{ "1005/1005.4692_arXiv.txt": { "abstract": "We use \\textit{Chandra} X-ray and \\textit{Spitzer} infrared observations to explore the AGN and starburst populations of XMMXCS J2215.9-1738 at $z=1.46$, one of the most distant spectroscopically confirmed galaxy clusters known. The high resolution X-ray imaging reveals that the cluster emission is contaminated by point sources that were not resolved in \\textit{XMM-Newton} observations of the system, and have the effect of hardening the spectrum, leading to the previously reported temperature for this system being overestimated. From a joint spectroscopic analysis of the \\textit{Chandra} and \\textit{XMM-Newton} data, the cluster is found to have temperature $T=4.1_{-0.9}^{+0.6}$~keV and luminosity $L_{\\rm X} = (2.92_{-0.35}^{+0.24}) \\times 10^{44}$~erg~s$^{-1}$, extrapolated to a radius of 2~Mpc. As a result of this revised analysis, the cluster is found to lie on the $\\sigma_v-T$ relation, but the cluster remains less luminous than would be expected from self-similar evolution of the local $L_{\\rm X}-T$ relation. Two of the newly discovered X-ray AGN are cluster members, while a third object, which is also a prominent 24~$\\micron$ source, is found to have properties consistent with it being a high redshift, highly obscured object in the background. We find a total of eight $>5\\sigma$ 24~$\\micron$ sources associated with cluster members (four spectroscopically confirmed, and four selected using photometric redshifts), and one additional 24~$\\micron$ source with two possible optical/near-infrared counterparts that may be associated with the cluster. Examining the IRAC colors of these sources, we find one object is likely to be an AGN. Assuming that the other 24~$\\micron$ sources are powered by star formation, their infrared luminosities imply star formation rates $\\sim100$~M$_\\sun$~yr$^{-1}$. We find that three of these sources are located at projected distances of $< 250$~kpc from the cluster center, suggesting that a large amount of star formation may be taking place in the cluster core, in contrast to clusters at low redshift. ", "introduction": "\\label{s_intro} The universe at high redshift is a much more active place than we see locally. Radio and far-infrared studies have shown that the comoving star formation rate density increases by a factor $\\approx 10$ from $z=0$ to $z \\approx 1$ \\citep[e.g.][]{Seymour_2008, Magnelli_2009}, and continues rising towards higher redshifts, peaking at $z \\approx 3$ \\citep{Bouwens_2009}. The fraction of galaxies with Active Galactic Nuclei (AGN) is also seen to be much higher when the universe was young: current observations suggest that the space density of the brightest AGNs peaks at $z \\approx 2-3$ \\citep{Assef_2010, Silverman_2005}. Observations in the local universe show that the stellar mass of galactic bulges is correlated with the mass of nuclear super-massive black holes \\citep[e.g.][]{Ferrarese_2000, Gebhardt_2000}, suggesting that AGN play a major role in the growth of stellar mass. Simulations suggest that interactions and mergers between gas-rich galaxies drive this process, by providing both fuel for AGNs and triggering starburst activity in galactic nuclei \\citep[e.g.][]{Granato_2004, Hopkins_2008}. Eventually, gas heating by the AGN (feedback) reaches such a level that star formation is shut down, leaving the stellar population of the host galaxy to evolve passively. The quenching of star formation in massive galaxies by AGN feedback is a crucial ingredient in the latest semi-analytic models of galaxy formation, bringing the models into much closer agreement with observations \\citep[e.g.][]{DeLucia_2006, Croton_2006, Bower_2006, Somerville_2008}. A connection between local environment, star formation and AGN is therefore expected in this theoretical picture. Intriguingly, recent studies of the dependence of star formation rate (SFR) on local galaxy density using large samples of field galaxies indicate that SFR increases with increasing local density at $z \\approx 1$ \\citep{Elbaz_2007, Cooper_2008} -- whereas the opposite is observed in the local universe \\citep[e.g.][]{Lewis_2002, Balogh_2004}. This situation is mirrored to some extent in galaxy clusters. Although these systems are dominated by quiescent, early type galaxies -- primarily `red-and-dead' ellipticals -- and continue to host substantial populations of these objects up to the highest redshifts observed \\citep[e.g.][]{Blakeslee_2003, Lidman_2008, Mei_2009, Hilton_2009}, an increasing fraction of star forming, late type galaxies are seen in clusters at higher redshifts \\citep[e.g][]{ButcherOemler_1984, vanDokkum_2000, Ellingson_2001, Smith_2005}. However, recent mid-infrared observations of clusters conducted using the \\textit{Spitzer Space Telescope} show that a significant amount of this increased star formation at high redshift is obscured by dust \\citep{Geach_2006, Marcillac_2007, Saintonge_2008, Haines_2009}, and therefore the amount of star formation in clusters has previously been underestimated. The AGN fraction in clusters is also seen to rise with redshift. \\citet*{Martini_2009} report an eight-fold increase in the number of AGNs in clusters at $z=1$ compared to in the local universe. Similarly, \\citet{Galametz_2009} find that IR-selected clusters in the IRAC Shallow Cluster Survey \\citep{Eisenhardt_2008} show a clear excess of AGNs at high redshift, with the AGN fraction increasing by a factor of three from $0.5 < z < 1.0$ to $1.0 < z < 1.5$. However, it is not yet clear if the AGN fraction is higher in clusters than in the field, nor if the AGN fraction is evolving more rapidly in dense environments. \\defcitealias{Stanford_2006}{S06} \\defcitealias{Hilton_2007}{H07} \\defcitealias{Hilton_2009}{H09} In this paper we use \\textit{Chandra} X-ray observations and mid-infrared observations taken using the \\textit{Spitzer Space Telescope} to characterize the AGN and dusty star forming population of the $z=1.46$ cluster XMMXCS J2215.9-1738 (hereafter J2215), which was discovered in the ongoing optical follow-up of the \\textit{XMM} Cluster Survey \\citep[XCS;][]{Romer_2001, Sahlen_2009}. J2215 is the second highest redshift X-ray selected cluster known, following the recent discovery of a $z=1.62$ cluster in the Subaru/\\textit{XMM-Newton} Deep Field (\\citealt*{Tanaka_2010}, \\citealt{Papovich_2010}). The discovery of J2215 and an analysis of its X-ray properties using \\textit{XMM-Newton} data was reported in \\citet[][hereafter S06]{Stanford_2006}. We presented a first study of the dynamical state of the cluster and measurement of its velocity dispersion using additional VLT and Keck spectroscopy in \\citet[][hereafter H07]{Hilton_2007}. Most recently, we have used \\textit{Hubble Space Telescope} (\\textit{HST}) imaging and ground-based near-infrared data obtained at the \\textit{Subaru} telescope to perform a detailed examination of the morphologies of the cluster galaxies and the red-sequence within $<0.75$~Mpc of the cluster core \\citep[][hereafter H09]{Hilton_2009}. The structure of this paper is as follows. In Section~\\ref{s_XRayAnalysis}, we report revised measurements of the cluster X-ray properties obtained through a joint spectroscopic analysis of the \\textit{Chandra} and \\textit{XMM-Newton} data, taking into account the effect of X-ray point sources that were unresolved in the \\textit{XMM-Newton} data used in \\citetalias{Stanford_2006}. In Section~\\ref{s_imaging} we briefly review the optical and ground based near-infrared data used in this paper, before describing new photometry obtained using the Infrared Array Camera \\citep[IRAC;][]{Fazio_2004} that extends wavelength coverage of the cluster into the rest frame near-infrared. We describe additional spectroscopic observations obtained at the Keck and Gemini observatories during 2008-2009 in Section~\\ref{s_OpticalSpectroscopy}, where we also present an updated measurement of the cluster velocity dispersion. Mid-infrared (24~$\\micron$) observations of the cluster obtained using the Multiband Imaging Photometer for \\textit{Spitzer} (MIPS) are described in Section~\\ref{s_MIPSAnalysis}. We characterize the properties of cluster 24~$\\micron$ and X-ray sources in terms of star formation or AGN activity in Section~\\ref{s_starFormation}. Finally, we discuss our findings in comparison with lower redshift studies of clusters in Section~\\ref{s_discussion}. We assume a concordance cosmology of $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, and $H_0=70$~km~s$^{-1}$~Mpc$^{-1}$ throughout, where $\\Omega_\\Lambda$ is the energy density associated with a cosmological constant. All magnitudes are on the AB system \\citep{Oke_1974}, unless otherwise stated. ", "conclusions": "\\label{s_conclusions} We have explored the AGN and star forming populations of the cluster J2215.9-1738 at $z=1.46$ using high resolution X-ray data from the \\textit{Chandra} satellite and infrared observations from the \\textit{Spitzer Space Telescope}. This is the first study of star formation as traced by mid-infrared observations in a cluster at $z \\approx 1.5$. We found: \\begin{enumerate} \\item{The cluster emission is contaminated by X-ray point sources, leading to the X-ray temperature being overestimated in the analysis presented in \\citetalias{Stanford_2006}; however, these point sources only contribute $\\approx 15$\\% to the total flux. Two of the X-ray point sources revealed by the \\textit{Chandra} observations are cluster members, while a third has properties consistent with it being a high redshift, highly obscured AGN in the background. The cluster has temperature $T=4.1_{-0.9}^{+0.6}$~keV and bolometric luminosity $L_{\\rm X} = (2.92_{-0.35}^{+0.24}) \\times 10^{44}$~erg~s$^{-1}$ (extrapolated to 2~ Mpc radius) from a joint analysis of the \\textit{XMM-Newton} and \\textit{Chandra} data. The cluster is less luminous than expected from self similar evolution of the local $L_{\\rm X}-T$ relation at the $\\approx 2\\sigma$ level.} \\item{The velocity dispersion of the cluster is measured to be $\\sigma_v = 720 \\pm 110$~km~s$^{-1}$, from 31 galaxies within $R_{200}$. There is no clear evidence that the cluster velocity distribution is composed of two kinematically distinct components, although the \\citet{Hartigan2_1985} dip test of unimodality still hints that the cluster is not completely relaxed. Following the revised X-ray analysis presented in this paper, the cluster is found to lie on the $\\sigma_v-T$ relation, contrary to the result reported in \\citetalias{Hilton_2007}.} \\item{Mid-infrared imaging reveals a total of eight $>5 \\sigma$ 24~$\\micron$ sources that are cluster members, selected using spectroscopic or photometric redshifts. In addition, there is a prominent 24~$\\micron$ source (J221559.68-173758.9) with two possible optical/near-infrared counterparts located $\\approx 17\\arcsec$ from the cluster center that may be associated with the cluster. One of the 24~$\\micron$ sources is found to have the infrared colors expected of an AGN, and has an elliptical morphology. The remaining objects are most likely powered by star formation, and if this is the case have SFRs $\\sim 100$~M$_\\sun$~yr$^{-1}$, adopting the \\citet{CharyElbaz_2001} spectral templates and assuming that the \\citet{Kennicutt_1998} law holds at this redshift.} \\item{The cluster member AGN identified in the X-ray and infrared observations are located within $\\lesssim 2 \\sigma$ of the cluster red-sequence, as are three (possibly four) 24~$\\micron$ sources assumed to be powered by star formation. Three of the 24~$\\micron$ sources are also found within a projected distance of $< 250$~kpc from the cluster center, suggesting that the core of J2215 may be host to galaxies with very high star formation rates, in contrast to clusters at lower redshift.} \\end{enumerate}" }, "1005/1005.1886_arXiv.txt": { "abstract": "{ Over the past decade, astronomers have been using an increasingly\\ larger number of web-based applications and archives to conduct their\\ research. However, despite the early success in creating links across\\ projects and data centers, the promise of a single integrated digital\\ library environment supporting e-science in astronomy has proven\\ elusive. While some of the issues hampering progress in this area are\\ of technical nature, others are rooted in existing policies which\\ should be re-analyzed if further rapid progress\\ is to be made in this area. This paper describes a proposal that the\\ NASA Astrophysics Data System project has put forth in order to\\ improve its role as one of the primary discovery portals for astronomers,\\ focusing on those aspects which could benefit from an increased level\\ of involvement from the community, namely the effort to expose\\ astronomy resources as linked data, and the harvesting of\\ observational metadata.\\ } \\FullConference{Accelerating the Rate of Astronomical Discovery, sps5\\\\ August 11-14, 2009\\\\ Rio de Janeiro, Brazil} \\begin{document} ", "introduction": "Astronomy is an observational science. Our theories of the behavior of the universe and the objects in it are inspired by observations of that behavior, and are supported (or rejected) by observations, as well. Within the past twenty years systems of dense, interlinked data have become common and expected, fundamentally changing the way people think and services operate. It is now expected, for example, that from a listing of a movie at our local theater one can find a synopsis, reviews, a cast list, a list of all the director's films, recommendations, and the option to purchase tickets, as well. Astronomy was one of the first disciplines to benefit from the early developments of web-based technologies enabling cross-linking of resources across archives (Accomazzi et al, 1994). Sixteen years ago, thanks to a collaboration between the NASA Astrophysics Data System (ADS) and the Centre de Donn{\\'e}es Astronomiques de Strasbourg (CDS), it became possible to go from a list of articles to the abstract of an article to a list of astronomical objects described in that article to a set of measurements on one of those objects. Thirteen years ago, again thanks to a collaboration between the ADS and several major data centers, including NED, HEASARC, MAST, ESO, and Chandra, it became possible to go from an article abstract to the actual observational data used to write the article, and then back to all publications describing each observation. These connections have enabled astronomers to use the search capabilities of any of the main archives to locate a dataset or publication of interest, and then follow the appropriate links to find related information provided by another archive. For instance, using ADS, one could search for papers on X-rays emission in Abell clusters and filter results to obtain just papers which have links to data products. While this is a very useful way to narrow bibliographic searches, selecting which of these papers has links to images or spectra or catalogs still requires a person to click through all the data links provided in the list of returned papers. Automating this activity is currently not practical since connections between one archive and another are purely defined as links between URLs, and lack semantic and contextual information between the resources they represent. In this paper we argue that a tighter integration of observational and bibliographic metadata will enable the creation of new connections between resources in astronomy and applications that will allow users to search, browse, and reason over them. The existence of these connections between and within papers and data products represents more that just a convenient feature to the end-user. The research process in science today involves the generation, retrieval, manipulation, analysis, and publication of digital artifacts which are typically stored in different archives on the web. In order to enable scientists to access these scholarly products, and to document and recreate the analysis that was carried out on them, a formal description of these resources and processes is necessary (Pepe et al., 2009). Establishing relationships and links among these resources provides a way to satisfy a number of different goals, among them: aggregation (linking together all artifacts used in a study, including data products, notes, draft papers, software tools); attribution (properly acknowledging prior work, tools being used, or datasets being analyzed); preservation (maintaining provenance information and reconstructing the workflow used in the research process); and discovery (following connections between resources may uncover previously unknown relationships). ", "conclusions": "In this coming era of data intensive science, it is increasingly important to be able to seamlessly move between scientific results, the data used to publish them, and the processes used to produce them. In addition, scientific research requires that we are able to establish the provenance of data sources and processes operating on this data, so that research may be repeated, variations on the research be carried out, and new research on existing datasets be enabled. While the Virtual Observatory has provided us with the standards and protocols needed to model and exchange astronomical datasets, the problem of information discovery in our data-centric world still looms large. The key to creating such a system is having access to the metadata describing the datasets at the appropriate granularity level so that the proper resources can be identified, described, and connections between them can be made. The technology which can be used to expose and link these metadata has now matured and is the foundation of the semantic web framework and in particular in the linked data model. These technologies promise to become widely-adopted standards enabling the creation and growth of a ``web of data'' in which machine-readable content parallels the human-readable content available today. We believe that the astronomical community would benefit from embracing this philosophy and adopting these technologies. In particular, three sets of metadata resources currently stored in our archives should be made available machine-readable format for harvesting and indexing: observing proposals, their attributes and links to observations; observations and their attributes; instruments, their attributes and capabilities. From a technical standpoint this metadata should be exposed as RDF and should make use of community-developed (lightweight) ontologies. This will allow the use of linked-data principles to properly connect these resources. However, even before these technical aspects are worked out, we believe that the availability of the metadata to organizations such as the ADS, the CDS, NED, and other IVOA members will enable the creation of knowledge bases and applications demonstrating the value of such an interlinked data system. Therefore we urge the community to support this effort by recommending that observatories and archives make available and permit the harvesting and indexing of metadata describing publicly accessible datasets, including observing proposals, observing logs, and FITS headers." }, "1005/1005.3551_arXiv.txt": { "abstract": "{ Massless interacting scalar fields in de Sitter space have long been known to experience large fluctuations over length scales larger than Hubble distances. A similar situation arises in condensed matter physics in the vicinity of a critical point, and in this better-understood situation these large fluctuations indicate the failure in this regime of mean-field methods. We argue that for non-Goldstone scalars in de Sitter space, these fluctuations can also be interpreted as signaling the complete breakdown of the semi-classical methods widely used throughout cosmology. By power-counting the infrared properties of Feynman graphs in de Sitter space we find that for a massive scalar interacting through a $\\lambda \\, \\phi^4$ interaction, control over the loop approximation is lost for masses smaller than $m \\simeq \\sqrt \\lambda \\, H/2\\pi$, where $H$ is the Hubble scale. We briefly discuss some potential implications for inflationary cosmology.} \\begin{document} ", "introduction": "Some aspects of quantum fields on de Sitter space remain controversial, long after their first investigation more than 30 years ago \\cite{dSorig}, but their potential relevance as an explanation of the detailed properties of the fluctuations observed by precision measurements \\cite{CMBmeas, CMBth1} in the Cosmic Microwave Background (CMB) radiation has stimulated a recent re-examination of the issues \\cite{Wbg, dSrecent1, dSrecent2, KLR,AKH, Dolgov, MS, petri, us, SZ} (see also \\cite{Review} and references therein) The main interpretational difficulties for de Sitter space arise for massless fields, or for those that are very light compared with the Hubble scale: $m \\ll H$. There are two related ingredients that complicate calculations with such fields: the presence of various types of infrared singularities, and the presence of large fluctuations over extra-Hubble distances. For example, if a massless scalar field is prepared in an initial state for which fluctuations are small, then quantities like $\\langle \\phi^2(t) \\rangle$ grow linearly with cosmic time, $t$. For massive scalars this growth eventually saturates at a $t$-independent value $\\langle \\phi^2 \\rangle \\propto H^4/m^2$, which is parametrically large if $m \\ll H$ \\cite{lint}. Since these fluctuations are uncorrelated, $\\langle \\phi(x) \\phi(y) \\rangle \\simeq \\langle \\phi(x) \\rangle \\langle \\phi(y) \\rangle$, on scales longer than $H^{-1}$ and since gradients quickly redshift away, one finds a picture in which the field takes an approximately constant value within each Hubble volume, with different volumes evolving independently of one another in an essentially random and uncorrelated way \\cite{superhubble}. Although this has long been recognized as the appropriate physical picture, essentially all of what we know about fields in de Sitter backgrounds is based on calculations performed within the semi-classical approximation. This approximation assumes that classical field theory captures the dominant physics, with small calculable corrections arising from quantum fluctuations. The size of these corrections is believed to be kept small because of their systematic dependence on small dimensionless quantities, like powers of coupling constants, $\\lambda/(4\\pi)^2$, and (because gravity is non-renormalizable) of small energy ratios, $E/M_p$ \\cite{GREFT}. In particular, such an approximation underpins calculations of inflaton fluctuations during inflation, and their implications for the properties of the CMB at recombination. In this paper we argue that for massless scalar fields in de Sitter backgrounds subject to non-derivative self-interactions --- like $V_{\\rm int} \\simeq \\lambda \\, \\phi^4$ --- the presence of large extra-Hubble fluctuations undermines the entire semi-classical approximation. This is because the semiclassical approximation is at heart a mean-field approximation, within which a quantum field is represented as a dominant classical configuration plus a small quantum fluctuation, $\\phi(x) = \\varphi(x) + \\delta\\phi(x)$. But this kind of a description fundamentally breaks down over distances larger than $H^{-1}$ due to the large fluctuations occurring on these scales. Notice that the breakdown of semiclassical methods we have in mind does not merely mean that the classical approximation is inadequate, with the situation being saved if we compute just a few more loops than usual. Instead, for massless scalars the danger is that the assumption that higher loops are suppressed by a small quantity that breaks down, meaning that {\\em all} orders in the semiclassical expansion have similar sizes. This makes semiclassical calculations inherently unreliable because the truncation of the loop expansion omits contributions that are as large as those that are kept. Quantum field theory is a cruel but fair master, so (as usual) the formalism contains within itself the news about the breakdown of semiclassical methods. The messenger is in this case the infrared divergences that commonly plague de Sitter calculations with massless scalars. These indicate a singular dependence on the mass in the more general case of a massive, but very light, scalar field. We argue that at least some of these divergences reflect the dominance of fluctuations over the contributions of the classical background, pointing to a fundamental breakdown of mean-field methods. To make our claim precise, we consider a scalar field in de Sitter space, that self-interacts through a quartic scalar potential: $V = \\frac12 \\, m_0^2 \\phi^2 + \\frac{1}{4!} \\, \\lambda \\, \\phi^4$. Perturbing in $\\lambda$ and using propagators for fields of mass $m_0$ shows that in the small-mass limit the usual loop-suppressing factor of $[\\lambda/(2\\pi)^2]$ for each loop in an $L$-loop graph is systematically modified by factors of $(H^2/m_0^2)$, indicating that higher loops are not suppressed once the scalar mass is sufficiently light. We explicitly identify contributions to $L$-loop correlators that are proportional to \\be \\left( \\frac{\\lambda H^4}{4 \\pi^2 m_0^4} \\right)^{L} \\,, \\ee which, taken at face value, would indicate that perturbation theory fails once $m_0^2 \\sim \\sqrt\\lambda \\, (H/2\\pi)^2$. However we argue that the breakdown of perturbative methods at this higher mass arises because the physical mass scale that cuts off IR effects is really $m^2 = m_0^2 + \\delta m^2$, with $\\delta m^2 \\simeq \\lambda H^4/m_0^2$, and so it is only the expansion $H^2/m^2 = (H^2/m_0^2) [1 - \\delta m^2/m_0^2 + \\cdots]$ that breaks down when $m_0^2 \\sim \\sqrt\\lambda \\, (H/2\\pi)^2$. This particular breakdown can be resummed: that is, it can be removed by reorganizing the perturbative expansion so that the unperturbed lagrangian involves the mass term $m^2 \\phi^2$ rather than $m_0^2 \\phi^2$. This particular reorganization does not also remove the potential breakdown of perturbation theory at $m_0^2 \\simeq \\lambda (H/2\\pi)^2$. Finite-temperature field theory provides a well-understood precedent for these conclusions. The small-momentum limit of the Bose-Einstein distribution function, $n_\\ssB(k) \\propto T/k$, implies that infrared divergences are stronger at finite temperature than they are at zero temperature. As a result, an $L$-loop Feynman graph for a self-interacting scalar field at finite temperature comes with the systematic factor $(\\lambda \\, T/4\\pi^2 m_0)^{L}$, again indicating a breakdown of the loop expansion if $m_0 \\lsim \\lambda \\,T/(2\\pi)^2$. In the thermal case it is also known that the loss of perturbative control is only partial when $m_0 = 0$ since then the complete thermal mass is $m^2 = \\delta m^2 \\propto \\lambda T^2$ and the IR divergences can be weakened by reorganizing perturbation theory so that the unperturbed fields have this mass. In this case the same power-counting leads to the loop factor $\\lambda T/m \\propto \\sqrt \\lambda$, leading to controlled (but non-analytic) results for small $\\lambda$. This same resummation fails, however, if $m_0^2 < 0$ is adjusted so that $m^2 = 0$, leading to a {\\em bona fide} breakdown of semiclassical methods, such as is known to occur in condensed matter systems in the vicinity of a critical point. The breakdown of the loop expansion in this case underlies the well-known failure of mean-field methods to describe critical exponents \\cite{RGCrit}. The physical origin of this breakdown is the dominance of large fluctuations near the critical point, similar to the fluctuations found in de Sitter space. The analogy between thermal field theory and de Sitter space is robust, apart from one potentially important difference: in de Sitter space the difference $\\delta m^2 = m^2 - m_0^2 \\propto \\lambda H^4/m_0^2$ itself depends singularly on $m_0^2$, unlike for thermal field theory where $\\delta m^2 \\propto \\lambda T^2$. Because of this difference the map between the two cases is not simply to think of the Hubble scale as a temperature. In particular, although $m^2$ can be adjusted to vanish in the thermal case by appropriately choosing $m_0^2$, it is not clear that this can be done in the de Sitter situation. We discuss in the conclusions the open problem of the extent to which these same regimes of resummation also apply in the de Sitter case (however, see \\cite{Akhmedov:2008pu}). Although gravity resembles a massless scalar in many ways, we emphasize that we do {\\em not} expect this same failure to arise for pure gravity in a de Sitter background. The difference arises because the gravitational self-interactions are largely derivative couplings, and so are typically not as divergent in the infrared. Since Goldstone scalars similarly couple derivatively, they also need not share the same mean-field breakdown. Massless spin-1 fields can couple without derivatives, and infrared effects are also known to ruin mean-field methods for a hot plasma of charged particles interacting through gauge interactions \\cite{HTL}. We leave it open whether a similar breakdown occurs on de Sitter space, but any such a failure would require the existence of very light charged degrees of freedom to survive the exponential de Sitter red-shifting. There is some work for SQED in de Sitter space that shows that the photon can indeed get a mass there (see \\cite{PW} for a review). The remainder of the paper is organized as follows. The next section, \\S2, briefly reviews the power-counting of infrared divergences for self-interacting scalar fields at finite temperature, to show how these track the breakdown of mean-field methods. \\S3 then provides a similar power-counting for self-interacting scalars in de Sitter space. Dimensional analysis is first used to argue that the worst divergences are logarithmic for any correlation functions. For a massive propagator, this translates into inverse powers of mass at each loop order. A class of graphs is then displayed that verifies this dependence through explicit calculation. Finally, \\S4\\ draws some preliminary implications for inflationary calculations, and summarizes our conclusions. ", "conclusions": "The body of this paper argues that an $L$-loop contribution to a correlation function for a scalar field in de Sitter spacetime with $\\lambda \\, \\phi^4$ self-interactions carries a systematic factor of $(\\lambda H^2/4\\pi^2 m_0^2)^{L}$, indicating a fundamental breakdown of semiclassical methods once $m_0 \\lsim \\sqrt\\lambda \\; H/2\\pi$. The origin of the breakdown of perturbative methods is the infrared-singular behavior of these graphs which arises due to the large extra-Hubble fluctuations experienced by very light scalars in de Sitter space. These fluctuations dominate the contributions to correlation functions, invalidating the semiclassical approximation which is at its heart a mean-field description. This story is qualitatively similar to what happens in finite temperature field theory, although the power of coupling constant that defines the boundary of the semiclassical region differs. In the finite temperature case, the variance of the field, $\\Expect{\\phi^2}$, goes like $T^ 2$ while in the de Sitter case we find $\\Expect{\\phi^2}\\sim H^4/m^2$. This difference means that the mass, $m_{\\rm dyn}$, which is comparable to the corrections to $m$ scales differently with $\\lambda$ in these two cases: for the thermal case $m_{\\rm dyn}^2 \\propto \\lambda T^2$, while for a de Sitter background $m_{\\rm dyn}^2 \\propto \\lambda^{1/2} H^2$. It also may mean that the structure of de Sitter space precludes access to the regime $m^2 \\simeq 0$ for any choice of $m_0^2$, unlike for finite-temperature systems, see Figure (\\ref{fig:range}). However, as is shown above, the two theories have a similar perturbative structure, whose relation is sketched in Figure (\\ref{fig:chart}). \\FIGURE[ht]{ \\epsfig{file=massplot.eps,angle=0,width=0.5\\hsize} \\caption{A plot of $m^2$ vs $m_0^2$. The horizontal band represents the regime $m^2 < \\cO(\\lambda H^2/4\\pi^2)$ while the curve is given by Eq. (\\ref{massshift}). It would be interesting to investigate the case of negative bare mass, which we do not display here, further} \\label{fig:range}} A natural question to ask about this perturbative breakdown at small $m^2$ is whether it can be resummed (like for $m_0^2 \\ge 0$ at finite temperature), or whether it reveals a complete breakdown of expansions based on powers of $\\lambda$ (like when $m_0^2 < 0$ is chosen so that $m^2 = 0$ at finite temperature). \\FIGURE[ht]{ \\epsfig{file=MassRangeFigSmall.eps,angle=0,width=0.6\\hsize} \\caption{A comparison of the behavior of loop corrections for a scalar field with a quartic interaction $\\lambda\\, \\phi^4$ in a thermal background or de Sitter space as a function of mass. The masses labelled on the chart are, from left to right, the mass below which perturbation theory breaks down, the dynamical mass generated in the event that the zero temperature or flat space mass was zero, and the maximum mass the field can have and still receive thermal/de Sitter fluctuations.} \\label{fig:chart} } Others have argued that there exist semiclassical methods that capture the leading infrared logs. For secular growing logs, one proposal solves the late-time physics using the classical equation of motion \\cite{MS}, in a similar way to the $\\delta N$ formalism. Another point of view uses a stochastic approach to inflation \\cite{superhubble}, which is argued to capture (and resum) the leading logs, by generating a dynamical mass \\cite{SY}. Although such a stochastic approach goes beyond mean field, it is not yet clear what combination of small parameters control the approximations made in its derivation. {}From the point of view of the arguments made here, derivative couplings are not as dangerous as are those of the scalar potential. This is because the momentum dependence of these couplings tends to ameliorate any IR divergence the graph would have otherwise had. This means that massless Goldstone bosons on de Sitter space would {\\em not} suffer from the same breakdown of perturbation theory as does the $\\phi^4$ model considered here. We would expect the self-interactions of massless gravitons to be similarly benign so long as these are derivative couplings, leading us to expect no perturbative breakdown for pure gravity on de Sitter space. This expectation seems to be borne out by ref.~\\cite{DB} and ref.~\\cite{GS} (the latter appeared just as our paper neared completion). These authors study one-loop infrared divergences for pure gravity and gravity coupled to scalars in de Sitter and slow-roll spacetimes, but find that IR divergences cancel in the absence of the self-interactions of the scalar potential (which appear in their calculations as slow-roll parameters). Based on the power-counting arguments presented here we expect this to continue to be true at higher loops, with the scalar self-interactions being the most dangerous in the infrared. It is interesting to consider in this light what implications our result might have for inflationary cosmologies. A complication in directly extracting these for simple single-field models is our neglect of classical evolution of the background scalar field and metric, due to our use of a simple de Sitter background. Because a homogeneous evolving scalar field can be used to define a notion of cosmic time, many of its effects can be gauged away. We therefore expect a naive application of the above arguments to simple models are likely to cancel from gauge invariant quantities like the curvature fluctuations of physical interest for cosmology. Nevertheless, it may be possible to have important infrared effects appear in curvature correlations in multi-field models, particularly those involving nontrivial post-inflationary dynamics (such as curvaton models). We leave for future work the detailed question of whether and how the infrared effects we find above `propagate' into late-time curvature perturbations. In the remainder of this section we put these issues aside, and ask what the condition $m^2 > \\lambda H^2/4\\pi^2$ implies for the parameters of a single scalar field described by a quartic potential \\be V = V_0 + \\frac12 \\, m^2 \\phi^2 + \\frac{1}{4!} \\, \\lambda\\, \\phi^4\\,. \\ee For this model we ask how the condition $M^2 \\gg \\lambda (H/2\\pi)^2$, where $M^2 = V'' = m^2 + \\frac12 \\lambda \\, \\phi^2$, compares to other conditions to which inflationary models must be subject. To see what this implies for potential inflationary applications, consider two extreme cases: $(i)$ $\\phi$ so large that $V \\simeq \\frac{1}{24} \\, \\lambda \\, \\phi^4$; and $(ii)$ $\\phi$ small enough that $V \\simeq V_0$. (This last example can be regarded as part of a model of hybrid inflation \\cite{hybrid}, with inflation ending as another field starts to roll as its $\\phi$-dependent squared-mass goes negative.) \\subsubsection*{Large-field inflation} For the large-$\\phi$ regime we have $V \\simeq \\frac{1}{24} \\, \\lambda \\, \\phi^4$, and so $H^2 = V/3M_p^2 \\simeq \\lambda \\, \\phi^4/72 M_p^2$. The slow-roll parameters are \\be \\varepsilon := \\frac12 \\left( \\frac{M_p V'}{V} \\right)^2 \\simeq 4 \\left( \\frac{M_p}{\\phi} \\right)^2 \\quad \\hbox{and} \\quad \\eta := \\frac{M_p^2 V''}{V} \\simeq 12 \\left( \\frac{M_p}{\\phi} \\right)^2 \\,, \\ee so the edge of the inflationary regime is $\\phi_\\SR /M_p \\simeq \\O(1)$. For $\\phi$ larger than this classical evolution satisfies \\be \\dot\\phi \\simeq \\frac{V'}{3H} \\simeq \\frac{\\frac16 \\, \\lambda \\, \\phi^3}{\\sqrt\\lambda \\; \\phi^2/2\\sqrt2 \\; M_p} \\simeq \\frac{\\sqrt{2}\\,}{3} \\, \\sqrt\\lambda \\; M_p \\phi \\,. \\ee Once $\\dot\\phi$ becomes smaller than $H^2$ fluctuations dominate classical evolution and inflation becomes eternal, which in this case occurs when $\\frac{\\sqrt{2}\\,}{3} \\sqrt\\lambda \\; M_p \\phi < \\lambda \\, \\phi^4/72 M_p^2$, or $\\phi^3 > \\phi_\\EI^3 \\simeq 24 \\sqrt2 \\; M_p^3/\\sqrt\\lambda$. How do the boundaries of the semiclassical approximation compare to these? There are two criteria to be satisfied. First, control over the low-energy approximation that underlies the gravitational loop expansion requires $V \\ll M_p^4$, or $\\phi^4 \\ll \\phi_\\HE^4 \\simeq 24 M_p^4/\\lambda$. We have seen in previous sections that the $\\lambda$ loop expansion fails unless $M^2 \\gg \\lambda \\, H^2/4 \\pi^2$ or $\\frac12 \\, \\lambda \\, \\phi^2 \\gg \\lambda^2 \\phi^4/288 \\pi^2 M_p^2$, and so $\\phi^2 \\ll \\phi^2_\\BD \\simeq (12 \\pi M_p)^2/\\lambda$. Since $\\phi_\\SR/M_p \\simeq \\O(1)$, $\\phi_\\EI/M_p \\simeq \\O(\\lambda^{-1/6})$, $\\phi_\\HE /M_p \\simeq \\O( \\lambda^{-1/4})$ and $\\phi_\\BD /M_p \\simeq \\O( \\lambda^{-1/2})$ we have \\be \\phi_\\SR \\ll \\phi_\\EI \\ll \\phi_\\HE \\ll \\phi_\\BD \\,, \\ee and so the condition $M^2 > \\lambda (H/2\\pi)^2$ is parametrically weaker than the condition $M_p^4 > V$. The fact that eternal inflation can occur before a total breakdown of perturbation theory is as expected. \\subsubsection*{Hybrid inflation} Consider next a hybrid model \\cite{hybrid}, involving two scalar fields, $\\phi$ and $\\chi$, interacting through the potential \\be U(\\phi,\\chi) = \\frac{1}{4} \\,\\zeta \\,( \\chi^2 - v^2 )^2 + \\frac{g^2}{2} \\, \\chi^2 \\phi^2 + \\frac12 \\, m^2 \\phi^2 + \\frac{1}{4!} \\, \\lambda\\, \\phi^4\\;. \\ee In this model the fields start in the trough defined by $\\chi = 0$, with $\\phi$ large and rolling towards smaller values subject to the effective potential \\be V = V_0 + \\frac12 \\, m^2 \\phi^2 + \\frac{1}{4!} \\, \\lambda\\, \\phi^4\\,, \\ee with $V_0 = \\frac{1}{4!} \\, \\zeta \\, v^4$. This roll continues until \\be \\phi = \\phi_\\SR = \\frac{\\sqrt\\zeta \\, v}{g} \\,, \\ee at which point the $\\chi$ mass, $\\mu^2 = - \\zeta \\, v^2 + g^2 \\phi^2$, becomes negative, allowing $\\chi$ to evolve quickly towards the absolute minimum at $\\chi = v$ and $\\phi = 0$. The dynamics of the inflaton in this model is governed by the same potential considered earlier, but our interest now is in the small-field regime, for which $m^2 \\gg \\frac12 \\, \\lambda \\, \\phi^2$ and $V \\simeq V_0$ (and so $H^2 \\simeq V_0/3M_p^2$). We assume parameters are chosen to keep $\\phi_\\SR$ small enough to ensure $\\chi$ remains zero well into this regime. In this case the slow-roll parameters are \\be \\varepsilon \\simeq \\frac12 \\left( \\frac{m^2 M_p \\phi}{V_0} \\right)^2 \\quad \\hbox{and} \\quad \\eta \\simeq \\frac{m^2 M_p^2}{V_0} \\,, \\ee so $2 \\varepsilon \\simeq \\eta^2 (\\phi/M_p)^2$. Clearly inflation only requires $m^2 M_p^2 /V_0 \\ll 1$, since the condition $\\frac12 \\, \\lambda \\phi^2 \\ll m^2$ automatically ensures $\\phi \\ll M_p/\\eta$. Inflation ends once $\\phi$ reaches $\\phi_\\SR = \\sqrt\\zeta \\; v/g$. In this case the validity of the $\\lambda$ loop expansion requires $m^2 \\gg \\lambda \\, H^2/4 \\pi^2 \\simeq \\lambda V_0/12 \\pi^2 M_p^2$, or \\be \\eta \\simeq \\frac{m^2 M_p^2}{V_0} \\gg \\frac{\\lambda}{12 \\pi^2} \\,. \\ee Notice that this lower bound to the slow-roll parameters is not simply a naturality condition, corresponding to a regime for which small loop corrections dominate the smaller classical potential. Rather, in this regime the problem is not fixed simply by including one- or two-loop corrections to $V$. Instead it is the entire loop expansion itself that fails. It would be of great interest to see whether the semiclassical criterion plays any role in more general inflationary contexts." }, "1005/1005.2411_arXiv.txt": { "abstract": "{We have determined the luminosity function of 250$\\,\\mu$m-selected galaxies detected in the $\\sim$14\\,deg$^2$ science demonstration region of the Herschel-ATLAS project out to a redshift of $z=0.5$. Our findings very clearly show that the luminosity function evolves steadily out to this redshift. By selecting a sub-group of sources within a fixed luminosity interval where incompleteness effects are minimal, we have measured a smooth increase in the comoving 250$\\,\\mu$m luminosity density out to $z=0.2$ where it is $3.6^{+1.4}_{-0.9}$ times higher than the local value.} ", "introduction": "Measurement of the galaxy luminosity function (LF) constitutes one of the most fundamental statistical constraints that can be placed on models of galaxy formation, and hence the build up of large scale structure, in the universe. Since half of the energy ever emitted by galaxies has been absorbed by dust and re-radiated at far-infrared and sub-millimetre (submm) wavelengths (Fixsen et al. \\cite{fixsen98}) and, because knowledge of the statistical properties of submm sources is relatively sparse, determination of the submm LF provides a crucial missing piece in a fully comprehensive model of galaxy evolution. Following their detection in the first deep submm and mm surveys (e.g., Smail et al. \\cite{smail97}; Hughes et al. \\cite{hughes98}; Eales et al. \\cite{eales99}; Bertoldi et al. \\cite{bertoldi00}), much has been learned about the dusty high-redshift sources selected at such wavelengths. Although many studies have argued that these sources are likely ancestors of local ellipticals (e.g., Scott et al. \\cite{scott02}; Dunne et al. \\cite{dunne03}) little progress in verifying this assertion has been made since. The main reason for this is the preponderance of high redshift submm/mm-selected sources, owing to the strong negative k-correction and small survey areas. The resulting low numbers of sources at redshifts $z<1$ has therefore precluded evolutionary studies over the last $\\sim$60\\% of the Universe's history. In particular, despite the local submm LF being first determined a decade ago (Dunne et al. \\cite{dunne00}), little has been added to our comprehension of how the LF has evolved over the last $\\sim 7$\\,Gyr, until very recently. Observations conducted using the Balloon-borne Large Aperture Submm Telescope (BLAST; Devlin et al. \\cite{devlin09}) have made significant improvements with much enhanced sensitivity to $z<1$ sources. As a result, direct estimates of the LF at 250, 350 and 500$\\,\\mu$m were made by Eales et al. (\\cite{eales09}) who detected strong evolution, particularly among the higher luminosity systems, from $z=1$ to the present day. However, the accuracy of these findings is limited by the small number of sources used ($\\sim 50$ at $z<0.5$) and source confusion due to the angular resolution of BLAST. In this letter, we present our measurement of the LF of 250$\\,\\mu$m-selected galaxies detected by the Herschel Space Observatory ({\\sl Herschel}; Pilbratt et al. \\cite{pilbratt10}) over a $\\sim 14\\,$deg$^2$ region acquired as part of the science demonstration observations of the {\\sl Herschel}-Astrophysical Terahertz Large Area Survey (H-ATLAS; Eales et al. \\cite{eales10}). These data offer a significant improvement over the BLAST data in terms of their increased sensitivity, higher angular resolution and greater areal coverage, resulting in $\\sim 20$ times the number of sources with which to compute the LF. Throughout this letter, the following cosmological parameters have been assumed; ${\\rm H}_0=71\\,{\\rm km\\,s}^{-1}\\,{\\rm Mpc}^{-1}$, $\\Omega_{\\rm m}=0.27$, $\\Omega_{\\Lambda}=0.73$. ", "conclusions": "One of the key goals of H-ATLAS will be to understand the nature of the evolution detected in this letter. In turn, we aim to improve our understanding of the evolutionary link between high redshift and local submm systems. This letter has only considered sources selected at 250$\\,\\mu$m, merely one of the five wavebands on offer from H-ATLAS. Furthermore, the 14\\,deg$^2$ of survey data analysed in this work represent only 2.5\\% of the final, proposed H-ATLAS survey area. A repeat of the analysis presented here with the final survey data, would therefore result in the quoted uncertainties falling by at least a factor of five. In light of these considerations, it is clear that our results offer only a small glimpse of the anticipated wealth of science that H-ATLAS has to offer." }, "1005/1005.0622_arXiv.txt": { "abstract": "In this article, we test the hypothesis that Cepheids have infrared excesses due to mass loss. We fit a model using the mass-loss rate and the stellar radius as free parameters to optical observations from the OGLE-III survey and infrared observations from the 2MASS and SAGE data sets. The sample of Cepheids have predicted minimum mass-loss rates ranging from zero to $10^{-8}M_\\odot$ $yr^{-1}$, where the rates depend on the chosen dust properties. We use the predicted radii to compute the Period-Radius relation for LMC Cepheids, and to estimate the uncertainty caused by the presence of infrared excess for determining angular diameters with the infrared surface brightness technique. Finally, we calculate the linear and non-linear Period-Luminosity (P-L) relations for the LMC Cepheids at $VIJHK$ + IRAC wavelengths and we find that the P-L relations are consistent with being non-linear at infrared wavelengths, contrary to previous results. ", "introduction": "Classical Cepheids are powerful standard candles because they follow the Period-Luminosity (P-L) relation or Leavitt Law \\citep{Leavitt1908}. The P-L relation has been used as a tool for Galactic \\citep{Feast1997b}, extragalactic \\citep[and references therein]{Gieren2009} and cosmological studies \\citep{Freedman2001, Sandage1998}. The LMC Cepheids have been central for deriving these relations, initially using optical data \\citep{Udalski1999a,Sandage2004,Kanbur2003, Kanbur2006, Fouque2007}, then near-infrared observations \\citep{Groenewegen2000, Persson2004, Fouque2007}, and recently using infrared observations \\citep{Freedman2008, Ngeow2008a}. The infrared P-L relations were determined by matching sources from the SAGE (Surveying the Agents of a Galaxy's Evolution) archival data \\citep{Meixner2006} with the known Cepheids. \\cite{Ngeow2008b} computed the infrared P-L relations for a larger sample of 1848 Cepheids from the OGLE-III Database \\citep{Soszynski2008} by matching those sources with the two published epochs of SAGE data, and also averaging the two epochs of infrared fluxes to bring the fluxes closer to the mean flux. They tested the P-L relations for non-linearity, and found that the IR P-L relations are linear, while the relations at wavelengths shorter than $K$-band are non-linear. However, there are a number of Cepheids with infrared fluxes that are $>3$ standard deviations brighter than predicted by the P-L relations. Similar results were found using AKARI N-band data \\citep{Ngeow2010}. It is important to understand the nature of this excess infrared brightness because the intrinsic infrared flux of Cepheids is less metallicity dependent than optical P-L relations \\citep{Freedman2008}, which makes the IR P-L relations powerful tools for extragalactic studies. The James Webb Space Telescope will be able to observe Cepheids in distant galaxies, making it possible to use the IR P-L relations to determine distances, but the unknown source of infrared excess will increase the uncertainty of the results. One hypothesis for the source of the infrared excess is the existence of a circumstellar envelope (CSE) of dust around the LMC Cepheids, analogous to the CSEs observed around Galactic Cepheids \\citep{Kervella2006, Merand2006, Merand2007}. Multi-wavelength observations of RS Puppis and $l$ Car confirm the presence of CSEs around these Cepheids \\citep{Kervella2009}. There is more evidence of infrared excess in Galactic Cepheids using IRAS observations \\citep{McAlary1986, Deasy1988} and Spitzer observations \\citep{Marengo2009}. On the other hand, \\cite{Marengo2009b} argued that there is no evidence from IR excess from warm dust based on Spitzer observations and instead was due to carbon monoxide emission. The Cepheid RS Pup is also associated with a nebula \\citep{Havlen1972, Kervella2008}. \\cite{Kervella2006} speculated that CSEs surrounding Galactic Cepheids were caused by mass loss that can leads to an infrared excess. \\cite{Neilson2008a} proposed a driving mechanism for this wind that is a combination of radiative acceleration, pulsation and shocks in the atmosphere of the Cepheid. This proposed driving mechanism was tested by developing an analytic mass-loss model that was used to calculate rates of mass loss for a sample of Galactic Cepheids, with values between $10^{-10}$ to $10^{-7}M_\\odot/yr$ being found. This model predicts the fluxes of CSEs around the Cepheids that agree well with those observed. \\cite{Neilson2008b} applied this mass-loss model to theoretical models of Galactic, LMC, and Small Magellanic Cloud Cepheids and found that Cepheids in all three galaxies have significant mass-loss rates and infrared excesses. There is other evidence for Cepheid mass loss in addition to infrared excess; \\cite{Nardetto2008b} observed H$\\alpha$ line profiles of Galactic Cepheids with the purpose of detecting a hydrogen circumstellar medium. They found that the H$\\alpha$ line profiles are asymmetric and blue-shifted in long period ($P>10$ day) Cepheids, possibly caused by stellar wind. While it is important to characterize infrared excess of Cepheids in order to use them for various applications, it is just as important to understand how mass loss affects the structure and evolution of Cepheids. Mass loss is a potential solution to the mass discrepancy of Cepheids, which is the difference between Cepheid masses estimated using stellar evolution calculations and estimates using stellar pulsation calculations \\citep{Cox1980}. Currently, stellar pulsation calculations predict masses about $10$-$20\\%$ smaller than stellar evolution models \\citep{Keller2006, Keller2008}. Measurements of dynamic masses of Cepheids in binary systems tend to agree with pulsation calculations \\citep{Evans2006, Evans2008}, suggesting the discrepancy is in the stellar evolution modeling. If Cepheids lose an average of $10^{-8}$ to $10^{-7}M_\\odot/yr$, then mass loss could solve the mass discrepancy. Mass loss is hypothesized to create optically thin circumstellar dust shells that cause infrared excess in Cepheids but do not affect the visual extinction of the Cepheids. This excess should be seen in infrared observations of the LMC from the SAGE survey \\citep{Meixner2006} for which there are currently two epochs of observations. In fact, infrared excess has been detected in a number of evolved AGB stars in the LMC \\citep{Vijh2008, Groenewegen2009}. In our earlier paper testing mass loss in LMC Cepheids, we found that mass loss may be important, with about $10\\%$ of OGLE-II Cepheids showing statistically significant evidence for dust shells. However, the uncertainty of mean infrared flux of the Cepheids is a limiting factor. The predicted gas mass-loss rates of the LMC Cepheids in that sample range from $10^{-11}$ to $10^{-8}M_\\odot/yr$ assuming a dust-to-gas ratio of $1/250$. Mass-loss rates in this range affect both the zero-point and the slope of the infrared P-L relations. When the infrared excess was removed from the OGLE-II sample, \\cite{Neilson2008c} found that the scatter of the data was reduced. We also found that the $3.6$ to $5.8$ $ \\mu m$ relations are consistent with being non-linear while the $8.0$ $\\mu m$ relation are linear, differing from the observed relations \\citep{Ngeow2008a}. The infrared excess caused by mass loss appeared to have a greater effect at shorter periods than at longer periods. However, the infrared stellar luminosity is an increasing function of period, making the infrared excess is less noticeable for long-period Cepheids, even for larger mass-loss rates ($\\approx 10^{-8}~M_\\odot/yr$). These results suggest that the near-infrared fluxes, $JHK$, are potentially affected by mass loss. This would mean distance estimates using the Infrared Surface Brightness (IRSB) technique have uncertainty due to mass loss and infrared excess. The infrared surface brightness technique uses an empirical fit of the dependence of surface brightness on $(V-K)_0$ to determine the mean value and amplitude of a Cepheid's angular diameter \\citep{Gieren1999}. This angular diameter information is then converted to a radius and a distance by using radial velocity observations to compute the amplitude of the radius variation. This technique has been shown to be robust, but infrared excess from mass loss suggests that the color $(V-K)_0$ may be overestimated. The objective of this paper is to analyze fundamental-mode Cepheids in the LMC that have stellar fluxes from the OGLE-III survey and infrared data from the 2MASS and from two epochs of the SAGE surveys, and test if the Cepheids have infrared excess. We determine mass-loss rates from infrared excesses determined from a correlation of these data sets by assuming a mass-loss model. In the next section we describe the method for testing the model. In section 3 we fit the model assuming zero mass loss, and the results of fitting the mass-loss model are given in section 4. In section 5, we test how the removal of the $5.8$ and $8.0$ $\\mu m$ data affect the mass-loss model as well as how the unknown pulsation phase of the IR fluxes affects the predicted mass-loss rates. The statistical comparison of the two models and the uncertainty of the fits due to the unknown pulsation amplitudes of the IR fluxes is given in section 5. Using the results of the models, we calculate the Period-Radius relation (section 6), and test the uncertainty of the IRSB technique due to mass loss (section 7). In Section 8 we compute linear IR P-L relations from the data and compare the predictions with the observed results of \\cite{Ngeow2008b}, \\cite{Freedman2008}, and \\cite{Madore2008} in Section 9. In Section 10, we test the P-L relations for non-linearity. ", "conclusions": "" }, "1005/1005.0552_arXiv.txt": { "abstract": "The Cherenkov radio pulse emitted by hadronic showers of energies in the EeV range in ice is calculated for the first time using full three dimensional simulations of both shower development and the coherent radio pulse emitted as the excess charge develops in the shower. A Monte Carlo, \\zhaires, has been developed for this purpose combining the high energy hadronic interaction capabilities of AIRES, and the dense media propagation capabilities of TIERRAS, with the precise low energy tracking and specific algorithms developed to calculate the radio emission in ZHS. A thinning technique is implemented to allow the simulation of radio pulses induced by showers up to 10 EeV in ice. The code is validated comparing the results for electromagnetic and hadronic showers to those obtained with GEANT4 and ZHS codes. The contribution to the pulse of other shower particles in addition to electrons and positrons, mainly protons, pions and muons, is found to be below 3 $\\%$ for 10 PeV and above proton induced showers. The characteristics of hadronic showers and the corresponding Cherenkov frequency spectra are compared with those from purely electromagnetic showers. The dependence of the spectra on shower energy and high-energy hadronic model is addressed and parameterizations for the radio emission in hadronic showers in ice are given for practical applications. ", "introduction": "The observation of neutrinos of EeV-scale energies is one of the main priorities in Astroparticle Physics. The detection of neutrinos will open a new window to observe parts of the Universe shielded by large depths of matter, unaccessible to conventional astronomy. The measurement of diffuse neutrino fluxes will provide further clues to the identification of the sources of ultra-high energy cosmic ray production, their composition and their production mechanisms~\\cite{becker08}. In addition such detections could have important implications for fundamental particle physics. A very promising and cost-effective method to detect high-energy neutrino interactions was first proposed by G.~A.~Askar'yan in the 1960's \\cite{Askaryan62}. The idea is to detect the Cherenkov radiation at radio wavelengths generated by the excess number of electrons in the cascade of particles resulting from a high-energy particle interaction in a dense medium transparent to radio waves. The development of this excess charge, due to the interactions with matter electrons, is often referred to as the Askar'yan effect. The effect has been experimentally confirmed in accelerator experiments at SLAC in media such as sand \\cite{Saltzberg_SLAC_sand,Miocinovic_SLAC_sand}, rock salt \\cite{Gorham_SLAC_salt} and ice \\cite{Gorham_SLAC_ice}, with results in good agreement with theoretical calculations \\cite{alvz06}. At radio frequencies (MHz-GHz), the emission is coherent and the radiated power scales with the square of the primary particle energy. This makes this method very promising for the detection of neutrinos and cosmic rays of the highest energies (EeV) \\cite{zheleznykh,ZHS91,frichter96}. Several experiments have already exploited this technique searching for ultra-high energy neutrinos, but no positive detection has been reported so far. These include experiments using the ice cap at the South Pole, such as the ANITA balloon experiment \\cite{ANITAlite,ANITAlong,ANITA_2009_limits} and the RICE array of antennas buried under the Antarctic ice \\cite{RICE03,RICElimits}. Other arrays are beginning to be constructed, such as the Askar'yan Radio Array ARA \\cite{ARA} and ARIANNA \\cite{ARIANNA}. There are also a number of projects using the Moon as target for neutrino interactions, along with radio telescopes on Earth as radiation detectors, such as the pioneering Parkes \\cite{Parkes96,Parkes07}, GLUE \\cite{GLUElimits}, Kalyazin \\cite{Kalyazin}, NuMoon \\cite{NuMoon}, LUNASKA \\cite{LUNASKA} and RESUN \\cite{RESUN}. The interpretation of data from these experiments requires a detailed knowledge of the magnitude, angular distribution and frequency-dependence of the emitted Cherenkov radiation, which can then be related back to the properties of the induced cascade. This calls for accurate simulations of the Fourier-spectrum of coherent Cherenkov radiation from EeV showers in different dense media. In the past, full simulations of electromagnetic (EM) showers in ice were done up to PeV energies, using the well-known and well-tested ZHS code \\cite{ZHS91,ZHS92} and different versions of the GEANT code \\cite{razzaque04,almvz03}. Also, full simulations of hadronic and neutrino-induced showers in ice were carried out up to 100 TeV with GEANT \\cite{McKay_radio}. Hybrid Monte Carlo \\cite{alvarez_hybrid} and thinning techniques \\cite{Hillas} were also developed to simulate electromagnetic \\cite{alz97,aljpz09,acorne07}, hadronic \\cite{alz98,acorne07} and neutrino-induced showers \\cite{acorne07,alvz99} above PeV energies (mainly in ice and water). Above these energies the LPM effect \\cite{LPM,Stanev_LPM} starts to be important in the media under experimental consideration \\cite{Stanev_LPM,RalstonLPM,alz97}. Semi-analytical calculations of the radio-emission have also been performed \\cite{alvz00,buniy02}. Very recently, simulations of not only the Fourier-components of the spectrum but also of the radio-pulse in the time-domain emitted in electromagnetic showers, have been performed with the ZHS code\\cite{alrwz10}. Modeling hadronic showers up to EeV energies is of utmost importance for neutrino detection, since they are induced by all neutrino flavors in neutral current (NC) interactions, as well as in the hadronic vertex of charged current (CC) interactions of muon and tau neutrinos. Although at EeV energies only $\\sim 20\\%$ of the neutrino energy is on average carried by the hadronic shower, these showers are known to be less affected by the LPM effect \\cite{alz98}, and their Cherenkov emission does not suffer from the shrinking of the Cherenkov cone, which would otherwise reduce the solid angle for observation. On the other hand, mixed showers induced in CC electron neutrino interactions are composed of a purely electromagnetic shower, produced by an energetic electron carrying on average $\\sim 80\\%$ of the energy of the neutrino, and a hadronic shower, initiated by the debris of the interacting nucleon. The advantage for neutrino detection of these types of interaction is that all the neutrino energy is channeled into the resulting shower. However, the observation of the electromagnetic shower is expected to be very difficult, since the LPM effect shrinks the angular distribution of the electric field, reducing dramatically the available solid angle for detection. For these reasons, experiments aiming at detecting neutrinos using the radio technique gain most of their acceptance from neutrino-induced hadronic showers \\cite{RICElimits,ANITAlong,alvarez_icrc09}. Clearly, accurate simulations of hadronic showers with energies above which the LPM effect becomes effective in different dense media are needed. In this paper we present the first steps in that direction. We have used the well-known AIRES code \\cite{aires} in combination with the TIERRAS package \\cite{TIERRAS} to simulate proton-induced showers in dense media, such as ice. We have implemented the algorithms to calculate the Fourier components of the electric field produced by charged particle tracks in the shower, developed by Zas-Halzen-Stanev in the well-known and well-tested ZHS code \\cite{ZHS91,ZHS92}. The result is a flexible and powerful code named \\zhaires, which allows the simulation of electromagnetic, hadronic and neutrino-induced showers in a variety of media, along with their associated coherent radio emission due to the Askar'yan effect. It is important to stress that in \\cite{alz97,alz98} the calculations of the emission properties of showers in ice up to EeV energies were only done in an approximate way. They were based on the Fourier-transform of simulated longitudinal and lateral shower profiles. With \\ZHAireS we can obtain the radio-pulse features in a consistent manner within a well-tested simulation, computing the electric field emitted by charged particles in the shower. In Ref.\\cite{McKay_radio}, GEANT simulations were used to calculate the radio-pulse properties in the same way as the \\ZHAireS code. However these calculations were limited to energies below 100 TeV due to GEANT limitations, while with \\ZHAireS we can simulate showers at EeV energies and above. Also, and for the first time, the contribution to the radio pulse of charged pions, muons and protons is accounted for and quantified, and parameterizations of the frequency spectrum of the radio pulse due to Askar'yan effect in hadronic showers are presented. Previous parameterizations of hadronic showers exist in the literature~\\cite{alz98}, but these are based on less accurate simulations than those presented in this work. This paper is structured as follows. In Section \\ref{newcode} we describe the new \\ZHAireS code and compare its performance and results for electromagnetic showers with those of the well-tested codes GEANT 4 \\cite{GEANT4} and ZHS \\cite{ZHS92,aljpz09}. We also explore thinning techniques in the \\ZHAireS code, essential for the simulation of EeV showers. Section \\ref{hadronic} is devoted to the simulation of hadronic showers up to EeV energies, emphasizing their differences with respect to purely electromagnetic showers. Section \\ref{conclusions} summarizes and concludes the paper. In Appendix A, we also give parameterizations of the intensity of the Fourier-spectrum of the Cherenkov electric field emitted in hadronic showers, as a function of shower energy and observation angle. In this paper we concentrate on hadronic showers in ice. We defer to future papers the study of coherent radio emission from EM and neutrino-induced showers. However we show that our results for proton showers can be applied to approximately model the Cherenkov emission from hadronic showers induced in high-energy neutrino interactions. The code can also be applied to the simulation of radio emission in extensive air showers~\\cite{zhairesarena2010,zhairesair2011}, despite the different emission mechanism \\cite{allan71,ARENA08}. ", "conclusions": "\\label{conclusions} We have presented \\zhaires, a Monte Carlo code that merges the high energy hadronic interaction and tracking capabilities of AIRES\\cite{aires}, and the dense media propagation capabilities of TIERRAS \\cite{TIERRAS} with the precise low energy $e^\\pm$ tracking and radiation calculation capabilities of ZHS\\cite{ZHS91,ZHS92}. This combination allows a precise full simulation of radio emission of electromagnetic, and for the first time also hadronic showers up to EeV energies. \\ZHAireS has been used in conjuction with the TIERRAS package \\cite{TIERRAS} to simulate showers in dense media, obtaining the radiation emitted due to the Askar'yan effect. We have compared \\ZHAireS results for electromagnetic showers in ice against ZHS up to 10 EeV, obtaining a good agreement between them. In the case of hadronic showers, we have compared the results from \\ZHAireS and GEANT4 up to 90 TeV \\cite{McKay_radio}, the highest energy for full simulations of hadronic showers with GEANT in the literature, also with very good quantitative agreement. Unlike GEANT4, \\ZHAireS is capable of simulating the emission from hadronic showers up to EeV energies. By comparing the results of thinned \\ZHAireS simulations with un-thinned ones up to 1 PeV, we obtained thinning parameters which give a good compromise between accuracy and CPU time. We then tested the applicability of these parameters at higher energies, by comparing our results against ZHS, which has its own, well established \\cite{aljpz09}, thinning algorithms and parameters. We confirmed the expectation that the EM energy ratio of hadronic showers keeps approaching that of EM showers at the highest energies, increasing the magnitude of the electric field (normalized by $E_0$) as energy grows. We also showed that, in the case of UHE purely EM showers, the EM energy ratio deviates from a linear dependence on energy, due to photohadronic interactions. By comparing EM and hadronic showers up to 10 EeV we have found that, at the Cherenkov angle, the cut-off frequency tends to be smaller in hadronic showers and increases slowly as the energy rises because the lateral distribution becomes narrower. For angles away from the Cherenkov angle, as expected, the cut-off frequency decreases with energy, due to the logarithmic growth of the longitudinal profile. In contrast, away from the Cherenkov angle, the cut-off frequency for EM showers decreases rapidly above PeV energies due to the LPM effect, which is much more pronounced in EM showers than in hadronic ones. We also found that the fluctuations in hadronic showers are almost independent of energy, while in EM showers they grow rapidly, as the LPM effect becomes important. We have also analyzed the influence of charged pions, muons and protons on the radio emission of showers. We found that most of their contribution is due to the excess of protons, with pions being the next contribution in importance, that induce a decrease of the (total) negative excess track length of the shower of $\\sim1-2\\%$ above 1 PeV. Although the actual field emitted by the hadrons in the shower is very small, the hadronic component carries a significant fraction of the shower energy. The importance of this energy balance is illustrated by the effects of the EM energy fraction on the radio emission. Given our results, following the actual tracks of hadrons in the shower turns out not to be essential for the field calculation, but accounting for the hadronic energy of the shower is. We have compared the results obtained using SIBYLL 2.1 \\cite{SIBYLL21} and QGSJET01 \\cite{QGSJET01}, and found that there is a small dependence of the normalization of the emitted electric field on the hadronic model, up to $\\sim 4\\%$ lower in QGSJET01 at the Cherenkov angle. This is due to the higher number of protons, pions and muons when QGSJET01 is used, which increases the positive projected excess track-length of protons, pions and muons, with protons dominating the difference. Also more protons, pions and muons means means less electrons and positrons in the shower, decreasing the negative excess track. The net effect is a lower negative projected excess track as a whole, diminishing the field. At angles away from the Cherenkov angle, the longitudinal excess profile becomes important, and the shorter showers obtained using QGSJET lead to the maximum of the emission at higher frequencies, when compared to SIBYLL showers. Finally, we have used our simulations of the emission due to the Askar'yan effect in hadronic showers to obtain new parameterizations of the radio pulse frequency spectrum which are described in Appendix A. We note that our results for proton showers can also be applied to approximately model the Cherenkov emission from hadronic showers induced in high-energy neutrino interactions. This is shown in the bottom panel of Fig.~\\ref{fig:MC_vs_param} where we have plotted the frequency spectrum obtained in \\ZHAireS simulations of a neutrino-induced shower in a neutral current interaction (or a charged current interaction of a muon neutrino) in which the secondaries from the fragmentation of the nucleon carry 2 EeV of energy in total. The products of the neutrino-nucleon interaction, mainly pions, were obtained with the HERWIG interaction Monte Carlo code \\cite{HERWIG}. The frequency spectrum obtained in \\ZHAireS simulations is compared to the parameterization of a proton shower with energy $E_0=2$ EeV as given in Appendix A." }, "1005/1005.4767_arXiv.txt": { "abstract": "Type I X-ray bursts are thermonuclear stellar explosions driven by charged-particle reactions. In the regime for combined H/He-ignition, the main nuclear flow is dominated by the {\\it rp-process} (rapid proton-captures and $\\beta^+$-decays), the $3\\alpha$-reaction, and the {\\it $\\alpha$p-process} (a suite of ($\\alpha$,p) and (p,$\\gamma$) reactions). The main flow is expected to proceed away from the valley of stability, eventually reaching the proton drip-line beyond A = 38. Detailed analysis of the relevant reactions along the main path has only been scarcely addressed, mainly in the context of parameterized one-zone models. In this paper, we present a detailed study of the nucleosynthesis and nuclear processes powering type I X-ray bursts. The reported 11 bursts have been computed by means of a spherically symmetric (1D), Lagrangian, hydrodynamic code, linked to a nuclear reaction network that contains 325 isotopes (from $^1$H to $^{107}$Te), and 1392 nuclear processes. These evolutionary sequences, followed from the onset of accretion up to the explosion and expansion stages, have been performed for 2 different metallicities to explore the dependence between the extension of the main nuclear flow and the initial metal content. We carefully analyze the dominant reactions and the products of nucleosynthesis, together with the the physical parameters that determine the light curve (including recurrence times, ratios between persistent and burst luminosities, or the extent of the envelope expansion). Results are in qualitative agreement with the observed properties of some well-studied bursting sources. Leakage from the predicted SbSnTe-cycle cannot be discarded in some of our models. Production of $^{12}$C (and implications for the mechanism that powers superbursts), light p-nuclei, and the amount of H left over after the bursting episodes will also be discussed. ", "introduction": "Type I X-ray bursts (hereafter, XRBs) are cataclysmic stellar events. They are powered by thermonuclear runaways (TNR) in the H/He-rich envelopes accreted onto neutron stars in close binary systems (see reviews by Bildsten 1998; Lewin et al. 1993, 1995; Psaltis 2006; Schatz \\& Rehm 2006; Strohmayer \\& Bildsten 2006). These events constitute the most frequent type of thermonuclear stellar explosion in the Galaxy (the third, in terms of total energy output after supernovae and classical novae), in part because of their short recurrence period (hours to days). About $90$ Galactic low-mass X-ray binaries exhibiting such bursting behavior (with burst durations of $\\tau_{burst} \\sim$ 10 - 100 s) have been found since the discovery of XRBs by Grindlay et al. (1976), and independently, by Belian et al. (1976). Type I X-ray bursts and their associated nucleosynthesis have been extensively modeled by different groups (see pioneering work by Woosley \\& Taam 1976, Maraschi \\& Cavaliere 1977, and Joss 1977), reflecting the astrophysical interest in determining the nuclear processes that power the explosion, the light curve, as well as in providing reliable estimates for the chemical composition of the neutron star surface (see Schatz et al. 1999, Parikh et al. 2008, and references therein). With a neutron star hosting the explosion, temperatures and densities in the accreted envelope reach high values: T$_{peak} > 10^9$ K, and $\\rho \\sim 10^6$ g cm$^{-3}$. As a result, detailed nucleosynthesis studies require the use of hundreds of isotopes, linked by thousands of nuclear interactions, extending all the way up to the SnSbTe-mass region (Schatz et al. 2001) or beyond (the extent of the nuclear activity\\footnote{The nuclear activity reflects the changes in composition driven by different nuclear processes (i.e., p- and $\\alpha$-capture reactions, $\\beta$-decays, ...) that take place in the envelope at different stages of the burst. In this work, the extent of the nuclear activity is arbitrarily defined by the heaviest nucleus that achieves a mass fraction $>10^{-9}$.} in the XRB nucleosynthesis study of Koike et al. 2004 reaches $^{126}$Xe). Indeed, the extent of the rp-process in XRBs is still not clear: recent experimental work now shows that it will be more difficult to reach the SnSbTe-mass region (Elomaa et al. 2009). Because of computational constraints, XRB nucleosynthesis studies have been traditionally performed using limited nuclear reaction networks, truncated near Ni (Woosley \\& Weaver 1984; Taam et al. 1993; Taam, Woosley, \\& Lamb 1996 --all using a 19-isotope network), Kr (Hanawa, Sugimoto, \\& Hashimoto 1983 --274-isotope network; Koike et al. 1999 --463 nuclides), Cd (Wallace \\& Woosley 1984 --16-isotope network), or Y (Wallace \\& Woosley 1981 --250-isotope network). On the other hand, Schatz et al. (1999, 2001) have carried out very detailed nucleosynthesis calculations with a network containing more than 600 isotopes (up to Xe, in Schatz et al. 2001), but using a one-zone approach. Koike et al. (2004) have also performed detailed one-zone nucleosynthesis calculations, with temperature and density profiles obtained from a spherically symmetric evolutionary code, linked to a 1270-isotope network extending up to $^{198}$Bi. Until recently, it has not been possible to couple hydrodynamic stellar calculations (in 1-D) and detailed networks. Recent efforts include Fisker et al. (2004, 2006, 2007, 2008), and Tan et al. (2007) ($\\sim$ 300 isotopes, up to $^{107}$Te), Jos\\'e \\& Moreno (2006) (2640 reactions and 478 isotopes, up to Te), or Woosley et al. (2004) and Heger et al. (2007) (up to 1300 isotopes with an adaptive network). This has prompted a detailed analysis of the nuclear activity powering the bursts. The most detailed work to date is that of Fisker et al. (2008), in the context of the 1-D general relativistic hydrodynamic code AGILE (Liebend\\\"orfer et al. 2002), linked to a nuclear reaction network containing 304 isotopes: a thorough analysis of the main nuclear activity in one characteristic burst is reported (although details for a sequence of 5 consecutive, 'representative' bursts are also outlined). However, because of the specific choice of metallicity (Z = $10^{-3}$, for the accreted matter) and mass-accretion rate ($\\dot M \\sim 10^{17}$ g s$^{-1}$) adopted, the nuclear activity does not extend much beyond mass $A \\sim 65$, as a result of compositional inertia effects, that quench further extension of the nuclear path. Hence, the flow does not reach the {\\it SnSbTe}-mass region, which was suggested as a natural endpoint in XRB nucleosynthesis studies (see Schatz et al. 1999,2001). Clearly, the identification of the most relevant reactions in the $A \\sim 65 - 100$ mass region remains to be addressed in detail in the framework of hydrodynamic simulations. This is particularly relevant since, as first pointed out by Hanawa et al. (1983), proton captures on heavy nuclei (i.e., the {\\it rp}-process) have a dramatic effect on the shape of XRB light curves. To this end, a new set of type I X-ray bursts have been computed with SHIVA, a 1-D, spherically symmetric, hydrodynamic, implicit, Lagrangian code, used extensively in the modeling of classical nova outbursts (see Jos\\'e \\& Hernanz 1998). The code has been linked to a fully updated nuclear reaction network containing 324 nuclides and 1392 nuclear processes, a subset of that used in Parikh et al. (2008), and includes the most relevant charged-particle induced reactions occurring between $^1$H and $^{107}$Te, as well as their corresponding reverse processes. It is worth noting that the size of this network is similar (though slightly larger) to that adopted by Fisker et al. (2008). In order to set up the reaction rate library for our study, we started by adopting the proton drip line predicted by Audi et al. (2003a, 2003b). Experimental rates are available for a small subset of reactions (adopted from Angulo et al. 1999, Iliadis et al. 2001, and some recent updates for selected reactions). For all other reactions for which experimental rates are not available, we used the rates from the Hauser-Feshbach codes MOST (Goriely 1998; Arnould \\& Goriely 2006) and NON-SMOKER (Rauscher \\& Thielemann 2000; for details see Parikh et al. 2008). Neutron captures are disregarded since our early test calculations revealed that they play a minor role in XRB nucleosynthesis. All reaction rates incorporate the effects of thermal excitations in the target nuclei (Rauscher \\& Thielemann 2000). Screening factors are taken from Graboske et al. (1973) and DeWitt, Graboske, \\& Cooper (1973). For the weak interactions, $\\beta$-delayed nucleon emission and laboratory decay rates (Audi et al. 2003a) have been adopted. For a discussion of employing stellar versus laboratory decay rates, see Woosley et al. (2004). It is worth noting, however, that many computed stellar decay rates ( Fuller et al. 1982a, 1982b; Langanke \\& Martinez-Pinedo 2000) do not converge to their laboratory values at lower temperatures and densities, calling into question the model used for these calculations. Studies employing properly converging stellar decay rates for all isotopes relevant to XRB nucleosynthesis have not been performed by any group yet, and would certainly be interesting, although the results presented in this work would not be dramatically affected by their inclusion. SHIVA uses a time-dependent formalism for convective transport whenever the characteristic convective timescale becomes larger that the integration time step. Partial mixing between adjacent convective shells is treated by means of a diffusion equation (Prialnik, Shara, \\& Shaviv 1979). No additional semiconvection or thermohaline mixing is considered. Models make use of Iben's (1975) opacity fits, better suited than the OPAL opacities for astrophysical environments that exhibit strong variations in metallicity, as in XRB nucleosynthesis. However, plans to incorporate these more realistic opacities are currently underway. The adopted equation of state includes contributions from the electron gas (with different degrees of degeneracy; Blinnikov et al. 1996), a multicomponent ion plasma, and radiation; Coulomb corrections to the electronic pressure are also taken into account. Accretion is computed by redistributing material through a constant number of envelope shells (see Kutter \\& Sparks 1980, for details). To handle this, a tiny envelope, containing $1.1 \\times 10^{18}$ g of material (less than 1 permil of the total envelope mass accreted during the first bursting episode), distributed through all the envelope shells, is put initially in place(the influence of the number of envelope shells on burst properties will be discussed in Section 3). The model is then relaxed using a few, very large timesteps, to guarantee hydrostatic equilibrium. The temperature at the bottom of the envelope barely reaches $2.7 \\times 10^7$ K, whereas the density is just $1.4 \\times 10^3$ g cm$^{-3}$ (corresponding to a pressure of $5.7 \\times 10^{18}$ dyn cm$^{-2}$). Mass accretion and nuclear reactions are then initiated. Special emphasis is placed on the effect of the initial metallicity of the accreted matter on the main nuclear path, which in turn, will affect the final post-burst envelope composition and the shape of the light curves. The structure of the manuscript is as follows: in Section 2, we analyze the main features (nuclear path, nucleosynthesis, light curves, etc) of a series of 4 bursts computed in a model with solar-like accreted material. The effect of the resolution adopted in this model is discussed in Section 3. A detailed analysis of the impact of the metallicity of the accreted material on burst properties is given in Section 4. Finally, a comparison with previous work, together with a thorough analysis of the corrections posed by general relativity, are discussed in Section 5. ", "conclusions": "\\subsection{General relativity corrections} The calculations reported here have been performed assuming Newtonian gravity. Since the envelope layers are very thin, it is easy to introduce general relativity corrections to this Newtonian framework (see Ayasli \\& Joss 1982, Lewin et al. 1993, Taam et al. 1993, Cumming et al. 2002, and Woosley et al. 2004). To this end, the surface gravity is rewritten as $g = G M_*/R_*^2 (1+z)$, where M$_*$ is the mass, R$_*$ is the stellar radius (defined in such a way that the surface area is 4$\\pi R_*^2$), and $z$ is the gravitational redshift given by $1 + z = (1 - 2G M_*/R_*c^2)^{-1/2}$. Our models of M$_* = 1.4$ M$_\\odot$ require R$_* = 14.3$ km, and a gravitational redshift of $z = 0.19$. Following Woosley et al. (2004), once the redshift and radius are determined, it is straightforward to derive the set of correcting factors to the physical magnitudes described above for a suitable observer at infinity. Hence, recurrence times and burst durations should be increased by a factor $1 + z$. The mass-accretion rate as well as the burst luminosity have to take into account both the difference in surface area (compared to the Newtonian framework) and the gravitational redshift term. The energy and rest mass-accretion rate scale as R$_*^2$/$(1+z)$, while the luminosity $\\propto$ R$_*^2$/$(1+z)^2$. However, when M$_*$ is taken exactly as $M_{NS}$ (Newtonian framework), the surface area and redshift corrections for energy and mass accretion rate cancel out, since g $\\propto$ $(1+z)/$R$_*^2 = const$, and hence, no correction to the observed burst energy or mass-accretion rate is necessary, while the luminosity correction is simply given by $1/(1+z) = 0.84$. In addition, the accretion luminosity for an observer at infinity changes only by a factor 1.012, that is, the ratio between gravitational energy released per unit mass in general relativity, $c^2.z/(1+z)$, and the Newtonian value, $GM_{NS}/R_{NS}$. Finally, the luminosity measured at infinity will be smaller by a factor of $(1+z)=1.19$. \\subsection{Comparison with previous work} For consistency, the results discussed in this paper have been compared with those reported in previous work (obtained with similar hydrodynamic codes or in the framework of one-zone models). As emphasized in Section 2, model 1 is qualitatively similar to model ZM of Woosley et al. (2004). The twelve bursts computed by Woosley et al. (2004) in a Newtonian frame were characterized by recurrence times of about $\\sim 2.7$ hr, peak luminosities of L$_{peak} \\sim (1.5 - 2) \\times 10^{38}$ erg s$^{-1}$, and ratios between persistent and burst luminosities of $\\alpha \\sim 60-65$. Our calculations (model 1, Newtonian frame) yield $\\tau_{rec} \\sim 5 - 6.5$ hr, L$_{peak} \\sim (3 - 7) \\times 10^{38}$ erg s$^{-1}$, and $\\alpha \\sim 35-40$. The role played by the metallicity of the accreted material (model 3, with Z = Z$_\\odot$/20 = 0.001) qualitatively agrees with the pattern reported by Woosley et al. (2004) (see also, Heger et al. 2007). Longer recurrence times of $\\sim 9$ hours, peak temperatures of about $(1.3-1.4) \\times 10^{9}$ K, and ratios between persistent and burst luminosities of $\\alpha \\sim 20-30$ (with L$_{peak}$ $\\sim 10^{38}$ erg s$^{-1}$) have been obtained in the 5 bursts computed in model 3. In turn, the fifteen bursts computed by Woosley et al. (2004) for model zM are characterized by recurrence times of about $3 - 3.5$ hr, peak luminosities of L$_{peak} \\sim 10^{38}$ erg s$^{-1}$, and ratios between persistent and burst luminosities of $\\alpha \\sim 50-60$. Results reveal a dependence of burst properties on the metallicity of the accreted material: the smaller the metal content, the larger the recurrence time (and the smaller the $\\alpha$). In turn, explosions in metal-deficient envelopes (i.e., model 3) are characterized by lower peak luminosities and longer decline times, in agreement with the pattern described in Woosley et al. (2004) and Heger et al. (2007). Model 3 bears as well a clear resemblance with the model computed by Fisker et al. (2008). In that work, five representative bursting sequences were analyzed, with $\\tau_{rec} \\sim 3.5 - 4$ hr, L$_{peak} \\sim (7 - 8) \\times 10^{37}$ erg s$^{-1}$, and $\\alpha \\sim 65-70$, as measured at infinity. Despite the qualitative similaries in the gross properties of the bursts presented in this paper (as well as in the role played by the metallicity of the accreted material) and those reported in previous work, a quantitative comparison reveals some discrepancies that are worth analyzing. In model 1 (with Z = Z$_\\odot$), our computations yield systematically larger (by a factor of $\\sim$ 2) recurrence times and peak luminosities (and hence, lower $\\alpha$) than model ZM of Woosley et al. (2004). Similar results are found in the low-metallicity case (model 3, with Z = Z$_\\odot$/20) when compared with model zM of Woosley et al. (2004), except for the peak luminosities that turn out to be very similar. It is also worth noting that the values reported by Fisker et al. (2008) show discrepancies with respect to Woosley et al. (2004), in particular, lower peak luminosities (and larger $\\alpha$). A major difference concerns the much larger effect played by the metallicity of the accreted material in this work as compared with Woosley et al. (2004), who explained the moderate effect found as due to compositional inertia washing out the influence of the initial metallicity. Another striking issue concerns the extremely large differences in the gross physical characteristics -nucleosynthesis, energies or recurrence times- between the first and subsequent bursts, as reported by Woosley et al. (2004). In terms of nucleoynthesis or nuclear activity, Figs. \\ref{fig:abun_b3}, \\ref{fig:lum_m3_5b}, \\& \\ref{fig:LUMI_OVER}, reveal a similar behavior for the different bursts (although a somewhat lower production of intermediate-mass elements as well as of the heaviest elements is reported for the first burst computed in model 3). Very limited information on the nucleosynthetic yields obtained in model ZM is given in Woosley et al. (2004). Thus, we will restrict the discussion on the extent of the nuclear activity and on the resulting chemical abundance pattern to model 3, through a brief comparison with the work reported by Schatz et al. (2001), Fisker et al. (2008) and Woosley et al. (2004) (for model zM). It is worth noting that both the nucleosynthetic end-point (located in the SnSbTe-mass region) and the main nuclear path in the A $\\sim 50 - 100$ mass region obtained in this work (passing through a suite of different nuclei such, as $^{55}$Co, $^{60}$Zn, $^{70}$Br, $^{75}$Rb, $^{85}$Mo, $^{90}$Rh, or $^{100,105}$Sn) are similar to those reported by Schatz et al. (2001) in the framework of one-zone calculations. Whereas the main nuclear path in the first burst of model zM (Woosley et al. 2004) is very similar to the one reported in this work, compositional inertia causes a more limited extension of the nuclear activity in the successive bursts of Woosley et al.: hence, while the three most abundant nuclei at the bottom of the envelope are $^{106}$Sn and $^{104,106}$In at the end of the first burst, this switches to $^{64}$Zn, $^{68}$Se, and $^{32}$S (a similar trend is also reported by Fisker et al. 2008). In this work, the mass-averaged composition at the end of the first burst computed for model 3 (see Table 8) is dominated (aside from some residual H and $^{4}$He) by the presence of $^{104}$Pd (0.05, by mass) and $^{105}$Ag (0.08), while X($^{64}$Zn) $\\sim$ 0.04. But the peak at the end of the abundance distribution (see Fig. \\ref{fig:abun_b3}) increases with subsequent bursts up to a plateau value, which indicates that these heavy nuclei are still produced in similar quantities. This is very different to the results reported by Woosley et al. (2004). Indeed, at the end of the fifth burst, the abundance pattern, shows still a significant presence of heavy species (i.e., X($^{105}$Ag) $\\sim$ 0.1, X($^{104}$Pd) $\\sim$ 0.08, and X($^{94}$Tc) $\\sim$ 0.05), together with a simultaneous increase in the abundances of intermediate-mass elements, such as $^{60}$Ni (0.06), $^{64}$Zn (0.09), $^{68}$Ge (0.07), or $^{72}$Se (0.04). It is finally worth noting that, in agreement with all previous hydrodynamic studies, both models 1 and 3 yield very small post-burst abundances of $^{12}$C, below the threshold amount required to power superbursts. Even though only a few bursts have been computed for these models, they already show a trend on the amount of $^{12}$C that may be expected after many more bursts. Finally, it is also worth mentioning that large differences exist between the hydrodynamic simulations reported here (see also Woosley et al. 2004, and Fisker et al. 2008) and those based on one-zone models (i.e., Schatz et al. 1999, 2001) as regards the shape of the light curve accompanying the bursting episodes (the primary difference being the presence of a long-lasting plateau in the latter). The origin of the discrepancies reported is not totally clear and would require additional hydrodynamic studies. Notice, however, that the local surface gravity of our model is somewhat smaller than that adopted in the abovementioned works: whereas a 10 km radius is assigned to the 1.4 M$_\\odot$ neutron star in Woosley et al. (2004) ($g = 1.86 \\times 10^{14}$ cm s$^{-2}$), the integration of the neutron star structure from the core to its surface, in hydrostatic equilibrium, yielded 13.1 km (14.3 km, after general relativity corrections are introduced; see Subsection 5.1), for our 1.4 M$_\\odot$ neutron star (corresponding to a surface gravity of $g = 1.08 \\times 10^{14}$ cm s$^{-2}$); in turn, the calculations reported by Fisker et al. (2008), in a general relativity framework, relied on a 11 km (1.4 M$_\\odot$) neutron star, for which $g = 1.53 \\times 10^{14}$ cm s$^{-2}$. Although XRB properties depend weakly upon the neutron star mass (or surface gravity), part of the differences outlined between the three studies can be attributed to the combined effect of the adopted neutron star size (surface gravity) and to differences in the input physics (i.e., nuclear reaction network, opacities, treatment of convection). In particular, the use of Iben's opacities may have some effect on the peak luminosities achieved since the larger OPAL opacities will likely decrease the amount of energy radiated away from the star. Moreover, the inclusion of semiconvection and thermohaline mixing would have a minor effect in the properties of the explosions, likely affecting the appearance of marginal convective transport between bursts (see Woosley et al. 2004, Fisker et al. 2008). It is however worth noting that the convective pattern shown in Figs. \\ref{fig:conv_tem_b4} \\& \\ref{fig:LUMI_OVER} is similar to those reported in previous work: namely that convection sets in as soon as superadiabatic gradients are established in the envelope, following the early stages of the TNR and the corresponding rise in temperature; it reaches the surface and begins to recede before the observed burst properly commences, shutting off thereafter (Woosley et al. 2004, Fisker et al. 2008). The potential impact of XRB nucleosynthesis on Galactic abundances is still a matter of debate. Matter accreted onto a neutron star of mass $M$ and radius $R$ releases $G M m_p/R \\sim 200$ MeV nucleon$^{-1}$, whereas only a few MeV nucleon$^{-1}$ are released from thermonuclear fusion. Thus ejection from a neutron star is unlikely. However, it has been suggested that radiation-driven winds during photospheric radius expansion may lead to ejection of a tiny fraction of the envelope (containing nuclear processed material; see Weinberg et al. 2006; MacAlpine et al. 2007). Indeed, XRBs have been proposed as a possible source of the light p-nuclei $^{92,94}$Mo and $^{96,98}$Ru (Schatz et al. 1998, 2001). No matter is ejected in any of the models reported in this work, a result fully independent of the adopted resolution and in agreement with all previous hydrodynamic simulations (Woosley et al. 2004, Fisker et al. 2008). Moreover, it is worth noting that, as shown in Figs. \\ref{fig:abun_b1} \\& \\ref{fig:abun_b3}, the abundances of many species synthesized during the bursts decrease remarkably towards the outer envelope layers, because of inefficient convective transport (see Figs. \\ref{fig:conv_tem_b4} \\& \\ref{fig:LUMI_OVER}). To assess the possible contribution to the Galactic abundances, one has to rely on the abundances of the outer envelope layers (the only ones that have a chance to be ejected by radiation-driven winds). This shows the limitations posed by one-zone nucleosynthesis calculations, in which the chemical species synthesized in the innermost layers are, by construction, assumed to represent the whole (chemically homogeneous) envelope. The mass fractions of these p-nuclei, obtained in model 3, drop by more than an order of magnitude in the outer envelope layers (as compared with the values achieved at the innermost envelope; see Fig. \\ref{fig:MORU}); the resulting overproduction factors, $f \\sim 10^6$, are several orders of magnitude smaller than those required to account for the origin of these problematic nuclei (see Weinberg et al. 2006, Bazin et al. 2008), in sharp contrast with the results obtained on the basis of one-zone calculations (Schatz et al. 1998, 2001)." }, "1005/1005.3830_arXiv.txt": { "abstract": "{Planets are formed in disks around young stars. With an age of $\\sim$~10\\,Myr, \\object{TW~Hya} is one of the nearest T~Tauri stars that is still surrounded by a relatively massive disk. In addition a large number of molecules has been found in the \\object{TW~Hya} disk, making \\object{TW~Hya} the perfect test case in a large survey of disks with {\\it Herschel}--{\\it PACS} to directly study their gaseous component. We aim to constrain the gas and dust mass of the circumstellar disk around \\object{TW~Hya}. We observed the fine-structure lines of [\\ion{O}{I}] and [\\ion{C}{II}] as part of the Open-time large program {\\it GASPS}. We complement this with continuum data and ground-based $^{12}$ CO 3--2 and $^{13}$CO 3--2 observations. We simultaneously model the continuum and the line fluxes with the 3D Monte-Carlo code {\\it MCFOST} and the thermo-chemical code {\\it ProDiMo} to derive the gas and dust masses. We detect the [\\ion{O}{I}] line at 63 $\\mu$m. The other lines that were observed, [\\ion{O}{I}] at 145~$\\mu$m and [\\ion{C}{II}] at 157~$\\mu$m, are not detected. No extended emission has been found. Preliminary modeling of the photometric and line data assuming [$^{12}$CO]/[$^{13}$CO]=69 suggests a dust mass for grains with radius $<$ 1~mm of $\\sim$ 1.9 $\\times$ 10$^{-4}$ M$_{\\odot}$ (total solid mass of 3 $\\times$ 10$^{-3}$ M$_{\\odot}$) and a gas mass of (0.5--5) $\\times$ 10$^{-3}$ M$_{\\odot}$. The gas-to-dust mass may be lower than the standard interstellar value of 100.} ", "introduction": "Planets are formed in the disks that surround a large fraction of T~Tauri stars. Knowledge of the gas mass available at different disk ages is essential to constrain giant planet formation models. Most studies estimate the dust mass from millimeter continuum emission and assume the gas mass is a factor of 100 times larger. This conversion factor has been calibrated for the interstellar medium but is likely not valid for disks, especially those that are evolving toward debris disks or where most of the gas has accreted onto the planetary atmosphere. Disk gas mass estimates derived from observations of $^{12}$CO and optically thinner $^{13}$CO emission are at least a factor of 10 lower than the mass derived from dust observations assuming the interstellar medium conversion factor. The discrepancy has been ascribed to CO photodissociation at disk atmosphere and freeze-out onto cold dust grains in the disk midplane \\citep[e.g.,][]{Qi2004ApJ...616L..11Q,Thi2001ApJ...561.1074T}. An alternative explanation is that the CO abundance is not different and the gas in disks has been depleted. The {\\it PACS} instrument \\citep{Poglistch2010} on-board the {\\it Herschel Space Telescope} \\citep{Pilbratt2010} makes it possible to observe lines from species that result from the photodissociation of CO (atomic oxygen and singly ionized carbon). With observations of all the major gas-phase carbon and oxygen-bearing species, we can more precisely constrain the disk gas mass. At a distance of $\\sim$~56~pc \\citep{Wichmann1998MNRAS.301L..39W}, \\object{TW~Hya} is one of the nearest classical T~Tauri stars with an estimated age of 10~Myr \\citep{Barrado2006A&A...459..511B}. Its proximity allows us to attain an order of magnitude higher mass sensitivity than objects in the Taurus molecular cloud. Fits to the spectral energy distribution (SED) provide an estimate of the gas disk mass of 6 $\\times$ 10$^{-2}$ M$_{\\odot}$ after applying a conversion factor of $\\sim$~75 \\citep{Calvet2002ApJ...568.1008C}. This large disk mass at this advanced age is surprising as the median disk lifetime is only 2-3 Myr \\citep{Haisch2001ApJ...553L.153H}. \\object{TW~Hya} is considered a transition object with an optically thin inner cavity and an optically thick outer disk \\citep{Calvet2002ApJ...568.1008C,Ratzka2007A&A...471..173R}. The fit to the SED also suggests that grains have grown to at least $\\sim$~1~cm. The star \\object{TW~Hya} was observed as a Science Demonstration Project object and is part of the {\\it Herschel-GASPS} program \\citep{Dent2010}. {\\it Herschel} observations of the disk around the Herbig~Ae star \\object{HD169142} are presented by \\citet{Meeus2010}. In this letter we use fine-structure lines in addition to continuum data and CO (sub)millimeter lines to directly constrain the gas mass and compare it to the dust mass derived from fits to the SED. \\begin{figure}[!ht] \\centering {\\includegraphics[angle=90,width=8cm]{14578fig1.ps}} \\caption{{\\it Herschel-PACS} spectrum centred around the OI 63 $\\mu$m line on the upper-left panel.} \\label{fig_OI63_herschel} \\end{figure} \\begin{table}% \\centering \\small \\caption{Lines observed by {\\it Herschel-PACS}. The errors and upper limits are 3~$\\sigma$. The calibration error adds an extra $\\sim$~40\\% uncertainty. The CO data also have uncertainties of 30\\%.} \\label{table_results} \\begin{tabular}{lllll} \\hline \\noalign{\\smallskip} \\multicolumn{1}{c}{Line} & \\multicolumn{1}{c}{Cont. flux} & \\multicolumn{1}{c}{Obs.}&\\multicolumn{1}{c}{GH08}&\\multicolumn{1}{c}{M08}\\\\ & \\multicolumn{1}{c}{(Jy)}& \\multicolumn{3}{c}{(10$^{-18}$ W m$^{-2}$)}\\\\ \\noalign{\\smallskip} \\hline \\noalign{\\smallskip} O{\\sc I}\\ $^3$P$_1 \\rightarrow ^3$P$_2$ & 2.99~$\\pm$~0.14 & \\phantom{$<$}36.5~$\\pm$~12.1 & 124-161 & 412\\\\ O{\\sc I}\\ $^3$P$_0 \\rightarrow ^3$P$_1$ & 7.00~$\\pm$~0.05 & $<$\\ \\ 5.5 & 25-41 & 11\\\\ C{\\sc II}\\ $^2$P$_{3/2} \\rightarrow ^2$P$_{1/2}$ & 8.79~$\\pm$~0.08 & $<$\\ \\ 6.6& 0.8-12 & 0.06\\\\ CO 3--2 & \\multicolumn{1}{c}{n.a.} & \\phantom{$<$}\\ \\ 0.43 & 0.3-0.6 & n.a.\\\\ $^{13}$CO 3--2 & \\multicolumn{1}{c}{n.a.} & \\phantom{$<$}\\ \\ 4.4 $\\times$ 10$^{-2}$& n.a.& n.a.\\\\ \\noalign{\\smallskip} \\hline \\end{tabular} \\end{table} \\normalsize ", "conclusions": "The {\\it Herschel-PACS} spectral observations were used to constrain the gas disk mass surrounding the 10 Myr T~Tauri star \\object{TW~Hya}. We estimate the gas mass to be (0.5--5) $\\times$ 10$^{-3}$ M$_\\odot$ compared to the dust mass ($a_{\\mathrm{max}}<$~1mm) of 1.9 $\\times$ 10$^{-4}$ M$_\\odot$. The gas-to-dust mass ratio is $\\sim$2.6--26, lower than the standard interstellar value of 100. The ratio gas-to-total-mass in solids is $\\sim$0.17--1.7. Although the disk is still massive, a significant fraction of the primordial gas has already disappeared. A large fraction of the primordial gas may have been evaporated due to the strong X-ray flux from \\object{TW~Hya}. \\object{TW~Hya} is the first example where the disk gas mass around a transitional T~Tauri star can be determined accurately and directly from gas phase lines. However, more detailed modeling that includes X-ray physics and $^{13}$CO photochemistry is needed to confirm the low gas mass." }, "1005/1005.2188_arXiv.txt": { "abstract": "We present a complete numerical study of cosmological models with a time dependent coupling between the dark energy component driving the present accelerated expansion of the Universe and the Cold Dark Matter (CDM) fluid. Depending on the functional form of the coupling strength, these models show a range of possible intermediate behaviors between the standard $\\Lambda $CDM background evolution and the widely studied case of interacting dark energy models with a constant coupling. These different background evolutions play a crucial role in the growth of cosmic structures, and determine strikingly different effects of the coupling on the internal dynamics of nonlinear objects. By means of a suitable modification of the cosmological N-body code {\\small GADGET-2} we have performed a series of high-resolution N-body simulations of structure formation in the context of interacting dark energy models with variable couplings. Depending on the type of background evolution, the halo density profiles are found to be either less or more concentrated with respect to $\\Lambda $CDM, contrarily to what happens for constant coupling models where concentrations can only decrease. However, for some specific choice of the interaction function the reduction of halo concentrations can be larger than in constant coupling scenarios. We also find that different types of coupling evolution determine specific features in the growth of large scale structures, like peculiar distortions of the matter power spectrum shape or different time evolutions of the halo mass function. Furthermore, also for time dependent couplings baryons and CDM develop a bias already on large scales, which is progressively enhanced for smaller and smaller scales, and the effect can be significantly larger compared to constant coupling scenarios. The same happens to the baryon fraction of halos, which can be more significantly reduced below its universal value in variable coupling models with respect to constant coupling cosmologies. In general, we find that time dependent interactions between dark energy and CDM can in some cases determine stronger effects on structure formation as compared to the constant coupling case, with a significantly weaker impact on the background evolution of the Universe, and might therefore provide a more viable possibility to alleviate the tensions between observations and the $\\Lambda $CDM model on small scales than the constant coupling scenario. ", "introduction": "\\label{i} According to the present interpretation of the vast amount of cosmological data that have been collected over the last years, the Universe in which we live has a nearly flat spatial geometry, it expands with a rate of $\\sim 70 $ km s$^{-1}$ Mpc$^{-1}$, and the total amount of matter it contains accounts only for $\\sim 27\\%$ of the critical energy density that is required to justify its spatial flatness. Furthermore, the expansion seems to have entered an accelerated phase since about $6$ billion years, and the missing $\\sim 73\\%$ of the critical energy density is assumed to be in the form of some dark energy (DE) component able to drive such accelerated expansion. This detailed picture can be obtained by combining a wide variety of different and complementary cosmological datasets, ranging from Cosmic Microwave Backround (CMB) \\citep{wmap5,wmap7}, to Large Scale Structure surveys \\citep{Percival_etal_2001,Cole_etal_2005,Reid_etal_2010}, to observations of Type Ia Supernovae (SnIa) \\citep{Riess_etal_1998,Perlmutter_etal_1999,SNLS,Kowalski_etal_2008} and Baryon Acoustic Oscillations (BAO) \\citep[\\eg][]{Percival_etal_2009}. While the matter fraction of the Universe is known to be composed only for $\\sim 15\\%$ by baryonic particles, with the remaining mass in the form of some collisionless, weakly interacting Cold Dark Matter (CDM) component, the nature of the DE fraction is yet completely unknown, and its understanding constitutes one of the major challenges in modern cosmology. The simplest possibility of a cosmological constant $\\Lambda $, whose energy density remains constant throughout the whole expansion history of the Universe, is in good agreement with a very large amount of cosmological and astrophysical data, and this is the reason for the establishment of the $\\Lambda $CDM scenario as the present standard cosmological model. Nevertheless, the nature of the cosmological constant raises two fundamental questions concerning the very finely tuned value of its energy density (the ``fine tuning problem\") and the beginning of its domination over CDM only at relatively recent cosmological epochs (the ``coincidence problem\"). For this reason, alternative possibilities of dynamically evolving DE components have been proposed, in particular models where the DE is identified with a classical scalar field as for the case of quintessence \\citep{Wetterich_1988,Ratra_Peebles_1988} or k-essence \\citep{ArmendarizPicon_etal_2000,kessence}. As a further extension of these dynamical DE scenarios, different possible forms of interaction between the DE component and the matter sector of the Universe have been suggested and investigated in the literature, as \\eg the generalized Chaplygin gas \\citep[see \\eg][]{Kamenshchik_etal_2001, Bilic_Tupper_Viollier_2002,Bento_etal_2002,Carturan_Finelli_2003,Amendola_etal_2003b}, unified dark matter models \\citep{Mainini_Bonometto_2004,Bertacca_etal_2007,Bertacca_Bartolo_Matarrese_2010}, extended quintessence models \\citep{Perrotta_etal_2000,Baccigalupi_Matarrese_2000,Pettorino_etal_2005}, and coupled DE \\citep{Wetterich_1995,Amendola_2000,Amendola_2004,Pettorino_Baccigalupi_2008}. Any form of interaction between the DE sector and other matter species, like CDM or massive neutrinos \\citep[as in \\eg ][]{Amendola_Baldi_Wetterich_2008} would leave distinctive features in the background expansion history of the Universe and in the growth of cosmic structures \\citep[see \\eg ][]{Brax_etal_2010} which could provide new ways to tackle the problem of the nature of DE. It is therefore essential to understand the impact that the interaction would have on observable quantities as \\eg the properties of CMB \\citep{Amendola_etal_2003,Amendola_Quercellini_2003,Bean_etal_2008,LaVacca_etal_2009}, of large scale structure formation \\citep{Koivisto_2005,Bean_etal_2008,Baldi_Pettorino_2010,Baldi_Viel_2010}, and of the nonlinear newtonian dynamics at small scales \\citep{Perrotta_etal_2003,Mainini:2006zj,Maccio_etal_2004,Saracco_etal_2010,Baldi_etal_2010}. In the present work, in particular, we consider the case of a cosmological scenario where the DE scalar field interacts with the CDM fluid with a coupling strength that evolves in time, therefore generalizing the very widely studied case of constant couplings for the DE interaction. Other forms of effectively time dependent couplings, where the interaction depends on linear or nonlinear combinations of the energy densities of the interacting fluids and of their time derivatives have been studeid in \\eg \\citet{Barrow_Clifton_2006,CalderaCabral_2009,Chimento_2010}. Here we want to focus on general functional forms of the coupling strength in the context of coupled quintessence models where the interaction term is proportional to the energy density of the matter coupled fluid. One of the main motivations behind the introduction of a time dependence of the DE coupling to the matter sectors -- besides the fact that an evolving coupling is {\\em per se} a more general and natural assumption than a constant interaction strength -- lies in the recent discovery \\citep{Baldi_etal_2010} that the effects of the DE-CDM interaction on the formation and evolution of structures at small scales, in particular in the nonlinear regime, might help alleviating the tensions between the $\\Lambda $CDM model and a series of astrophysical observations. These range from \\eg the abundance of satellites in CDM halos \\citep{Navarro_Frenk_White_1995}, to the observed low baryon fraction in large galaxy clusters \\citep{Ettori_2003,Allen_etal_2004,Vikhlinin_etal_2006,LaRoque_etal_2006,McCarthy_etal_2007}, to the so called ``cusp-core\" problem for the density profiles of the CDM halos of dwarf galaxies \\citep{Moore_1994,Flores_Primack_1994,Simon_etal_2003}, of spiral galaxies \\citep{Navarro_steinmetz_2000,Salucci_Burkert_2000,Salucci_2000,Binney_Evans_2001}, and of galaxy clusters \\citep{Sand_etal_2002,Sand_etal_2004,Newman_etal_2009}, or to the so called ``Dark Flow\" problem \\citep{Watkins_etal_2008}. Furthermore, some new potential challenges to the $\\Lambda $CDM scenario have been recently reported based on the detection of very massive high-redshift clusters \\citep[see \\eg][]{Jee_etal_2009,Rosati_etal_2009} or on the observed dynamical properties of CDM halo satellites \\citep{Lee_Komatsu_2010,Lee_2010}. In their recent study \\citet{Baldi_etal_2010} showed by means of a series of high-resolution N-body simulations how the DE-CDM interaction -- for the case of constant coupling -- could alleviate some of these problems; in particular, it was shown how the interaction can reduce the ``cuspyness\" of massive CDM halos, thereby going in the right direction for a solution of the ``cusp-core\" problem. Nevertheless, the magnitude of this effect is strongly limited by the tight observational constraints on constant coupling models \\citep[as \\eg][]{Bean_etal_2008,LaVacca_etal_2009} that put a firm bound to the maximum allowed value of the coupling. It is therefore natural to speculate about the possibility that a time dependent coupling with a large value during the late stages of structure formation and a progressively smaller value at high redshift might increase significantly the impact of the new physics introduced by the interaction on \\eg the density profiles of CDM halos without perturbing the overall evolution of the Universe beyond the present observational limits. The present work constitutes the natural extension of the analysis done by \\citet{Baldi_etal_2010} to the case of time dependent couplings. In this paper we perform a complete numerical study of some quite general classes of coupling functions, starting from their background evolution up to the nonlinear regime of structure formation, and we present the first high-resolution hydrodynamical N-body simulations of structure formation in the context of interacting DE models with a time dependent coupling to date. The manuscript is organized as follows. In Section~\\ref{cde} we describe the main features of the coupled DE models under investigation with a particular stress on the differences with respect to the standard constant coupling models that have been widely studied in the literature. More specifically, in Sec.~\\ref{bkg} we discuss the background equations and in Sec.~\\ref{integration} we illustrate the numerical methods used to integrate such equations backwards in time. In Sec.~\\ref{obs} we discuss observational constraints on the coupled DE scenario and a possible way to use the bounds derived for constant coupling models to check the viability of the variable coupling cosmologies investigated in the present work. In Sec.~\\ref{prt} we study linear perturbations equations and we discuss the evolution of linear matter density fluctuations in variable coupling models.\\\\ In Sec.~\\ref{sim} we briefly summarize the numerical methods used in the N-body simulations, and in Sec.~\\ref{results} we present and discuss the results of our runs. Finally, in Sec.~\\ref{concl} we draw our conclusions. ", "conclusions": "\\label{concl} In the context of interacting DE models we have studied a few general classes of time evolution of the interaction strength between DE and CDM, generalizing the widely studied case of constant couplings to the more natural scenario of a variable coupling. Following the idea -- already discussed in previous works -- that a large value of the coupling could leave distinctive features in the properties of observable structures and even possibly alleviate the tensions between the $\\Lambda $CDM cosmological model and astrophysical observations at small scales, we have designed a few general forms of coupling functions $\\beta _{c}(\\phi )$ that grow in time, thereby having a significantly weaker impact on the overall background expansion of the Universe as compared to constant coupling models with the same interaction strength at $z=0$. In particular, we have investigated three classes of time evolution of the coupling, where the interaction strength is proportional either to a power of the cosmological scale factor $a(t)$, or to the fractional DE density $\\Omega _{\\phi }$, or to an exponential function of the DE scalar field $\\phi $. The first two classes are purely phenomenological parametrizations of the time evolution of the coupling, while the latter one represents a more physical situation where the interaction depends on the dynamical evolution of the scalar field.\\\\ We have performed a complete numerical analysis of the background evolution for these three different types of coupling functions by solving the full system of coupled dynamic equations in the presence of a variable coupling, generalizing previous works. Even in the absence of analytic solutions for variable coupling models, our numerical integrations allow to identify the main background features of these cosmologies. More specifically, we have shown that the first two phenomenological parametrizations of the coupling evolution mentioned above, due to the very fast decrease of the coupling with increasing redshift, do not present the so called ``$\\phi $MDE\" scaling solution typical of constant coupling models, and for what concerns their background evolution are practically indistinguishable from $\\Lambda $CDM, thereby suffering of the same level of fine tuning of the cosmological constant. On the contrary, the more physical form of a coupling that depends on the evolution of the scalar field shows a background evolution with an intermediate behavior between $\\Lambda $CDM and a standard ``$\\phi $MDE\" phase which can be still well reproduced by the usual analytic solution for the fractional DE density $\\Omega _{\\phi }$ during the ``$\\phi $MDE\" phase, once generalized to the case of growing couplings. We have therefore called this intermediate type of background evolution a ``Growing $\\phi $MDE\" phase.\\\\ We have then studied the evolution of linear perturbations within variable coupling models, pointing out the main differences arising in the perturbations equations due to the time dependence of the coupling. In particular, we have shown how in general a growing coupling function could induce instabilities in the growth of scalar perturbations at large scales, due to the presence of a negative effective mass term in the linear scalar field equation, and we have discussed under which conditions these instabilities can be avoided. We have then numerically computed the growth factor of matter density perturbations at subhorizon scales for a few selected models. Among these, the computed growth factors again clearly shows two distinct classes: on one side the phenomenological couplings present no significant differences in the growth of density perturbations with respect to $\\Lambda $CDM, except at very low redshifts, while on the other side the exponential coupling models have a faster growth of density fluctuations during most of the expansion history of the Universe.\\\\ The main focus of the present paper is on the effects of variable couplings on nonlinear structure formation. Exploiting the implementation of coupled DE cosmologies into the N-body code {\\small GADGET-2} developed for previous works, we have run high-resolution hydrodynamical N-body simulations for some selected cosmological models belonging to different classes of coupling evolution. For all these cosmologies initial conditions have been generated discarding possible early effects of the coupling based on the consideration that these effects were shown by previous studies to have a minor impact on the final properties of nonlinear structures at the present level of numerical resolution. We have also chosen to normalize all the cosmologies to the same amplitude of the large-scale power at the present time. Although this is a common choice, other conventions are equally valid and could lead to different predictions for the same cosmological models.\\\\ We have shown that the power spectrum of matter density fluctuations evolves in a strikingly different way in the two different types of models with respect to $\\Lambda $CDM. In particular, the phenomenological couplings have a very similar evolution to $\\Lambda $CDM until $z\\sim 0.5$, followed by a very fast growth of the power spectrum amplitude at scales below $k \\sim 1.0~h~\\text{Mpc}^{-1}$. On the contrary, the exponential models have a lower amplitude than $\\Lambda $CDM at all scales during most of the cosmic evolution, and catch up with the $\\Lambda $CDM power spectrum only at large scales at $z=0$, while at small scales they still show a significant lack of power also at the present time. We have then confirmed that also for variable couplings the scalar fifth-force acting only between CDM particles induces a bias in the amplitude of density fluctuations in baryons and CDM at all scales, as it happens for constant coupling models. The bias shows a clear scale dependence that develops very quickly as the couplings approaches their large values at low redshift, and the enhancement of this effect when moving from the linear regime of very large scales to smaller and progressively more nonlinear scales is found to be stronger than for the constant coupling models studied in previous works. This bias can be detected also in the very nonlinear regime characterizing the inner parts of collapsed objects, and has an impact on the total amount of baryons contained in massive halos that could therefore influence the determination of the baryon fraction from cluster measurements. We have therefore computed the evolution of the average baryon fraction within the virial radius $r_{200}$ for all the halos arising in the simulations of the different cosmological models, finding a generally stronger reduction of the baryon fraction as compared to constant couplings, with a decrease up to $\\sim 14 - 16 \\%$ with respect to $\\Lambda $CDM. We have computed the mass functions at different redshifts for all of our selected models, showing how at $z=0$ all the cosmologies have similar shapes and amplitudes of the mass function, with relative differences of the order of $\\sim 10\\%$. However, at higher redshifts the mass functions of the exponential coupling models show a clear excess of small halos and a strong lack of large halos with respect to $\\Lambda $CDM, consistently with the later onset of structure formation in these models. We have also computed the multiplicity function for all the models and compared it with the theoretical predictions according to the \\citet{Sheth_Tormen_1999} and \\citet{Jenkins_etal_2000} fitting formulae, evaluated with the appropriate growth factor for all the models, and with the standard value of the extrapolated linear overdenisty at collapse. We found good agreement between the multiplicity functions in our simulated cosmologies and the theoretical fitting functions, with the only exception of one of the scale factor dependent models at $z=0$. This discrepancy might suggest the need to reconsider the spherical collapse formalism in the presence of strongly variable couplings, will be investigated in future works. Nevertheless, the usual mass function fitting formulae were found to be fairly accurate also for most of our variable coupling models, generalizing previous results. Finally, we have investigated the effects of variable couplings on the halo density profiles. As for the case of constant coupling models, we find that they are still remarkably well fit in all the different cosmologies by the NFW formula. However -- in contrast with what happens within constant coupling models -- we also find that variable coupling cosmologies do not always show a decrease of the inner overdensity of halos with respect to the standard $\\Lambda $CDM case, but present opposite trends for the two different classes of coupling functions. The more realistic and physically motivated exponential coupling models show a significant decrease of the inner overdensity of halos with respect to $\\Lambda $CDM, while the phenomenological models show on the contrary a clear increase of the density in the central regions. This strikingly different behavior can be explained by considering which are the main physical mechanisms that can account for a modification of the equilibrium state of a collapsed halo in the context of our variable coupling models. As it was shown before for the case of constant couplings, the mass decrease and the extra friction term in the equation of motion of CDM particles can only induce an increase of the total energy of a virialized system, which therefore restores its virial equilibrium by slightly expanding. This effect is the source of the lower densities in the cores of CDM halos in the presence of constant couplings. However, if the coupling is changing in time, there is an additional mechanism coming into play: the total potential energy of the system decreases due to the increase of the effective gravitational constant as a consequence of the growing scalar fifth-force. In our models, the effective gravitational constant grows by $\\sim $30 - 75 \\% during the whole cosmic evolution, which determines a corresponding decrease of the gravitational potential energy of collapsed systems. This decrease of total energy then determines a contraction of the halos. Interstingly, we have shown that the two opposite behaviors found for the inner overdensity of nonlinear structures are determined by the background evolution of the coupled DE-CDM system: in the models that do not present a ``Growing $\\phi $MDE\" phase the friction term, which is the main driver of the expansion of halos in constant coupling models, is strongly suppressed and cannot counteract the contraction induced by the strong increase of the effective gravitational constant. On the contrary, the exponential coupling models, due to the presence of a ``Growing $\\phi $MDE\", have a still efficient friction term that balances and sometimes overcomes the effect of the potential energy decrease, thereby determining an overall expansion of the halos. We have therefore shown how the nonlinear behavior of matter particles at small scales, in the context of coupled DE models with time dependent couplings, is directly influenced by the cosmological background evolution of the scalar field, and how the presence of a ``Growing $\\phi $MDE\" phase is essential to determine whether halos will contract or expand in these cosmologies. These considerations apply also to halo concentrations, which are found to be higher with respect to $\\Lambda $CDM in the phenomenological models due to the absence of a ``Growing $\\phi $MDE\", and lower in the exponential coupling models.\\\\ In conclusion, we have performed a complete numerical study of interacting DE cosmologies for a few general types of time dependence of the DE-CDM coupling, concerning background evolution, linear perturbations evolution, and nonlinear structure formation. We have presented the first high-resolution hydrodynamical N-body simulations of structure formation in the context of variable coupling models to date. In our analysis, we have found that differently from the constant coupling case, halo density profiles and halo concentrations do not evolve in the same direction with respect to $\\Lambda $CDM for all types of coupling evolution. In particular, depending on the type of background evolution determined by the coupling function, density profiles can be less overdense and correspondingly less concentrated than in $\\Lambda $CDM, or vice versa. Furthermore, the growth of structures at large scales is also affected in a significantly different way according to the different types of coupling evolution. Finally, we find that the decrease of the halo baryon fraction already found for constant coupling models can be significantly enhanced in variable coupling cosmologies. Some of these effects alleviate tensions between astrophysical observations and the $\\Lambda $CDM cosmology at small scales, and arise in cosmological models that contrarily to the constant coupling scenarios are not in stark conflict with present observational constraints on the background evolution of the Universe even in the presence of a significant coupling strength at low redshifts. Therefore, cosmological models with time dependent couplings in the dark sector might represent -- for some specific forms of coupling evolution -- a viable alternative to the standard $\\Lambda $CDM concordance model." }, "1005/1005.2141_arXiv.txt": { "abstract": "We present detailed observations of the bright short-hard gamma-ray burst GRB~090510 made with the Gamma-ray Burst Monitor (GBM) and Large Area Telescope (LAT) on board the \\Fermi\\ observatory. GRB~090510 is the first burst detected by the LAT that shows strong evidence for a deviation from a Band spectral fitting function during the prompt emission phase. The time-integrated spectrum is fit by the sum of a Band function with $\\Epeak = 3.9\\pm 0.3$\\,MeV, which is the highest yet measured, and a hard power-law component with photon index $-1.62\\pm 0.03$ that dominates the emission below $\\approx$\\,20\\,keV and above $\\approx$\\,100\\,MeV. The onset of the high-energy spectral component appears to be delayed by $\\sim$\\,0.1\\,s with respect to the onset of a component well fit with a single Band function. A faint GBM pulse and a LAT photon are detected 0.5\\,s before the main pulse. During the prompt phase, the LAT detected a photon with energy $30.5^{+5.8}_{-2.6}$ GeV, the highest ever measured from a short GRB. Observation of this photon sets a minimum bulk outflow Lorentz factor, $\\Gamma\\ga$\\,1200, using simple $\\gamma\\gamma$ opacity arguments for this GRB at redshift $z = 0.903$ and a variability time scale on the order of tens of ms for the $\\approx$\\,100\\,keV--few MeV flux. Stricter high confidence estimates imply $\\Gamma \\ga 1000$ and still require that the outflows powering short GRBs are at least as highly relativistic as those of long duration GRBs. Implications of the temporal behavior and power-law shape of the additional component on synchrotron/synchrotron self-Compton (SSC), external-shock synchrotron, and hadronic models are considered. ", "introduction": "A difficulty in trying to understand gamma-ray bursts is that, at least in terms of the temporal structure of their emission, all GRBs differ. When the overall time scales of the emission are considered, however, a pattern does emerge. The durations of the $\\sim 100$\\,keV--MeV emission from GRBs form a bimodal distribution and hence are divided into two classes, namely the short and long duration bursts. The short bursts formally have durations $< 2\\,$s with typical values around $\\sim 0.2$\\,s, whereas the long bursts have a distribution that peaks around $\\sim 30$\\,s with a tail extending to several hundreds of seconds \\citep{Kouveliotou93}. Although the physics of their gamma-ray emission is not well understood, these two classes of bursts likely originate from distinct types of progenitor systems \\citep{Lee07, Woosley06}. Long duration bursts are thought to be produced by the core collapse of massive stars as evidenced by the direct association of several nearby GRBs ($z<0.3$) with SN Ib/c events \\citep{Woosley06}. Consistent with this, the afterglow counterparts of long duration GRBs tend to lie in star forming regions of low mass, irregular galaxies \\citep{Kocevski09}. By contrast, short duration GRBs have been associated with both early and late type host galaxies, in proportions that reflect the underlying field galaxy distribution \\citep{Berger2009}. In the prevailing model for short bursts, they are produced in merger events of a compact binary systems composed of two neutron stars or a neutron star/black hole pair and so would tend to originate from older stellar populations. With the launch and successful operation of the \\Fermi\\ Gamma-ray Space Telescope, a wider observational window has been opened through which a greater understanding of GRBs may be obtained. The Large Area Telescope aboard \\Fermi\\ provides significantly greater energy coverage (20 MeV to $>$\\,300\\,GeV), field-of-view ($2.4$\\,sr) and effective area ($8000$\\,cm$^2$ at 1\\,GeV) than its predecessor EGRET \\citep{2009ApJ...697.1071A}. Owing to its substantially lower deadtime (26\\,$\\mu$s vs 100\\,ms for EGRET), the LAT can probe the temporal structure of even the shortest GRBs. In addition, the LAT can localize GRBs with sufficiently high precision to enable follow-up observations by \\Swift\\ and ground-based observatories; and at energies $\\ga$\\,few\\,GeV, the LAT can distinguish GRB photons from background with little ambiguity. The Gamma-ray Burst Monitor, the other science instrument on \\Fermi, comprises an array of 12 NaI scintillators and two BGO scintillators and can detect gamma-rays from 8\\,keV to 40\\,MeV over the full unocculted sky. The combined capabilities of the LAT and GBM enable \\Fermi\\ to measure the spectral parameters of GRBs over seven decades in energy. For sufficiently bright bursts, time-resolved spectral analysis is possible over the entire energy range. The \\Fermi\\ observations of the short gamma-ray burst, GRB~090510, take full advantage of the GBM and LAT capabilities. The LAT emission shows temporal structure on time scales as short as 20\\,ms. In addition to the usual Band function component \\citep{Band:93}, spectral fits reveal a hard power-law component emerging in the LAT band 0.1\\,s after the onset of the main prompt emission in the GBM band. Moreover, a $\\approx 0.2$\\,s delay is observed between the brightening of the $\\approx 200$ MeV--GeV emission with respect to the strong count increases in the NaI and BGO. These behaviors present severe challenges for emission models of GRBs. A photon with energy $30.5_{-2.6}^{+5.8}$\\,GeV was detected by the LAT 0.829\\,s after the GBM trigger. This event arrived during the prompt phase and is temporally coincident with a sharp feature in the GBM and LAT light curves. Given this energy, the temporal structure of the burst light curve and the known distance to the burst, \\citet{Nature..GRB090510} have used this photon and its arrival time to set limits on a possible linear energy dependence of the propagation speed of photons due to Lorentz-invariance violation that would require a quantum-gravity mass scale significantly above the Planck mass. Similar to several of the long bursts seen by the LAT, GRB~090510 shows a high energy extended emission component that is detected by the LAT as late as 200\\,s after the GBM trigger. In the context of GRB outflow models, the properties of this GeV emission and the optical and X-ray afterglow observations by \\Swift\\ place significant constraints on possible internal and external shock models for the late time emission of this source \\citep{2009arXiv0909.0016G,2010ApJ...709L.146D,2009arXiv0911.4453C,2009ApJ...706L..33G,2009MNRAS.400L..75K}. In this paper, we report on and discuss the GBM and LAT observations of GRB~090510 during the prompt emission phase. In section~2, we give the basic observational details. In section~3, we present a timing analysis, including discussion of its designation as a short-hard GRB and a cross-correlation analysis between various energy bands to characterize any energy-dependent temporal lags. In section~4, we perform spectral analyses of both the time-integrated and the time-resolved data; and we demonstrate the significance of the additional power-law component and how the spectra evolves over the course of the burst. In section~5, we derive a lower limit on the bulk Lorentz factor given the variability time scales and observations of the highest energy photons, describe the constraints imposed on leptonic, hadronic and other models of the prompt emission, and discuss implications of the detection of the 30.5\\,GeV photon for models of the extragalactic background light (EBL). Finally, we summarize our results in section~6. ", "conclusions": "\\label{sec:discussion} The emergence of a distinct high-energy spectral component in the prompt-phase spectrum of GRB~090510 establishes that a hard emission component in addition to a Band component is found in the short hard class of GRBs. Hard power-law components are also found in three long-duration GRBs, namely GRB~090902B \\citep{Fermi_GRB090902B}, GRB~090926A \\citep{Fermi..GRB090926A}, and GRB~941017 \\citep{Gonzalez:03}. In GRB~090510, the LAT data show that the $\\gamma$-ray flux above 100 MeV brightens $\\approx 0.2$\\,s after the start of the bright phase of GBM emission. This behavior is similar, but on a shorter timescale, to the delayed onset in the long-duration GRB~080916C \\citep{2009Sci...323.1688A} and most other \\Fermi\\ GRBs, including GRB~081024B, the first short GRB observed with the LAT \\citep{2010ApJ...712..558A}. Furthermore, GRB~090510 displays $>$\\,100\\,MeV emission significantly extended beyond the duration of the GBM flux \\citep{2010ApJ...709L.146D}, as observed in other \\Fermi\\ GRBs and earlier from GRB 940217 with the EGRET instrument on the {\\it Compton Observatory} \\citep{Hurley:94}. These behaviors provide important constraints for high-energy emission models and could help answer whether the high-energy $\\gamma$ rays have a leptonic or hadronic origin. Though the afterglow radiation in both long and short GRBs is probably nonthermal synchrotron emission from an external shock \\citep{1998ApJ...497L..17S}, the situation is less clear in the prompt and early afterglow phases when the GRB engine is most powerful. This radiation could be from the thermal photosphere made by the powerful relativistic wind \\citep{2002ApJ...578..812M,2007ApJ...664L...1P}, from magnetic reconnection in Poynting-flux dominated outflows \\citep{2003astro.ph.12347L}, or from nonthermal leptonic emissions formed by internal or external shocks \\citep[e.g.,][]{2009MNRAS.400L..75K,2009arXiv0909.0016G,2009arXiv0911.4453C} in the relativistic jet of a GRB. \\citet{2010ApJ...709L.146D} describe and present models for the \\Swift\\ and \\Fermi\\ observations of GRB~090510 during the afterglow phase. Here we consider the implications of the prompt phase and early afterglow emission for GRB 090510. After deriving the minimum bulk Lorentz factor $\\Gamma_\\min$ and considering the various uncertainties that enter into this calculation, we use the observations to constrain leptonic synchrotron/SSC model and hadronic models of short duration GRBs. We do not discuss a thermal photospheric interpretation for the \\Fermi\\ results on GRB~090510. The photospheric interpretation overcomes the problem that the GBM spectra are harder than expected below $\\Epeak$ with the simplest synchrotron emission model (which is only the case at $\\gtrsim 2\\sigma$ in GRB~090510 during interval b; see Table~\\ref{tab:spectral fits}). Even if it explains much of the GBM emission, however, a different origin is needed for the separate hard spectral component observed at LAT energies. The coincident narrow spikes between the LAT all events and GBM lightcurves would not be easy to explain in a purely photospheric scenario, though Compton-scattered photospheric emission by internal shocked electrons could produce the coincident components \\citep{2010arXiv1002.2634T}. \\subsection{Lower Limit on the Bulk Lorentz Factor} The use of $\\gamma$-ray observations to constrain the bulk outflow speed of highly variable and energetic $\\gamma$-ray emission from GRBs has been studied by many authors \\citep[e.g.,][]{1997ApJ...491..663B,2001ApJ...555..540L,2004ApJ...613.1072R}. A detailed derivation involves an integration over the photon spectrum to calculate the opacity of $\\gamma$ rays emitted from sources with idealized geometries \\citep[see Supplementary Information in][]{2009Sci...323.1688A}. A $\\delta$-function approximation for the $\\gamma\\gamma$ opacity constraint gives values of $\\Gamma_\\min$ accurate to $\\sim 10$\\% whenever the target photon spectrum is softer than $\\nu F_\\nu \\propto \\nu$. In this case, $\\gamma\\gamma$ opacity arguments for a $\\gamma$-ray photon with energy $\\e_1$, in $m_ec^2$ units, imply a minimum bulk Lorentz factor, defined by $\\tau_{\\gamma\\gamma}(\\e_1) = 1$, of \\citep[e.g.][]{1995MNRAS.273..583D,2007PhR...442..166N}, \\begin{equation} \\Gamma_{\\rm min} \\cong \\left[ {\\sigma_{\\rm T} d_L^2 (1+z)^2 f_{\\hat \\e}\\e_1\\over 4t_v m_ec^4}\\right]^{1/6}\\;\\;,\\;\\; \\hat \\e = {2\\Gamma^2\\over (1+z)^2\\e_1}\\;. \\label{eqs2} \\end{equation} Here $f_\\e $ is the $\\nu F_\\nu$ flux at photon energy $m_ec^2 \\epsilon$, which is evaluated at $\\e = \\hat \\epsilon$ due to the peaking of the $\\gamma\\gamma$ cross section near threshold. While the local value of the photon index around $\\hat\\epsilon$ has some effect on the exact numerical coefficient, this effect is small provided that the target photon index is $< -1/2$. Because of the threshold condition used to relate the high-energy photon and the target photons, the solution to equation~(\\ref{eqs2}) is iterative but quickly converges. We use this expression to estimate $\\Gamma_\\min$ from \\Fermi\\ observations of GRB 090510 for comparison with more accurate calculations. For interval b, during which a 3.4\\,GeV photon was detected, spectral analysis of GBM and LAT data during this episode reveals distinct Band-function and power-law components (Figure~\\ref{fig:nuFnu spectra}). The Band function has $\\Epeak = 5.1$ MeV, $\\alpha = -0.48$ and $\\beta = -3.09$ (Table~\\ref{tab:spectral fits}). The combined Band plus power-law fit reaches a peak $\\nu F_\\nu$ flux of $\\approx 4\\times 10^{-5}$ erg cm$^{-2}$ s$^{-1}$. Writing the variability timescale $t_v$(s)$=0.01 t_{-2}$ s for 10 ms variability timescale, and $f_{\\hat \\e} = 10^{-5} f_{-5}$ erg cm$^{-2}$ s$^{-1}$, then equation~(\\ref{eqs2}) gives $\\Gamma_\\min \\cong 1100(f_{-5}/t_{-2})^{1/6}\\equiv 10^3\\Gamma_3$ with $\\epsilon_1 = 3400/0.511 \\cong 6650$. The target photon energy $\\hat \\e \\cong 2\\Gamma_\\min^2/(1+z)^2\\e_1 \\cong 110 (f_{-5}/t_{-2})^{1/3}$, or $\\approx 50$ MeV, corresponding to the Band $\\beta$ branch of the function. Depending on whether the 3.4\\,GeV photon is interacting with the total emission or just the photons in the power law, then $f_{-5} \\approx 0.7$ or $f_{-5}\\approx 0.1$, and $\\Gamma_\\min \\cong 950$ or $\\Gamma_\\min \\cong 720$, respectively. For interval c from \\interval{0.8}{0.9}, the same procedure with the 30.5\\,GeV photon gives $\\Gamma_\\min \\cong 1370$ or 1060 for $f_{-5} \\approx 0.5$ or $f_{-5}\\approx 0.1$ corresponding respectively to the combined Band plus power-law fit or the power-law component only. The results of numerical integrations to determine $\\Gamma_\\min$ using the more detailed expressions in \\citet{2009Sci...323.1688A} are shown in Table~\\ref{tab:Gamma_min_results}. As can be seen, the simple estimates given above are in good agreement with the detailed calculation. A number of issues arise in the use of equation~(\\ref{eqs2}) or the numerical integrations that are important for assessing the value and uncertainty in $\\Gamma_\\min$. These include the error incurred by the uncertainties in source spectral fitting parameters, which properly involves a covariance matrix to correlate uncertainties for different parameters of the Band-function and power-law fit. For $E_{\\rm max}$, we take the highest energy photon associated with that pulse. Table~\\ref{tab:Gamma_min_results} presents the values for $t_v$, $E_{\\rm max}$, and $\\Gamma_\\min$ for time intervals \\interval{0.6}{0.8} and \\interval{0.8}{0.9}, which are the only two for which a distinct pulse width could be measured. In interval \\interval{0.6}{0.8}, the Band + PL fit shown in Table~\\ref{tab:spectral fits} form the target photon spectrum, while in interval \\interval{0.8}{0.9}, we present results for both the Band and Band + PL fits since each fits the data reasonably well. We also give values for $\\Gamma_\\min$ assuming that only the hard power-law component forms the target photon source. Here we assume that $t_v$ is the same as that measured from the BGO emission, which is primarily associated with photons in the Band portion of the spectrum. If the variability timescale of the power-law emission is different than assumed, which would be compatible with the two components originating from different locations, then the minimum Doppler factor limit would change as indicated by equation~(\\ref{eqs2}). Furthermore, for calculations of $\\Gamma_\\min$ we use the spectrum derived on 0.2 s (time interval b) and 0.1 s (time interval c) timescales rather than on the shorter variability timescales during which the high-energy photons are measured. This is required for accurate spectral analysis, but could underestimate the flux (and therefore $\\Gamma_\\min$) during the bright narrow spikes, as can be seen from Fig. \\ref{fig:light curves}. The derivation of $\\Gamma_\\min$ depends crucially on the assumption that the high-energy radiation and target photons are made in the same emitting region. Correlated variability between different wavebands would support the cospatial assumption (see Figure~\\ref{fig:light curves}), but no strong evidence for this behavior was found in the \\ccf\\ analysis described in Section~3. A conservative assumption would be to suppose that the high-energy photon is part of the power-law component and that it can potentially interact only with target photons that are part of the same power-law emission component. Even in cases where the target MeV photons are made at smaller radii than the high-energy photons, or in different regions within the Doppler cone of the emitting surface, spacetime overlap will add to opacity, so this should represent the most conservative assumption. Further complicating the derivation of $\\Gamma_\\min$ is the assumed emitting geometry and the temporal evolution of the radiating plasma. For a blast-wave geometry, the precise value of $\\Gamma_\\min$ depends on whether high-energy photons are produced throughout the ``shell'' or from the inner edge of the ``shell,'' and on the dynamical behavior of the target photons \\citep{Granot:08}. Finally, a significant uncertainty on $\\Gamma_\\min$ can arise if the photon with observed energy $E_{\\rm max}$ is a random fluctuation of the underlying true spectrum that corresponds to $\\Gamma \\lesssim \\Gamma_\\min$ and $\\tau_{\\gamma\\gamma}(E_{\\rm max}) \\gtrsim 1$. The confidence we have on the value of $\\Gamma$ depends on the radiative transport and escape of $\\gamma$ rays from the emitting region. For interval c from \\interval{0.8}{0.9}, $\\Gamma_\\min = 1218 \\pm 61$ for the Band plus power-law fit (Table~\\ref{tab:Gamma_min_results}). Assuming that the intrinsic spectrum extrapolates as a power law to high-energies, a likelihood ratio test assuming an exponential escape probability gives $\\Gamma/\\Gamma_\\min = 0.96$, $0.88$, and $0.80$ and a spherical escape probability gives $\\Gamma/\\Gamma_\\min = 0.89$, $0.69$, and $0.49$ at the 1, 2, and $3\\sigma$ confidence levels, respectively. The presence of two photons with energies between 1 and 2 GeV in interval b and a 7 GeV photon in interval c reduces the likelihood that the highest energy photon is a fluctuation and can be used to independently estimate $\\Gamma_\\min$, giving a value $\\Gamma_\\min \\gtrsim 1000$. GRB 090510 is the second short GRB observed with LAT, after GRB~081024B \\citep{2010ApJ...712..558A}, but the first with a redshift, which is required to derive $\\Gamma_\\min$. The value of $\\Gamma_\\min \\cong 1200$--1300 for GRB 090510 is comparable to, and slightly larger than the values of $\\Gamma_\\min \\cong 900$ and $\\Gamma_\\min \\cong 1000$ derived for GRB 080916C \\citep{2009Sci...323.1688A} and GRB 090902B \\citep{Fermi_GRB090902B} using corresponding $\\gamma\\gamma$ opacity arguments. This has led to suggestions that the GRBs with the most luminous LAT emission are those with the largest bulk Lorentz factors \\citep{2009MNRAS.400L..75K,2009arXiv0909.0016G}. \\subsection{Models for the Prompt Radiation from GRB 090510} In addition to the requirement of bulk outflow Lorentz factors $\\Gamma \\gtrsim \\Gamma_\\min$, models for GRB 090510 should explain the $\\approx 0.2$\\,s delay of the onset of the $\\gtrsim 100$\\,MeV emission compared to the start of the main GBM emission at \\Time{0.5}, the appearance of a hard component, and the high-energy radiation extending to $\\approx$\\,\\Time{150}. \\subsubsection{Synchrotron/SSC Model} A standard GRB model for the prompt phase assumes that the keV--MeV emission is nonthermal synchrotron radiation from shock-accelerated electrons \\citep[e.g.,][]{1996PhRvL..76.3478T}. This emission is necessarily accompanied by SSC radiation. The SSC component is stronger for a large ratio of nonthermal electron to magnetic-field energy density, as expressed by the condition $\\epsilon_e \\gg \\epsilon_B$ \\citep{2001ApJ...548..787S,Zhang:01}, and also when the GRB has a lower bulk Lorentz factor for an external shock origin. Here $\\epsilon_e$ and $ \\epsilon_B$ are the fractions of shocked energy transferred to nonthermal lepton and magnetic-field energy, respectively. Lower bulk Lorentz factors that give stronger SSC components result in greater attenuation of high-energy $\\gamma$ rays from $\\gamma\\gamma$ pair-production processes, as discussed in Section 5.1. For generic ($E_{\\rm iso} \\sim 10^{52}$ erg, $\\Gamma_0\\sim 300, n\\sim 1$ cm$^{-3}$) parameters expected in an external shock model, a hardening of the LAT spectrum due to the deceleration of the blast wave and the emergence of the SSC component in the LAT band was expected to take place in the afterglow phase, but not in the prompt phase \\citep{Dermer:00b}. With the larger initial Lorentz factors $\\Gamma_0\\gtrsim 10^3$ implied by $\\gamma\\gamma$ arguments for LAT GRBs and the earlier emergence of an external shock component, an SSC component would persist after the decline of the synchrotron component due to less scattering taking place in the Klein-Nishina regime as the blast wave decelerates, and to the peak frequency of the SSC flux decreasing from the TeV to the GeV range. The detailed observations of GRB~090510 can help determine whether the hard power-law component appearing at $\\approx$\\,\\Time{0.7} can be explained by SSC emission in the LAT waveband during the prompt phase. Figure~\\ref{fig:dermerSSC} shows results for a numerical model where synchrotron peak energy, peak flux and variability time are made to correspond to the observed values shown in Figure~\\ref{fig:nuFnu spectra}. This code employs a Compton kernel that accurately treats Compton-scattering of relativistic electrons throughout the Thomson and Klein-Nishina regimes, internal $\\gamma\\gamma$ opacity, synchrotron self-absorption, radiative escape for $\\gamma$ rays described by exponential and spherical escape probabilities, and second-order SSC. The code does not include reprocessing of internally-absorbed radiation or effects of attenuation of $\\gamma$-ray photons by the EBL (see Section 5.3). Parameters appropriate to the Band component in interval b (see Figures~\\ref{fig:light curves} and~\\ref{fig:nuFnu spectra}) are used, as shown by the dotted curves. The parameters are $t_v = 14$\\,ms, $\\Epeak = 5.10$\\,MeV, $E_\\max = 3.4$\\,GeV, $\\alpha = -0.48$, $\\beta = -3.09$, and peak synchrotron flux $\\approx 4\\times 10^{-5}$\\,erg\\,cm$^{-2}$\\,s$^{-1}$ (see Figure~\\ref{fig:nuFnu spectra} and Tables~\\ref{tab:spectral fits} and~\\ref{tab:Gamma_min_results}). Figure~\\ref{fig:dermerSSC} shows results for $\\Gamma = 500$ (dashed curves) and $\\Gamma = 1000$ (solid curves), and for magnetic fields $B^\\prime = 1$\\,kG and $B^\\prime = 1$\\,MG in the upper and lower panels, respectively. Note that subtraction of an underlying hard component will not change the peak $\\nu F_\\nu$ flux value by more than $\\approx 5$\\%. The unattenuated power of the SSC component is comparable to the synchrotron power when $B^\\prime = 1$ kG, whereas the SSC component is much weaker in the strong-field case. The isotropic jet power for the $B^\\prime = 1$ kG model is $\\cong 2.0\\times 10^{54}$ erg s$^{-1}$ and $\\cong 5.8\\times 10^{53}$ erg s$^{-1}$ for $\\Gamma = 500$ and $\\Gamma = 1000$, respectively, and is dominated by the energy in the escaping radiation. When $B^\\prime = 1$ MG, the isotropic jet power is $\\cong 1.1\\times 10^{55}$ erg s$^{-1}$ and $\\cong 7\\times 10^{56}$ erg s$^{-1}$ for $\\Gamma = 500$ and $\\Gamma = 1000$, respectively, and is dominated by magnetic-field energy. For $B^\\prime \\ll$ 1 kG, the SSC flux becomes much brighter than the synchrotron flux, and the jet power becomes dominated by particles, even assuming that all particle energy is in the form of relativistic electrons. Because the SSC component is strongly attenuated by $\\gamma\\gamma$ processes for the $\\Gamma$ factors considered, an electromagnetic cascade will be formed with $\\gamma$ rays emerging at lower energies where the system becomes optically thin. The timescale for the electromagnetic radiation to cascade to energy $E_\\gamma$ is shorter than the synchrotron time, given by $t_{\\rm syn} \\approx 0.006 (\\Gamma/1000)^{1/2}/[(B^\\prime/{\\rm kG})^{3/2} \\sqrt{E_\\gamma/100{\\rm~MeV}}]$ s. Unless $B^\\prime\\ll 1$ kG, in which case much more power is found in the cascading SSC emission than in the synchrotron emission, the cascading timescale is too short to explain the $\\approx 0.2$ s delay between the GBM and LAT emission. This model also faces the well-known line-of-death problem \\citep{Preece:98} that the standard synchrotron mechanism makes a spectrum softer than $\\alpha = -2/3$, whereas $\\alpha = -0.48\\pm 0.07$ in interval b, representing a nearly $3\\sigma$ discrepancy from the hardest expected synchrotron emissivity. For the strong-field case where the SSC component is weak, the separate hard component in GRB 090510 would then have to originate from a different mechanism. \\subsubsection{Afterglow Synchrotron Model} \\citet{2009MNRAS.400L..75K,2009arXiv0910.5726K} and \\citet{2009arXiv0909.0016G,2009arXiv0910.2459G} have proposed forward shock emission from the early afterglow as the origin of the delayed onset and the hard component extending into the LAT energy band. This possibility is also considered by \\citet{2010ApJ...709L.146D}, \\citet{2009arXiv0911.4453C}, and \\citet{2009ApJ...706L..33G}. In particular, \\citet{2009arXiv0909.0016G} calculate the coasting bulk Lorentz factor of the GRB jet by identifying the time of the peak LAT emission occurring at $\\approx$\\,\\Time{0.7} with the deceleration time of a relativistic blast wave. If $\\eta$ is the efficiency to convert bulk kinetic energy to $\\gamma$-ray energy, $\\Gamma_0 \\approx 2000 ~n^{-1/8} (t_{\\rm peak}/0.2~\\rm s)^{-3/8} (\\eta/0.2)^{-1/8} (E_{\\gamma,\\rm iso}/3.5\\times 10^{52}~\\rm erg)^{1/8}$, where $n($cm$^{-3}$) is the density of the surrounding medium. This expression uses an apparent isotropic $\\gamma$-ray energy release that excludes the LAT emission, which if due to synchrotron radiation from relativistic electrons, would require a radiative efficiency approaching unity. Depending on circumburst density, the implied Lorentz factor is about 2--4 times the value of $\\Gamma_\\min$ calculated in the previous section. For this model, the emission radius corresponding to the time of the peak LAT flux is $R \\approx 2.4\\times 10^{16} n^{-1/4}$ cm. At $t\\approx t_{\\rm peak}$, the minimum energy electrons in the forward shock radiate synchrotron photons at energies \\\\ $h\\nu_m \\approx 3.6(\\epsilon_e/0.1)^2\\xi_e^{-2}(\\epsilon_B/0.01)^{1/2}E_{53}^{1/2}(t/0.2\\,{\\rm s})^{-3/2}\\;$MeV assuming an electron injection index of $p \\approx 2.4$, where $\\xi_e$ is the fraction of the electrons that take part in the non-thermal power-law component responsible for the observed emission. Recent particle in cell simulations of relativistic collisionless shocks \\citep{2008ApJ...673L..39S,2008ApJ...682L...5S,2009ApJ...695L.189M} suggest that $\\xi_e$ is fairly small (of the order of a few percent), which may in our case allow $h\\nu_m$ of several MeV even for low values of $\\epsilon_B$. For $\\epsilon_e = 0.1$, $\\epsilon_B = 0.01$, and an isotropic equivalent kinetic energy of $E_{k,{\\rm iso}} = 10^{53}\\;$erg (similar to $E_{\\rm\\gamma,iso}$), the cooling break frequency is given by \\citep{2002ApJ...568..820G}, $h\\nu_c \\approx 0.4 n^{-1} (\\epsilon_e/0.1)^{-1} (\\epsilon_B/0.01)^{-1/2} E_{53}^{-1/2} (t/0.2{\\rm~s})^{-1/2}$\\,keV. This expression holds when $\\epsilon_B \\ll \\epsilon_e$ and Klein-Nishina effects are unimportant, so that $Y = [(1+4\\epsilon_e/\\epsilon_B)^{1/2}-1]/2 \\approx \\sqrt{\\epsilon_e/\\epsilon_B} > 1$. This would imply fast cooling $\\nu_c \\ll \\nu_m$ for $n \\sim 1$, which could result in a highly radiative shock for $\\epsilon_e \\sim 1$. However, the late time broad band spectrum at $t \\sim 100\\;$s from optical-UV through X-ray and up to the LAT $\\gamma$-ray energies \\citep{2010ApJ...709L.146D} suggests $h\\nu_c(100\\,{\\rm s}) \\sim 300\\;$MeV, which, even if with a large uncertainty, together with the overall afterglow modeling suggest a much lower external density of $n \\sim 10^{-5}$. Such a low density would imply $h\\nu_c \\approx 0.3 (n/10^{-5})^{-1} (\\epsilon_B/0.01)^{-3/2} E_{53}^{-1/2} (t/0.5{\\rm~s})^{-1/2}$\\,GeV (where the Klein-Nishina effect suppresses SSC cooling), so that $\\nu_c$ passes through the (low part of the) LAT energy range around $\\sim 0.5\\;$s from the onset of the main emission episode, or at $\\sim T_0+1\\;$s, when there is a softening in the LAT photon index, near the end of the prompt emission. Thus, after the prompt emission, the LAT energy range would be above both $\\nu_m$ and $\\nu_c$, accounting for the observed photon index. This would imply $h\\nu_c \\approx 18 (n/10^{-5})^{-1} (\\epsilon_B/0.01)^{-3/2} E_{53}^{-1/2} (t/100{\\rm s})^{-1/2}$\\,MeV which is consistent with the broad band spectrum at that time, especially since the spectral break around $\\nu_c$ is very smooth and gradual \\citep{2002ApJ...568..820G}. Interestingly, the inferred value of $h\\nu_m(100\\,{\\rm s}) \\approx 0.43\\;$keV and the $\\nu_m \\propto t^{-3/2}$ scaling gives $h\\nu_m(t_{\\rm peak}\\approx 0.2\\,{\\rm s}) \\approx 4.8\\;$MeV, which is very close to the measured value of $\\Epeak$ near $t_{\\rm peak}$. This model would not produce a spectrum softer than $\\nu F_{\\nu} \\propto \\nu^{4/3}$, so would have difficulty accounting for the emission near 10\\,keV in interval b, which appears to be a continuation of the hard spectral component. For an adiabatic blast wave and an electron injection index $p$, the synchrotron flux scales as $\\nu F_\\nu\\propto t^{(2-3p)/4} \\nu^{(2-p)/2}$ at frequencies $\\nu > \\nu_m,\\nu_c$. The measured late time LAT flux decay rate of $\\nu F_\\nu \\propto t^{-1.38\\pm 0.07}\\nu^{-0.1\\pm 0.1}$ would in this case require $p = 2.5\\pm 0.1$, while the spectral slope requires $p = 2.2\\pm 0.2$. Both are consistent with $p = 2.4$ at the $1\\sigma$ level. This value is also consistent with the late time afterglow of GRB~090510 \\citep{2010ApJ...709L.146D}. A radiative blast wave at early times is not required in order to account for the observed early LAT flux decay rate. \\subsubsection{Hadronic Models} In hadronic models, photohadronic and proton/ion synchrotron processes induce electromagnetic cascades, which lead to synchrotron and Compton emissions from secondary electron-positron pairs \\citep[e.g.,][]{2006NJPh....8..122D, Gupta:07, 2007ApJ...671..645A}. For a target photon energy distribution $n(\\epsilon) \\propto \\epsilon^{x}$, the efficiency for photopion processes is $ \\propto R^{-1} \\Gamma^{-2} E_{\\rm p}^{-1-x} \\propto \\Gamma^{-4} t_{\\rm v}^{-1} E_{\\rm p}^{-1-x}$, where $R\\propto c\\Gamma^2 t_{\\rm v}$ is the shock radius. In this expression, protons with energy $E_{\\rm p}$ preferentially interact with photons with energies $\\propto \\Gamma^2/E_{\\rm p}$ \\citep{1995PhRvL..75..386W, 2009PhRvL.103h1102M}. The large deduced values for $\\Gamma$ in GRB~090510 make the photopion efficiency low, so that a very large energy release is required if the LAT radiation from GRB 090510 is assumed to be from a photomeson-induced cascade. A stronger magnetic field shortens the acceleration timescale, leading to a larger maximum particle energy $E_{\\rm max}$, thus enhancing the photopion production efficiency. In such a strong magnetic field, however, the effective injection index of secondary pairs tends to be about $-2$ \\citep{1992MNRAS.258..657C}, which yields a flat spectrum in a $\\nu F_\\nu$ plot, while the power-law index of the extra component in GRB~090510 is $\\sim -1.6$. In a weaker magnetic field, the Compton component from secondary pairs can harden the spectrum, though with reduced photopion production efficiency due to the smaller value of $E_{\\rm max}$. The slower cooling time of protons than electrons would produce an extended proton-induced emission feature \\citep{Bottcher:98}, though the SSC component would decay even more slowly \\citep{Zhang:01}. A recent numerical calculation by \\citet{2009ApJ...705L.191A} for GRB 090510 demonstrates that the proton injection isotropic-equivalent power is required to be larger than $10^{55}$\\,erg\\,s$^{-1}$ to explain the hard spectra of the extra component in GRB 090510, which is $\\approx 2$ orders of magnitude greater than the measured apparent isotropic $\\gamma$-ray luminosity. An alternative hadronic scenario is a proton synchrotron model with a very strong magnetic field \\citep{2009arXiv0908.0513R}. This model explains the delayed onset by proton synchrotron radiation in the prompt phase due to the time required to accelerate, accumulate, and cool the ultrarelativistic protons. For an onset times $t_{\\rm onset} \\approx 0.2$~s after the start of the GBM main emission at $T_0+0.5$~s, the required magnetic field is $B^\\prime \\approx 7.4\\times 10^5 ~(\\Gamma/1000)^{-1/3} (t_{\\rm onset}/0.1~\\rm s)^{-2/3} (E_\\gamma/100~\\rm MeV)^{-1/3} $~G in the shocked fluid frame. The corresponding total energy release, for a two-sided jet beaming factor $f_b \\approx 1.5\\times 10^{-4}(\\theta_j/1~{\\rm deg})^2$, is ${\\cal E}\\approx 1.3\\times 10^{54}(\\Gamma/1000)^{16/3} (t_{\\rm onset}/0.1~\\rm s)^{5/3} (E_\\gamma/100~\\rm MeV)^{-2/3} (\\theta_j/1~{\\rm deg})^2$~erg \\citep[see][for GRB 080916C]{2009ApJ...698L..98W}. Thus values of $\\Gamma \\lesssim 1000$ and narrow jet angles $\\approx 1^\\circ$ are required for a proton synchrotron scenario, and such strong beaming is not clearly found in the short hard class of GRBs \\citep{2007PhR...442..166N}. Proton-dominated GRB models for ultra-high energy cosmic rays therefore are plausible only with low values of $\\Gamma$ and narrow jet collimation to reduce the total energy. \\subsection{Implications for the Extragalactic Background Light} The EBL is dominated by direct starlight in the optical/ultraviolet and by stellar radiation that is reprocessed by dust in the infrared. The EBL is difficult to measure directly due to contamination by zodiacal and Galactic foreground light \\citep{2001ARA&A..39..249H}. For sources at sufficiently high redshifts, $\\gamma\\gamma$ absorption of high-energy $\\gamma$-rays by EBL photons can provide a means of constraining models of the EBL. Figure~\\ref{fig:taugg} shows the absorption optical depth, $\\tau_{\\gamma\\gamma}$, for various models of the EBL as a function of $\\gamma$-ray energy at the redshift $z=0.903$ of GRB 090510. We have included curves for the two models of \\citet{2006ApJ...648..774S} as well as the fiducial model of \\citet{2009arXiv0905.1144G}, the best fit model of \\citet{2004A&A...413..807K}, the model by \\citet{2008A&A...487..837F}, and ``Model C'' of \\citet{2009arXiv0905.1115F}. The models of \\citet{2006ApJ...648..774S} border on optically thick at 30.5\\,GeV; all other models considered here give a transmission probability of $e^{-\\tau_{\\gamma\\gamma}}\\gtrsim 0.85$. The baseline and fast evolution models of \\cite{2006ApJ...648..774S} give transmission probabilities of 0.37 and 0.30, respectively. Although a higher energy ($\\approx 30.5$\\,GeV) photon was found from this burst than for GRB 080916C ($13$\\,GeV; \\cite{2009Sci...323.1688A}), that burst was more constraining for EBL models due to its higher redshift, $z = 4.35 \\pm 0.15$ \\citep{2009A&A...498...89G}. However, the EBL does evolve with redshift, and this GRB provides an independent constraint from a later time and at a closer distance than GRB 080916C. The low optical depth for the highest energy photons in GRB~090510 justify neglecting EBL effects in Figure~\\ref{fig:dermerSSC}." }, "1005/1005.0144_arXiv.txt": { "abstract": "{The observed relation between the X-ray radiation from active galactic nuclei, originating in the corona, and the optical/UV radiation from the disk is usually described by the anticorrelation between the UV to X-ray slope $\\alpha_{ox}$ and the UV luminosity. Many factors can affect this relation, including: i) enhanced X-ray emission associated with the jets of radio-loud AGNs, ii) X-ray absorption associated with the UV broad absorption line (BAL) outflows, iii) other X-ray absorption not associated with BALs, iv) intrinsic X-ray weakness, v) UV and X-ray variability, and non-simultaneity of UV and X-ray observations. The separation of these effects provides information about the intrinsic $\\alpha_{ox}-L_{UV}$ relation and its dispersion, constraining models of disk-corona coupling.} {We use simultaneous UV/X-ray observations to remove the influence of non-simultaneous measurements from the $\\alpha_{ox}-L_{UV}$ relation.} {We extract simultaneous data from the second XMM-Newton serendipitous source catalogue (XMMSSC) and the XMM-Newton Optical Monitor Serendipitous UV Source Survey catalogue (XMMOMSUSS), and derive the single-epoch $\\alpha_{ox}$ indices. We use ensemble structure functions to analyse multi-epoch data.} {We confirm the anticorrelation of $\\alpha_{ox}$ with $L_{UV}$, and do not find any evidence of a dependence of $\\alpha_{ox}$ on $z$. The dispersion in our simultaneous data ($\\sigma\\sim 0.12$) is not significantly smaller than in previous non-simultaneous studies, suggesting that ``artificial $\\alpha_{ox}$ variability'' introduced by non-simultaneity is not the main cause of dispersion. ``Intrinsic $\\alpha_{ox}$ variability'' , i.e., the true variability of the X-ray to optical ratio, is instead important, and accounts for $\\sim 30\\%$ of the total variance, or more. ``Inter-source dispersion\", due to intrinsic differences in the average $\\alpha_{ox}$ values from source to source, is also important. The dispersion introduced by variability is mostly caused by the long timescale variations, which are expected to be driven by the optical variations.} {} ", "introduction": "The relationship between the X-ray and optical/UV luminosity of active galactic nuclei (AGNs) is usually described in terms of the index $\\alpha_{ox}=0.3838 \\log(L_X/L_{UV})$, i.e., the slope of a hypothetical power law between 2500 \\AA\\ and 2 keV rest-frame frequencies. The X-ray and UV monochromatic luminosities are correlated over 5 decades as $L_X \\propto L_{UV}^k$, with $k\\sim0.5-0.7$, and this provides an anticorrelation $\\alpha_{ox} = a \\log L_{UV}$ + const, with $-0.2\\la a\\la -0.1$ \\citep[e.g.,][]{avni86,vign03,stra05,stef06,just07,gibs08}. One of the main results of these analyses is that QSOs are universally X-ray luminous and that X-ray weak QSOs are very rare \\citep[e.g.,][]{avni86,gibs08}, but it is not yet known if the same is true for moderate luminosity AGNs. UV photons are generally believed to be radiated from the QSO accretion disk, while X-rays are supposed to originate in a hot coronal gas of unknown geometry and disk-covering fraction. The X-ray/UV ratio provides information about the balance between the accretion disk and the corona, which is not yet understood in detail. The $\\alpha_{ox}-L_{UV}$ anticorrelation implies that AGNs redistribute their energy in the UV and X-ray bands depending on the overall luminosity, with more luminous AGNs emitting fewer X-rays per unit UV luminosity than less luminous AGNs \\citep{stra05}. It has been proposed that the anticorrelation can be caused by the larger dispersion in the luminosities in the UV than the X-ray band for a population with intrinsically uniform $\\alpha_{ox}$ \\citep{la-f95,yuan98}; however, more recent analyses based on samples with wider luminosity ranges confirm the reality of the relationship \\citep{stra05}. \\citet{gibs08} stressed the quite large scatter in the X-ray brightness of individual sources about the average relation and investigated the possible causes of the dispersion. Part of this scatter, usually removed \\citep[e.g.,][]{stra05,stef06,just07,gibs08} is caused by radio-loud quasars, which are relatively X-ray bright because of the enhanced X-ray emission associated with their jets \\citep[e.g.,][]{worr87}, and to broad absorption line (BAL) quasars, which are relatively X-ray faint \\citep[e.g.,][]{bran00} due to X-ray absorption associated with the UV BAL outflows. Additional causes of deviations from the average $\\alpha_{ox}-L_{UV}$ relation include: i) X-ray absorption not associated with BALs, ii) intrinsic X-ray weakness, and iii) UV and X-ray variability, possibly in association with non-simultaneous UV and X-ray observations. In particular, \\citet{gibs08} estimate that variability may be responsible for 70\\%-100\\% of the $\\alpha_{ox}$ dispersion, and that a few percent ($<2$\\%) of all quasars are intrinsically X-ray weak by a factor of 10, compared to the average value at the same UV luminosity. A large fraction of intrinsically X-ray weak sources would suggest that coronae may frequently be absent or disrupted in QSOs. An extreme case is PHL 1811, which is X-ray weak by a factor $\\sim$70, studied in detail by \\citet{leig07}, who propose various scenarios, including disk/corona coupling by means of magnetic reconnections, Compton cooling of the corona by unusually soft optical/UV spectrum, and the photon trapping of X-ray photons and their advection to the black hole. The influence of variability on the $\\alpha_{ox}-L_{UV}$ relation can be divided into two different effects: i) non-simultaneity of X-ray and UV measurements, which we call ``artificial $\\alpha_{ox}$ variability'', and ii) true variability in the X-ray/UV ratio, which we refer to as ``intrinsic $\\alpha_{ox}$ variability''. It is beneficial to analyse simultaneously acquired X-ray and UV data to eliminate the effect of the artificial variability and search for the intrinsic X-ray/UV ratio and/or its variability. On a rest-frame timescale of a few years, the optical/UV variability of QSOs has been estimated to be $\\sim$30\\% \\citep[e.g.,][]{gial91,vand04}, while X-ray variability has been estimated to be $\\sim$40\\% for Seyfert 1 AGNs \\citep{mark03}. On intermediate timescales, the relation between X-ray and optical/UV variability may be due to either: i) the reprocessing of X-rays into thermal optical emission, by means of irradiation and heating of the accretion disk, or ii) Compton up-scattering, in the hot corona, of optical photons emitted by the disk. In the former case, variations in the X-ray flux would lead optical/UV ones, and vice versa in the latter case. Cross-correlation analyses of X-ray and optical/UV light curves allow us to constrain models for the origin of variability. The main results obtained so far, on the basis of simultaneous X-ray and optical observations, indicate a cross-correlation between X-ray and UV/optical variation on the timescale of days, and in some cases delays of the UV ranging from 0.5 to 2 days have been measured \\citep{smit07}. Simultaneous X-ray/UV data can be obtained by the XMM-Newton satellite, which carries the co-aligned Optical Monitor (OM). The second XMM-Newton serendipitous source catalogue (XMMSSC) \\citep{wats09} is available online in the updated incremental version 2XMMi\\footnote{http://heasarc.gsfc.nasa.gov/W3Browse/xmm-newton/xmmssc.html}. The XMM-Newton Optical Monitor Serendipitous UV Source Survey catalogue (XMMOMSUSS) is also available online\\footnote{http://heasarc.gsfc.nasa.gov/W3Browse/xmm-newton/xmmomsuss.html}. We look for simultaneous measurements of the $\\alpha_{ox}$ index from XMM/OM catalogues, to provide at least partial answers to the following questions: how large is the effect of non-simultaneous X-ray/UV observations on the dispersion about the average $\\alpha_{ox}-L_{UV}$ relationship? Is there any spectral X-UV variability for individual objects? Do their $\\alpha_{ox}$ harden in the bright phases or vice versa? Which constraints do these measurements place on the relationship between the accretion disk and the corona? The paper is organised as follows. Section 2 describes the data extracted from the archival catalogues. Section 3 describes the SEDs of the sources and the evaluation of the specific UV and X-ray luminosities. Section 4 discusses the $\\alpha_{ox}-L_{UV}$ anticorrelation and its dispersion. In Sect. 5, we present the multi-epoch data and discuss the intrinsic X/UV variability of individual sources. Section 6 provides notes about individual peculiar sources. Section 7 discusses and summarises the results. Throughout the paper, we adopt the cosmology H$_{o}$=70 km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{m}$=0.3, and $\\Omega_{\\Lambda}$=0.7. ", "conclusions": "The behaviour of $\\alpha_{ox}$, i.e., its dependence on luminosity and redshift, its dispersion and variability, are to be considered as symptoms of the relation between disk and corona emissions and their variabilities. It is generally believed that variable X-ray irradiation can drive optical variations by means of variable heating of the internal parts of the disk on relatively short timescales, days to weeks, while intrinsic disk instabilities in the outer parts of the disk dominate on longer timescales, months to years, propagating inwards and modulating X-ray variations in terms of Compton up-scattering in the corona \\citep{czer04,arev06,arev09,papa08,mcha10}. The structure functions of the light curves increase on long timescales both in the optical \\citep[e.g.,][]{dicl96,vand04,baue09} and X-rays \\citep[e.g.,][Vagnetti et al., in prep.]{fior98}. This, however, does not imply that the $\\alpha_{ox}$ SF also increases with time lag. Larger changes (on long time lags) in both X-ray and UV fluxes may occur without changes in the spectral shape (i.e., with constant $\\alpha_{ox}$). Our results shown in Fig. 9 indicate that this is not the case, i.e., that slope changes are indeed larger on longer timescales. Moreover, it is evident from Fig. 9 that most of the dispersion about the $\\alpha_{ox}-L_{UV}$ relation is due, in the present sample, to variations on timescales from months to years, which are associated with optically driven variations, according to the general belief. The $\\alpha_{ox}$ structure function does not distinguish between the hardening or softening of the optical to X-ray spectrum during brightening. This is instead described by the spectral variability parameter $\\beta=\\partial\\alpha/\\partial\\log F$ \\citep{trev01,trev02}, which can be adapted to the optical-X-ray case to become \\begin{equation} \\beta_{ox}={\\delta\\alpha_{ox}\\over\\delta\\log L_{UV}}\\,, \\end{equation} \\noindent where $\\beta_{ox}$ is the slope of the correlated variations $\\delta\\alpha_{ox}$ and $\\delta L_{UV}$, and describes whether a source hardens when it brightens or vice versa, i.e., if the X-ray luminosity increases more than the optical or less. For example, single source variations parallel to the $\\alpha_{ox}$ anticorrelation have a negative $\\beta_{ox}$, while variations perpendicular to the correlation have $\\beta_{ox}>0$. Both of these behaviours can be seen in Fig. 7. Of course, the different behaviour of the sources in the $\\alpha_{ox}-L_{UV}$ plane may correspond to a different time sampling. Constraining physical models of the primary variability source and disk-corona coupling would require the analysis of $\\beta_{ox}$ as a function of the time lag. This analysis does not look feasible, i.e., to have statistical reliability, with the present sparse sampling. We can propose more conventional scenarios and note that since most of the variability, in the present sample, occurs on long ($\\sim 1$ year ) timescale, it is presumably associated with optically driven variations. Considering all the measured variations $\\delta\\alpha_{ox}$ and $\\delta L_{UV}$, we obtain the ``ensemble\" average $\\langle\\beta_{ox}\\rangle=-0.240$. The negative sign implies that, on average, a spectral steepening occurs in the brighter phase. This is, in fact, consistent with larger variations in the UV band, driving the X-ray variability. The value of $\\langle\\beta_{ox}\\rangle$ can be compared with the average slope of the $\\alpha_{ox}-L_{UV}$ relation, Eq. (13), indicating that the UV excess in the brighter phase (steepening) is larger than the average UV excess in bright objects respect to faint ones. Finally we emphasise that, despite its limitations, the present analysis illustrates the feasibility of an ensemble analysis of the $\\alpha_{ox}-L_{UV}$ correlation, e.g., by considering the $\\beta$ parameter as a function of time lag. What is presently missing is an adequate simultaneous X-ray-UV sampling, at relatively short time lags, of a statistical AGN sample. An ensemble analysis may provide important constraints even when the total number of observations does not allow us to carry out a cross-correlation analysis of X-ray and UV variations of individual sources. We summarise our main results as follows: \\begin{itemize} \\item we have studied the $\\alpha_{ox}-L_{UV}$ anticorrelation with simultaneous data extracted from the XMM-Newton Serendipitous Source catalogues; \\item we confirm the anticorrelation, with a slope (-0.178) slightly steeper w.r.t. \\citet{just07}; \\item we do not find evidence for a dependence of $\\alpha_{ox}$ on redshift, in agreement with previous authors \\citep[e.g.][]{avni86,stra05,stef06,just07}; \\item there appears to be a flatter slope to the anticorrelation at low luminosities and low redshifts, in agreement with previous results by \\citet{stef06}; \\item the dispersion in our simultaneous data ($\\sigma\\sim 0.12$) is not significantly smaller w.r.t. previous non-simultaneous studies \\citep{stra05,just07,gibs08}, indicating that ``artificial $\\alpha_{ox}$ variability'' introduced by non-simultaneity is not the main cause of dispersion; \\item ``intrinsic $\\alpha_{ox}$ variability'' , i.e., true variability in the X-ray to optical ratio, is important, and accounts for $\\sim 30\\%$ of the total variance, or more; \\item ``inter-source dispersion\", due to intrinsic differences in the average $\\alpha_{ox}$ values from source to source, is also important; \\item the dispersion introduced by variability is mostly caused by the long timescale variations, which are expected to be dominated by the optical variations; the average spectral softening observed in the bright phase is consistent with this view; \\item distinguishing the trends produced by optical or X-ray variations may be achievable using the ensemble analysis of the spectral variability parameter $\\beta_{ox}$ as a function of time lag; crucial information would be provided by wide field simultaneous UV and X-ray observations with relatively short (days-weeks) time lags. \\end{itemize}" }, "1005/1005.2363_arXiv.txt": { "abstract": "We present manganese abundances in 10 red-giant members of the globular cluster $\\omega$ Centauri; 8 stars are from the most metal-poor population (RGB MP and RGB MInt1) while two targets are members of the more metal rich groups (RGB MInt2 and MInt3). This is the first time Mn abundances have been studied in this peculiar stellar system. The LTE values of [Mn/Fe] in $\\omega$ Cen overlap those of Milky Way stars in the metal poor $\\omega$ Cen populations ([Fe/H]$\\sim$-1.5 to -1.8), however unlike what is observed in Milky Way halo and disk stars, [Mn/Fe] declines in the two more metal-rich RGB MInt2 and MInt3 targets. Non-LTE calculations were carried out in order to derive corrections to the LTE Mn abundances. The non-LTE results for $\\omega$ Cen in comparison with the non-LTE [Mn/Fe] versus [Fe/H] trend obtained for the Milky Way confirm and strengthen the conclusion that the manganese behavior in $\\omega$ Cen is distinct. These results suggest that low-metallicity supernovae (with metallicities $\\le$ -2) of either Type II or Type Ia dominated the enrichment of the more metal-rich stars in $\\omega$ Cen. The dominance of low-metallicity stars in the chemical evolution of $\\omega$ Cen has been noted previously in the s-process elements where enrichment from metal-poor AGB stars is indicated. In addition, copper, which also has metallicity dependent yields, exhibits lower values of [Cu/Fe] in the RGB MInt2 and MInt3 $\\omega$ Cen populations. ", "introduction": "The abundance ratios of chemical elements which have different nucleosynthetic origins can be used to constrain the history of chemical evolution in different types of stellar populations. This is because stellar sources, such as massive stars that undergo core-collapse and explode as supernovae of Types II or Ib/Ic, or binary systems that become Type Ia supernovae (SNe Ia), or asymptotic giant branch (AGB) stars, produce and return to the interstellar medium different element abundance ratios on differing timescales. Abundance ratios of certain elements can thus be used to explore which stars and in which proportions have contributed to chemical evolution within stellar populations in galaxies. Of particular interest to chemical evolution of a stellar system are those elements whose yields may depend on the metallicity of the progenitor star. Metallicity-dependent abundance patterns can retain `memory' of the metallicity distributions of their stellar populations, regardless of how the overall metallicity of the parent galaxy evolves. One important element whose production may be metallicity dependent is manganese (Z=25), which falls within the iron peak. Manganese is produced in both core collapse supernovae (CC SNe) and SNe Ia, but the relative amounts from these two sources are not well constrained. Manganese has one stable isotope, $^{55}$Mn; this element is synthesized as a result of explosive incomplete silicon burning (see, e.g., Thielemann et al. 2007 for a more complete discussion). Woosley \\& Weaver (1995), for example, model nucleosynthesis of massive stars exploding as supernovae of Type II (SNe II) and find increasing Mn yields with increasing metallicity. Convolving the Mn yields from the Woosley \\& Weaver (1995) models with a Salpeter mass function leads to values of [Mn/Fe] that decline steadily to -0.3 at [Fe/H]=-1 and then to -0.5 at [Fe/H]=-2. A decrease of [Mn/Fe] ratios with metallicity is reported in the LTE Mn I abundance analyses of Galactic stars such as Gratton (1989), Reddy et al. (2003; 2006), Johnson (2002), Cayrel et al. (2004), or Fetzing et al. (2007). Recent results from Bergemann \\& Gehren (2008), however, indicate that the LTE approximation underestimates the abundances based on Mn I lines by 0.1 - 0.4 dex, with non-LTE effects being most pronounced in stellar atmospheres with low metal content. In the metallicity range investigated by Bergemann \\& Gehren (2008), -2.5 $<$ [Fe/H] $<$ 0, the non-LTE [Mn/Fe] ratios in late-type stars are approximately solar. The flat [Mn/Fe] trend with [Fe/H] is in general agreement with the SNe II yields of Chieffi \\& Limongi (2004) who do not predict a strong depletion of Mn relative to Fe at low metallicities. Manganese yields from SNe Ia models can span a range of values depending on the explosion mechanism, such as slow or fast deflagrations, or deflagration-detonation events. Iwamoto et al. (1999) investigate a variety of models and find no large differences in Mn/Fe yields (i.e., less than a factor of $\\sim$2) from models driven by both slow and fast deflagrations, as well as deflagration-detonations, having different initial metallicities. Badenes et al. (2008) computed nucleosynthesis from 4 delayed detonation models, with different metallicities, as well as one deflagration model. This group finds that Mn yields decline with decreasing metallicity, despite the differences in the explosion mechanisms and initial conditions of the models. In addition to the results for the Galactic thin disk, thick disk and halo noted above, McWilliam et al. (2003) added two other populations to the studies of manganese by measuring LTE [Mn/Fe] abundances for stellar members of the Sagittarius dwarf spheroidal galaxy and stars from the Galactic bulge. Both systems exhibited somewhat different behaviors of [Mn/Fe] with [Fe/H], with the Sgr dwarf galaxy stars having values of [Mn/Fe] that fall below the general LTE trend found for Milky Way disk or halo stars at a given [Fe/H]. Bulge stars exhibit opposite behavior, with their values of [Mn/Fe] falling above the trend defined by the Milky Way disk and halo stars. The goal of this study is to add another distinct stellar population to analyses of manganese and its chemical evolution in different Galactic environments. $\\omega$ Cen exhibits some peculiar characteristics in the nature of its chemical evolution, with perhaps the most striking being the large increase in the abundances of the heavy s-process elements (such as Ba or La) as the overall metallicity of cluster stars, measured by such elements as Fe, Ca, or Ti, increases (e.g. Norris \\& Da Costa 1995). $\\omega$ Centauri, although historically classified as a globular cluster, is now thought possibly to be a surviving remnant of a captured small galaxy, with multiple populations spanning a large range in metallicity (for a more detailed discussion see the review by Smith 2004). Recently, Carretta et al. (2010) have pointed out similarities between the $\\omega$ Cen populations and those from the Sagittarius dwarf galaxy. Five distinct stellar populations each with a different metallicity have been identified by Pancino et al. (2000) and Sollima et al. (2005). These studies label the distinct red giant branches from the $\\omega$ Cen populations as RGB MP (metal-poor); RGB MInt1; RGB MInt2; RGB MInt3 (intermediate metallicities); and RGB-a (anomolous, with the highest metallicity). Manganese abundances are presented here for the first time in $\\omega$ Cen stars, with the sample consisting of 10 targets; 8 red giants are from the most metal poor populations (RGB MP and RGB MInt1) and 2 stars are from the more metal rich RGB MInt2 and MInt3. These $\\omega$ Cen stars have been analyzed in previous studies (Smith et al. 2000; Cunha et al. 2002), however Mn was not included in the analysis. ", "conclusions": "Manganese abundances have been measured for the first time in the peculiar globular cluster $\\omega$ Cen, with the analysis of 10 red giants spanning a range in metallicity from [Fe/H]= -1.9 to -0.9. The analysis is based on Mn I lines using both LTE and non-LTE calculations to derive abundances. In addition, the possible effects of enhanced He abundances on derived Mn abundances were investigated for the more metal-rich $\\omega$ Cen giants and were found to be negligible. The novel result is that two members from the more metal-rich populations with $\\omega$ Cen (RGB MInt2 and MInt3) exhibit low ratios of Mn/Fe in comparison to Galactic field stars, as well as other globular cluster stars at the same metallcitiy ([Fe/H]$\\sim$-1). Differences between $\\omega$ Cen and the other Milky Way populations exist whether the comparison is made using LTE or non-LTE abundances. The low abundances of Mn may indicate that low-metallicity progenitors to supernovae (of either core collapse or SNe Ia) dominated the production of manganese within $\\omega$ Cen. This result for Mn is similar to what has been noted previously for the behavior of copper (which in some nucleosynthesis processes has metallicity-dependent yields) in $\\omega$ Cen (Cunha et al. 2002). The behavior of Mn in the more extreme metal-rich $\\omega$ Cen population (RGB-a) remains to be probed. In future studies it would be of interest to determine Mn abundances in the most metal-rich $\\omega$ Cen population; for example, by analyzing Mn in the 3 most metal-rich red giants studied to date (ROA 300, WFI22068, and WFI222679) by Pancino et al. (2002). The behavior of the manganese abundances in these more metal-rich stars will provide further insight into both the origins of Mn and the nature of star formation within $\\omega$ Cen during the final throes of its chemical evolution. MB thanks Dr. Frank Grupp for providing MAFAGS-OS model atmospheres for selected stars. This research was supported in part by the National Science Foundation (AST 06-46790 to KC and VVS). DLL thanks the Robert A. Welch Foundation for support via grant F-634." }, "1005/1005.2964_arXiv.txt": { "abstract": "The survival of unbound density substructure against orbital mixing imposes strong constraints on the slope of the underlying gravitational potential and provides a new test on modified gravities. Here we investigate whether the interpretation that the stellar clump in Ursa Minor (UMi) dwarf spheroidal galaxy is a `dynamical fossil' is consistent with Modified Newtonian dynamics (MOND). For UMi mass models inferred by fitting the velocity dispersion profile, the stellar clump around the second peak of UMi is erased very rapidly, within $1.25$ Gyr ($6.5$ orbits), even with the inclusion of self-gravity. We find that the clump can hardly survive for more than $2$ Gyr even under more generous conditions. Alternative scenarios which could lead to a kinematically cold clump are discussed but, so far, none of them were found to be fully satisfactory. Our conclusion is that the cold clump in UMi poses a challenge for both $\\Lambda$CDM and MOND. ", "introduction": "The standard concordance cosmological model with cold dark matter ($\\Lambda$CDM model) is remarkably successful on scales larger than $1$ Mpc, but it faces challenges on smaller scales. For instance, it seems that the theory predicts a too cuspy density profile for the dark matter at the centres of galaxies (e.g., Trachternach et al.~2008). The MOdified Newtonian Dynamics (MOND) proposed by Milgrom (1983) has proven to be successful in reproducing the kinematics of spiral galaxies without any assumption of unseen matter (see Sanders \\& McGaugh 2002, for a review), from extremely low mass galaxies of low surface brightness (Milgrom \\& Sanders 2006) to high luminosity galaxies (Sanders \\& Noordermeer 2007). Gentile et al.~(2007) found that the observed rotation curves in tidal dwarf galaxies are quite naturally explained without any free parameters within MOND, and are inconsistent with the current $\\Lambda$CDM theory (see also Kroupa et al.~2005). If MOND is able to naturally account for all of the discrepancies faced by $\\Lambda$CDM on small scales, this would lend strong support to MOND. In the Newtonian dark matter scenario, dwarf spheroidal galaxies (dSph's) require the largest mass-to-light ratios. Hence, dSph's provide a unique testing ground for the nature of dark matter and its alternatives (Gerhard \\& Spergel 1992; Milgrom 1995; Lokas et al.~2006; S\\'anchez-Salcedo et al.~2006; Angus 2008). However, the mass-to-light ratios inferred in MOND are very sensitive to uncertainties on the structural parameters, luminosities, distances and internal velocity dispersions. It is, therefore, important to explore other gravitational effects, which, in principle, may offer independent tests. In the dark matter paradigm, the velocity dispersion profiles of the brightest dSphs are consistent with both a cuspy NFW halo and a cored dark halo. However, there exists some indirect evidence that dSph galaxies may possess a core (Kleyna et al.~2003, hereafter K03; Goerdt et al.~2006; S\\'anchez-Salcedo et al.~2006). In particular, K03 considered the survival of the cold density substructure detected in photometric data in Ursa Minor (UMi). This substructure appears as an off-centre localized stellar clump with low velocity dispersion $\\simeq 0.5$ km s$^{-1}$. K03 concluded that the secondary peak in UMi is a long-lived structure, surviving in phase-space because the underlying gravitational potential is close to harmonic. In the standard dark matter paradigm, this implies that the dark matter halo in UMi must have a large core because, if the dark halo has a central density cusp, the clump should have diluted in $\\sim 1$ Gyr. Even if dSph galaxies are influenced to some degree by the tidal forces exerted by the Milky Way, it is unlikely that such a large core may have a tidal origin (e.g., Stoehr et al.~2002; Hayashi et al.~2003; Pe\\~narrubia et al.~2008). It is worthwhile exploring if the competing MOND scenario can explain the survival of cold substructures in dSph galaxies in a natural way. The question that arises is whether the MOND gravitational potential can mimic the potential of a dark halo with a core in order to explain the very longevity of the dynamical fossil in UMi. In other words, can the inference of cored haloes in dSphs be accomodated naturally within MOND? This paper is organized as follows. In section \\ref{sec:tidalradius}, we give a general statement of the problem and outline some key analytic results for tidal dissolution of satellite systems. In section \\ref{sec:UMidm}, we briefly describe the observational properties of UMi dwarf and its clump. Some important issues regarding the assumptions and approximations made to study the evolution of the clump in MOND framework are given in section \\ref{sec:approximations}. Section \\ref{sec:survival} presents results on the evolution of a stellar clump in MOND. In section \\ref{sec:alternatives}, we discuss alternative scenarios to account for the persistence of kinematically cold substructure in dSph galaxies. Finally, our conclusions are summarized in section \\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} Observations of the rotation curves of spiral galaxies strongly support a one-to-one relation between gravity at any radius and the enclosed baryonic mass. Due to this empirical relation, modified gravity theories like MOND are able to account rather successfully for the amplitude and the shape of the rotation curves. There are some indirect phenomena that suggest that this empirical relation may break down at scales of dSph galaxies (e.g., Gilmore et al.~2007, and references therein), challenging the interpretation of a modification of gravity as a substitute for dark matter. Localized regions with enhanced stellar density and, where data permit, extremely cold kinematics have been detected in some dSph galaxies (e.g., Olszewski \\& Aaronson 1985; K03; Coleman et al.~2004; Walker et al.~2006). In the framework of CDM, K03 show that adopting a cored halo profile can preserve the UMi's clump incorrupted for a Hubble time. This cored halo can fit the observed stellar velocity dispersion and the persistence of the clump. However, although we are still unable to make robust predictions about how the dark matter distribution changes in the process of galaxy formation when the physics of baryons are included, the formation of a large core in a dark matter dominated galaxy, provides a hard challenge for $\\Lambda$CDM. Mashchenko et al.~(2008) showed that energy feedback in dwarf galaxies drives bulk gas motions and gravitational potential fluctuations large enough to turn the cusp into a flat core (see also Governato et al.~2009). They find that for a Fornax sized galaxy, the dark matter core has an average density of $0.2 M_{\\odot}$ pc$^{-3}$ at redshift $z=5.2$, and the density decays a factor $\\sqrt{2}$ in $\\sim 300$ pc. Still, such a core would not be enough to avoid orbital spreading of the clump. In the particular case of UMi dSph galaxy, the density excess around the secondary peak cannot be the remnant of a merger with UMi of a smaller, gas-rich system because the stars have the same properties in terms of color and magnitude as the body of the UMi population (Kleyna et al.~1998). One might wonder whether the density peak could instead be a projection effect and that what we are seeing is a cold, low-density tidal tail. Numerical experiments by Read et al.~(2006) have shown that this scenario is very unlikely. Another possibility is that the clump is a portion of the stationary debris of a disrupted globular cluster. This model is not implausible but cannot be proven yet. A more drastic alternative scenario consists in interpreting the survival of substructures as an internal inconsistency of $\\Lambda$CDM. We have explored if MOND can explain the longevity of unbound clumps in dSph galaxies. Following K03, we have assumed that this overdensity structure is a disrupted stellar cluster and simulate its evolution in the gravitational potential derived in MOND. Whichever the form of the orbit, the clump is tidally disrupted within $1.5$ Gyr, even if clump's self-gravity is included. One can decrease UMi mass by adopting a smaller $\\Upsilon_{\\star}$, slowing the process of orbital mixing because the crossing time for the clump becomes larger. However, even assuming a mass-to-light ratio of $0.8 M_{\\odot}/L_{\\odot}^{V}$, the clump is disrupted in $<2.5$ Gyr. Our conclusion that tidal forces should have disrupted the clump appears robust for the adopted value of the stellar mass-to-light ratio of UMi. The external field acceleration felt by UMi from the Milky Way depends on the galactocentric distance of UMi, which is time dependent if the orbit of UMi around the Galactic halo is elongated. However, the temporal variation of the EFE can hardly boost the longevity of the clump. In the absence of any alternative model to explain the origin of the cold clump, we conclude that it is challenging for both $\\Lambda$CDM and MOND to explain the nature and dynamics of the clump in UMi. Another alternative of gravity suggested by Moffat (2005) is the so-called Modified Gravity (MOG). It is a fully covariant theory which predicts a Yukawa-like modification of Newton's law. For a system with a baryonic mass of a few times $10^{6} M_{\\odot}$ like UMi dSph galaxy, MOG predicts little or no observable deviation from Newtonian gravity at galactic distances of $\\sim 200$ pc (e.g., Moffat \\& Toth 2008). Therefore, explaining the internal dynamics of dSph galaxies is problematic without advocating dark matter, weakening the appeal of the MOG model. Other authors achieve to find modified theories of gravity which seem to reproduce the rotation curves of galaxies (e.g., Capozziello et al.~2007). Nevertheless, an analysis of the dynamics of dSph galaxies in these theories is still missing." }, "1005/1005.5069_arXiv.txt": { "abstract": "The CMB maps obtained by observations always possess domains which have to be masked due to severe uncertainties with respect to the genuine CMB signal. Cosmological analyses ideally use full CMB maps in order to get e.\\,g.\\ the angular power spectrum. There are attempts to reconstruct the masked regions at least at low resolutions, i.\\,e.\\ at large angular scales, before a further analysis follows. In this paper, the quality of the reconstruction is investigated for the ILC (7yr) map as well as for 1000 CMB simulations of the $\\Lambda$CDM concordance model. The latter allows an error estimation for the reconstruction algorithm which reveals some drawbacks. The analysis points to errors of the order of a significant fraction of the mean temperature fluctuation of the CMB. The temperature 2-point correlation function $C(\\vartheta)$ is evaluated for different reconstructed sky maps which leads to the conclusion that it is safest to compute it on the cut-sky. ", "introduction": "The cosmic microwave background (CMB) provides one of the cornerstones of the cosmological concordance model. The statistical properties of our cosmological models have to match those of the CMB in order to give an admissible model. Thus, it is of utmost importance to reliably extract the statistical properties of the CMB. A main obstacle is the foreground emission of our galaxy and of other sources which restrict the area of the sky available for a sufficiently clean CMB signal, i.\\,e.\\ the full sky CMB signal has to be masked. One path of statistical analysis leads to the Fourier space in which the CMB is decomposed with respect to spherical harmonics $Y_{lm}(\\hat n)$ where the masked sky leads to a coupling between the Fourier modes since no full sky CMB is available. In this paper, we do not delve into these difficulties, where upon an extensive literature exists, but instead follow the alternative path which allows an analysis directly in the pixel space, which in turn is more adapted to a masked sky. This analysis is based on the temperature two-point correlation function $C(\\vartheta)$, which is defined as \\begin{equation} \\label{Eq:C_theta} C(\\vartheta) \\; := \\; \\left< \\delta T(\\hat n) \\delta T(\\hat n')\\right> \\hspace{10pt} \\hbox{with} \\hspace{10pt} \\hat n \\cdot \\hat n' = \\cos\\vartheta \\hspace{10pt} , \\end{equation} where $\\delta T(\\hat n)$ is the temperature fluctuation in the direction of the unit vector $\\hat n$. The most direct way to deal with a mask is just to use only those pixels which are outside the mask. In this way it was discovered by the COBE team \\citep{Hinshaw_et_al_1996} that the correlation function $C(\\vartheta)$ possesses surprisingly low power at large angles $\\vartheta \\gtrsim 60^\\circ$. A surprising observation is made by using the ILC map, which represents a full sky CMB map obtained by the WMAP team \\citep{Gold_et_al_2010}. Computing the correlation function $C(\\vartheta)$ using a mask leads to a correlation function having very low power at large scales, whereas using the complete ILC map leads to a correlation function which possesses higher large scale power being compatible with the concordance model \\citep{Spergel_et_al_2003}. One has to decide which correlation function $C(\\vartheta)$ corresponds to the true CMB sky: the one which is based on the safe pixels, i.\\,e.\\ those outside the mask, or the other one, which would imply that most of the large scale power is generated by those areas hidden by the galaxy, i.\\,e.\\ by those pixels which have experienced much larger corrections. This question has recently stimulated much discussions, e.\\,g.\\ \\citep{Copi_Huterer_Schwarz_Starkman_2006,Copi_Huterer_Schwarz_Starkman_2008,% Copi_Huterer_Schwarz_Starkman_2010,Hajian_2007,% Aurich_Janzer_Lustig_Steiner_2007,Aurich_Lustig_Steiner_2009,% Sarkar_Huterer_Copi_Starkman_Schwarz_2010,% Bennett_et_al_2010,% Efstathiou_Ma_Hanson_2009,Pontzen_Peiris_2010}. \\cite{Efstathiou_Ma_Hanson_2009} emphasise that one has to start with the cut sky, but before the correlation function $C(\\vartheta)$ is computed the low-order multipoles have to be reconstructed. In this way, a stable result is obtained as long as the mask is not too large. The obtained correlation function is then the one with large power at large scales. For the details of the method see \\cite{deOliveira-Costa_Tegmark_2006,Bielewicz_Gorski_Banday_2004}. Here we summarise only the most important ingredients. The data vector $\\vec x$ containing only the pixel values outside the mask is related to the spherical harmonic coefficients $a_{lm}$ represented as $\\vec a$ by \\begin{equation} \\label{Eq:x_as_a} \\vec x \\; = \\; Y \\, \\vec a \\, + \\, \\vec n \\hspace{10pt} , \\end{equation} where $Y_{ij}$ denotes the corresponding values of $Y_{l_jm_j}(\\hat n_i)$ and $\\vec n$ the noise. In the low-order multipole reconstruction, only the multipoles with $l \\leq l_{\\hbox{\\scriptsize max}} = 10\\dots20$ are taken into account. The methods of reconstruction differ by the choice of a square matrix $A$ which determines the reconstructed $\\vec a\\,^r$ by \\begin{equation} \\label{Eq:ar_by_A} \\vec a\\,^r \\; = \\; (Y^T A Y)^{-1} \\, Y^T A \\vec x \\hspace{10pt} . \\end{equation} Setting the matrix $A$ equal to the unit matrix leads to the method of ``direct inversion''. To take the correlations between the pixels into account, one can choose the covariance matrix $A^{-1}= <\\vec x \\cdot \\vec x\\,^T>$ for a reconstruction up to $l_{\\hbox{\\scriptsize max}}$, which leads to the method used by \\cite{deOliveira-Costa_Tegmark_2006} with \\begin{equation} \\label{Eq:covariance_matrix} A_{ij}^{-1} \\; = \\; \\sum_{l=l_{\\hbox{\\scriptsize max}}+1}^{l_{\\hbox{\\scriptsize cut}}} \\frac{2l+1}{4\\pi} \\, P_l(\\hat n_i \\cdot \\hat n_j) \\, C_l \\end{equation} ignoring the noise contribution. The sum in (\\ref{Eq:covariance_matrix}) runs only over those multipole moments $C_l$, $l>l_{\\hbox{\\scriptsize max}}$, that are not to be reconstructed. The reconstructed sky map up to $l_{\\hbox{\\scriptsize max}}$ is then obtained from the coefficients $\\vec a\\,^r$. \\cite{Efstathiou_Ma_Hanson_2009} compute the correlation function $C(\\vartheta)$ from such reconstructed maps for $l_{\\hbox{\\scriptsize max}}=5$ to $l_{\\hbox{\\scriptsize max}}=20$ by using the KQ85 or the KQ75 mask provided by the WMAP team, and it is found that these $C(\\vartheta)$ only show a negligible variation. This is interpreted as a sign for a stable method to compute the temperature correlation in the presence of masks. It should be emphasised that the WMAP team bases its 7 year investigation with respect to CMB anomalies \\citep{Bennett_et_al_2010} on the method of \\citep{Efstathiou_Ma_Hanson_2009}. This leads to the conclusion that the correlation function agrees well with the $\\Lambda$CDM concordance model and displays thus no anomalous behaviour. The system of equations (\\ref{Eq:ar_by_A}) is over-determined, since there are much more pixel values than multipole moments with $l \\leq l_{\\hbox{\\scriptsize max}}$ as long as the mask and $l_{\\hbox{\\scriptsize max}}$ are not too large. Thus, there is the hope that the pixels outside the mask already determine the low-order multipoles. The caveat in the demonstration of \\citep{Efstathiou_Ma_Hanson_2009} is, however, that a large Gaussian smoothing of $10^\\circ$ is applied to the maps {\\it before} the cut is applied. Due to this smoothing there is a transfer of information from pixel values within the mask to those outside the mask. Since the system of equations (\\ref{Eq:ar_by_A}) is already over-determined, this additional information is readily extracted and reveals the temperature structure within the mask after the ``reconstruction''. It should be noted that even a downgrade in the HEALPix resolution can lead to an information transfer by carrying out the downgrade before the mask is applied. \\begin{figure} \\begin{center} { \\begin{minipage}{11cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/wmap_temperature_analysis_mask_r9_7yr_v4_small.eps} \\end{minipage} \\put(-315,67){(a) KQ85 (7yr) mask, $N_{\\hbox{\\scriptsize side}} = 512$} } \\vspace*{5pt} { \\begin{minipage}{11cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/wmap_temperature_analysis_mask_r9_7yr_v4_s0.5_nside_16_small.eps} \\end{minipage} \\put(-315,67){(b) KQ85 (7yr) mask, $N_{\\hbox{\\scriptsize side}} = 16$, $x_{\\hbox{\\scriptsize th}}=0.5$ } } \\end{center} \\caption{\\label{Fig:KQ85_masks_nside_512_and_nside_16} The KQ85 mask of the WMAP 7 year data with a pixel resolution of $N_{\\hbox{\\scriptsize side}} = 512$ is displayed in panel \\ref{Fig:KQ85_masks_nside_512_and_nside_16}a. The masked region is pictured in black. Figure \\ref{Fig:KQ85_masks_nside_512_and_nside_16}b shows the KQ85 mask downgraded to a pixel resolution of $N_{\\hbox{\\scriptsize side}} =16$. All pixels with $x(i)\\le 0.5$, equation (\\ref{Eq:not_masked_pixel_downgr}), are considered as masked. } \\end{figure} The two masks used in this paper are the KQ85 (7yr), shown in figure \\ref{Fig:KQ85_masks_nside_512_and_nside_16}a, and the KQ75 (7yr) mask which are available at the LAMBDA website. These are stored in the HEALPix \\citep{Gorski_Hivon_Banday_Wandelt_Hansen_Reinecke_Bartelmann_2005} format with a HEALPix resolution of $N_{\\hbox{\\scriptsize side}} = 512$. The reconstruction algorithm requires masks in lower resolutions of $N_{\\hbox{\\scriptsize side}} = 16$ or 32, and a downgrade has to be carried out. A single downgraded pixel $i$ contains $N_{\\hbox{\\scriptsize total}}$ pixels of the higher resolution map and from these only $N_{\\hbox{\\scriptsize nm}}$ pixel values are used in the averaging process, i.\\,e.\\ those that are {\\it not} masked in the higher resolution map, in order to compute the value of the downgraded pixel. If the ratio \\begin{equation} \\label{Eq:not_masked_pixel_downgr} x(i) \\; = \\; \\frac{N_{\\hbox{\\scriptsize nm}}} {N_{\\hbox{\\scriptsize total}}} \\end{equation} is larger than a given mask threshold $x_{\\hbox{\\scriptsize th}}\\in(0.0, 1.0)$, i.\\,e.\\ $x(i)>x_{\\hbox{\\scriptsize th}}$, the resulting pixel is {\\it not} masked, and otherwise it is masked. The result of the KQ85 (7yr) mask for $N_{\\hbox{\\scriptsize side}} =16$ and $x_{\\hbox{\\scriptsize th}}=0.5$ is shown in figure \\ref{Fig:KQ85_masks_nside_512_and_nside_16}b. It is obvious that the size of the masked domains in masks with $N_{\\hbox{\\scriptsize side}} <512$ depends on this mask threshold $x_{\\hbox{\\scriptsize th}}$. \\begin{figure} \\begin{center} { \\begin{minipage}{11cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/lcdm_7yr_wmap_bao_h0_nside_512_lmax_1024_fwhm_60.0arcmin_plus_fwhm_600.0arcmin_seed_126_sw_small.eps} \\end{minipage} \\put(-325,75){(a) $\\Lambda$CDM simulation A} } \\vspace*{10pt} { \\begin{minipage}{11cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/lcdm_7yr_wmap_bao_h0_nside_512_lmax_1024_fwhm_60.0arcmin_plus_fwhm_600.0arcmin_seed_166_sw_small.eps} \\end{minipage} \\put(-325,75){(b) $\\Lambda$CDM simulation B} } \\vspace*{10pt} { \\begin{minipage}{11cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/lcdm_7yr_wmap_bao_h0_nside_512_lmax_1024_fwhm_60.0arcmin_seed_126_und_seed_166_maske_KQ85_7yr_600.0arcmin_sw_small.eps} \\end{minipage} \\put(-325,75){(c) $\\Lambda$CDM simulation A outside and B inside KQ85 mask} } \\end{center} \\caption{\\label{Fig:CutSky_Info_Input_maps} Figure \\ref{Fig:CutSky_Info_Input_maps}a shows the CMB simulation A of the $\\Lambda$CDM concordance model with a pixel resolution of $N_{\\hbox{\\scriptsize side}} = 512$ and an additional smoothing of $10^\\circ$. In figure \\ref{Fig:CutSky_Info_Input_maps}b a second map (CMB simulation B) using the same cosmological parameters is displayed. The map in figure \\ref{Fig:CutSky_Info_Input_maps}c results from the models used for figure \\ref{Fig:CutSky_Info_Input_maps}a and \\ref{Fig:CutSky_Info_Input_maps}b where the pixels outside the KQ85 mask are taken from simulation A and inside from simulation B at the resolution $N_{\\hbox{\\scriptsize side}} = 512$. Thereafter a Gaussian smoothing of $10^\\circ$ is applied. } \\end{figure} \\begin{figure} \\begin{center} { \\begin{minipage}{10cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/lcdm_7yr_wmap_bao_h0_nside_512_lmax_1024_fwhm_60.0arcmin_plus_fwhm_600.0arcmin_seed_126_kov_s0.5_lmax_10_sw_small.eps} \\end{minipage} \\put(-295,75){(a) $\\Lambda$CDM simulation A} } \\vspace*{10pt} { \\begin{minipage}{10cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/lcdm_7yr_wmap_bao_h0_nside_512_lmax_1024_fwhm_60.0arcmin_plus_fwhm_600.0arcmin_seed_166_kov_s0.5_lmax_10_sw_small.eps} \\end{minipage} \\put(-295,75){(b) $\\Lambda$CDM simulation B} } \\vspace*{10pt} { \\begin{minipage}{10cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/lcdm_7yr_wmap_bao_h0_nside_512_lmax_1024_fwhm_60.0arcmin_seed_126_und_seed_166_maske_KQ85_7yr_600.0arcmin_kov_s0.5_lmax_10_sw_small.eps} \\end{minipage} \\put(-295,75){(c) $\\Lambda$CDM simulation A outside and B inside KQ85 mask} } \\end{center} \\caption{\\label{Fig:CutSky_Info_Extraction_lmax_10} The three reconstructed maps are displayed. The reconstructions are carried out for $l_{\\hbox{\\scriptsize max}}=10$. These maps result from the maps in figure \\ref{Fig:CutSky_Info_Input_maps}a, b and c by downgrading these maps to a pixel resolution of $N_{\\hbox{\\scriptsize side}} =16$ and thereafter applying the reconstruction algorithm (\\ref{Eq:ar_by_A}) to the data outside the mask shown in figure \\ref{Fig:KQ85_masks_nside_512_and_nside_16}b. Note that panel (a) and (c) should show the same sky map if the reconstruction algorithm would not use information from the masked region. However, there are significant difference especially near the Galactic plane. } \\end{figure} In the next sections we will compare results of CMB simulations of the $\\Lambda$CDM concordance model with those of the ILC map. For this reason all maps of the $\\Lambda$CDM concordance model are produced in a FWHM resolution of $1^\\circ$ and $N_{\\hbox{\\scriptsize side}} =512$. But we also investigate these maps and the ILC map after an additional smoothing of e.\\,g. $10^\\circ$. In the following only the additional smoothing width is specified. Let us now return to the information transfer caused by the smoothing procedure and/or by carrying out the downgrade. The following simple numerical experiment presented in figures \\ref{Fig:CutSky_Info_Input_maps} and \\ref{Fig:CutSky_Info_Extraction_lmax_10} reveals the information transfer. Figure \\ref{Fig:CutSky_Info_Input_maps}a displays the CMB simulation A of the $\\Lambda$CDM concordance model after a smoothing of $10^\\circ$. This map is downgraded from a pixel resolution of $N_{\\hbox{\\scriptsize side}} =512$ to $N_{\\hbox{\\scriptsize side}} =16$. Then the KQ85 mask in the pixel resolution $N_{\\hbox{\\scriptsize side}} =16$, shown in figure \\ref{Fig:KQ85_masks_nside_512_and_nside_16}b, is applied. For $l_{\\hbox{\\scriptsize max}}=10$ the reconstruction algorithm (\\ref{Eq:ar_by_A}) is used to obtain the reconstructed map shown in figure \\ref{Fig:CutSky_Info_Extraction_lmax_10}a. Both the original and the reconstructed map agree within the mask reasonably well. The same procedure is repeated for a second simulation B where figure \\ref{Fig:CutSky_Info_Input_maps}b shows the simulation and figure \\ref{Fig:CutSky_Info_Extraction_lmax_10}b the reconstruction. In the next step the pixels of simulation A are replaced at a pixel resolution of $N_{\\hbox{\\scriptsize side}} =512$ within the KQ85 mask, shown in figure \\ref{Fig:KQ85_masks_nside_512_and_nside_16}a, by those of simulation B shown in figure \\ref{Fig:CutSky_Info_Input_maps}b. After this replacement the smoothing of $10^\\circ$ is applied. The last step transfers now the ``wrong'' information to the pixels outside the mask. This smoothed map is shown in figure \\ref{Fig:CutSky_Info_Input_maps}c. Downgrading this map to $N_{\\hbox{\\scriptsize side}} =16$ provides the data outside the mask which are used for the reconstruction. If the reconstruction would not use the information within the mask, the reconstructed map of figure \\ref{Fig:CutSky_Info_Extraction_lmax_10}a should reappear. However, as revealed in figure \\ref{Fig:CutSky_Info_Extraction_lmax_10}c, the reconstruction algorithm generates within the mask the main structures of simulation B, which is displayed in figure \\ref{Fig:CutSky_Info_Input_maps}b. This clearly demonstrates the information transfer, so that one has to be careful in testing the reconstruction algorithm. This leads to the question whether the reconstruction can be carried out using only unsmoothed maps where no information about pixels within the mask is encoded outside. However, then stability difficulties arise as shown below. \\begin{figure} \\begin{center} \\begin{minipage}{11cm} \\hspace*{-20pt}\\includegraphics[width=9.0cm]{psplots/wmap_ilc_7yr_v4_leak_600.0arcmin_small.eps} \\end{minipage} \\end{center} \\caption{\\label{Fig:wmap_ilc_7yr_v4_leak_600.0arcmin} The leak of the ILC temperatures from regions within the mask to those outside due to the smoothing process. } \\end{figure} Another way to demonstrate the flow of information from regions within the mask to those outside the mask, is the following. At first the pixels of the ILC map outside the KQ85 mask are set to zero. Then a subsequent $10^\\circ$ smoothing shows how much information about the ILC pixel values inside the mask leaks to those regions outside. This leak of information is shown in figure \\ref{Fig:wmap_ilc_7yr_v4_leak_600.0arcmin}. It is obvious that the main structures of the ILC map within the mask appear close to the boundary outside the mask. The paper is organised as follows. In section 2 the reconstruction errors of the CMB temperatures are evaluated outside and inside the masks using different smoothings and different resolutions. The reconstruction method using the covariance matrix is compared with the direct inversion method. In section 3 the influence of the reconstruction method onto the 2-point correlation function $C(\\vartheta)$ of the CMB is investigated where the focus is on large scales. An integrated measure of $C(\\vartheta)$ serves in section 4 as a further tool to demonstrate the drawbacks of the reconstruction methods. Finally, in section 5 we summarise our results. ", "conclusions": "In this paper we have investigated the reconstruction of pixel values within masks using inversion methods. We have compared the algorithm using the covariance matrix with the direct inversion method. The comparison shows that for the reconstruction of multipoles $l_{\\hbox{\\scriptsize max}}\\lesssim 8$ the algorithm using the covariance matrix gives slightly smaller errors than the direct inversion method. For this reason we focus on the reconstruction method which uses the covariance matrix. An important point of our analysis is that the additional smoothing of $10^{\\circ}$ of the ILC map in \\cite{Efstathiou_Ma_Hanson_2009} transfers information from pixels inside the mask to pixels outside. This should not happen since the pixel values inside the mask are considered as insecure and should be ignored. The smoothing, however, leads to a quantitatively better reconstruction of the ILC map, if the masked domain is considered as containing the true pixel values which have to be recovered by the reconstruction. But the contribution of information from masked pixels makes a reconstructed map obtained from data with additional smoothing unacceptable in a realistic application. The errors of the pixel values of the reconstructed map are compared with the mean temperature fluctuation $\\sigma_{\\hbox{\\scriptsize true}}(l_{\\hbox{\\scriptsize max}})$ defined in equation (\\ref{Eq:Ortsraum_Normierung}). In the case of the KQ75 (7yr) mask the reconstructed pixels inside the mask have errors larger than $\\sigma_{\\hbox{\\scriptsize true}}(l_{\\hbox{\\scriptsize max}})$ already for $l_{\\hbox{\\scriptsize max}} \\gtrsim 6$ such that the reconstruction is unusable as revealed by figure \\ref{Fig:Q_ilc_KQ75}. Even for smaller values of $l_{\\hbox{\\scriptsize max}}$ the errors are at least of the order of $\\frac 12\\sigma_{\\hbox{\\scriptsize true}}(l_{\\hbox{\\scriptsize max}})$. The culprit for this negative result is the size of the KQ75 (7yr) mask being too large for a reconstruction. In the case of the smaller KQ85 (7yr) mask the reconstruction without additional smoothing leads for $l_{\\hbox{\\scriptsize max}} \\gtrsim 13$ to errors larger than $\\sigma_{\\hbox{\\scriptsize true}}(l_{\\hbox{\\scriptsize max}})$ such that a rough reconstruction can be done for $l_{\\hbox{\\scriptsize max}}$ around 10. However, even with this restriction, errors larger than $\\frac 12\\sigma_{\\hbox{\\scriptsize true}}(l_{\\hbox{\\scriptsize max}})$ occur, see figure \\ref{Fig:Q_ilc_KQ85}. This might be too large for cosmological applications. It is also checked that refining the resolution from $N_{\\hbox{\\scriptsize side}} = 16$, which is used in the above computations, to $N_{\\hbox{\\scriptsize side}} = 32$ does not improve the reconstruction. Furthermore, the behaviour of the temperature 2-point correlation function $C(\\vartheta)$ is analysed with respect to the various reconstructed maps, i.\\,e.\\ with and without additional smoothing, for the ILC (7yr) map and for CMB simulations of the $\\Lambda$CDM concordance model. The correlation function $C(\\vartheta)$ is computed from the full ILC map, from the data outside the KQ85 mask, and from the reconstructed ILC maps which depend on $l_{\\hbox{\\scriptsize max}}$. In order to get a good approximation of the correlation function $C(\\vartheta)$ on large angular scales, the multipoles up to $l_{\\hbox{\\scriptsize max}} \\simeq 10$ are needed. The application of the large KQ75 mask is therefore excluded. The errors of the reconstructed 2-point correlation function are estimated by using 1000 CMB simulations of the $\\Lambda$CDM model. Since this error estimation is based on simulations which contain only the pure CMB signal, no further errors are taken into account e.\\,g.\\ resulting from residual foreground and detector noise. Therefore, the error estimation of the correlation function of the ILC map gives only a lower bound of the errors expected in a genuine application. As discussed above the reconstructed ILC map with additional smoothing of 600 arcmin uses information from inside the mask. This information transfer leads in turn to underestimated errors in the 2-point correlation function. Furthermore, the similarity of $C(\\vartheta)$ computed from the full ILC map and from the reconstructed map is due to this information transfer and not the achievement of the reconstruction algorithm. Without additional smoothing the ILC correlation function $C(\\vartheta)$, computed from the reconstructed map, does not match the one of the full map better than the one of the masked map within the estimated errors. This is caused by the large errors which increase with $l_{\\hbox{\\scriptsize max}}$. For $10 \\lesssim l_{\\hbox{\\scriptsize max}} \\lesssim 12$ the reconstructed 2-point correlation function is uncertain by 100$\\mu\\hbox{K}^2$ and for even larger values of $l_{\\hbox{\\scriptsize max}}$ useless as revealed by figure \\ref{Fig:correlation_function_ilc_KQ85_FWHM_000arcmin_nside_16}. This uncertainty is also reflected in the $S(60^{\\circ})$ statistic which integrates the power of the 2-point correlation function on angular scales larger than $60^{\\circ}$. The $S(60^{\\circ})$ statistic of the reconstructed ILC map does not converge to a stable value for $l_{\\hbox{\\scriptsize max}} \\gtrsim 6$ as it is the case for the full and for the masked ILC map as displayed in figure \\ref{Fig:s_statistik_1000lcdm_und_ilc_KQ85_FWHM_000arcmin_nside_16}a. In addition, the mean value of the $S(60^{\\circ})$ statistic, calculated from 1000 reconstructed CMB sky maps, reveals the instability of the algorithm for $l_{\\hbox{\\scriptsize max}} \\gtrsim 12$, where the power strongly increases, see the dashed curve in figure \\ref{Fig:s_statistik_1000lcdm_und_ilc_KQ85_FWHM_000arcmin_nside_16}b. This demonstrates the inability to reconstruct the temperature correlations. The conclusion of this paper is that by using a realistic mask for the WMAP data the reconstruction algorithm (\\ref{Eq:ar_by_A}) does not work well enough to obtain a prediction of the 2-point correlation function. For this reason a cosmological analysis of the WMAP data should use only data outside a mask, e\\,g.\\ the KQ85 (7yr) or KQ75 (7yr) mask. The Planck satellite measures the sky at more different frequencies than the WMAP satellite. This should allow a more secure reduction of the foreground in the CMB map. It is expected that the masked region in the Planck data is significantly smaller. It might be small enough in order to allow an acceptable reconstruction of the CMB within the mask which could provide a 2-point correlation function useable at large scales." }, "1005/1005.2225_arXiv.txt": { "abstract": "\\ammonia\\ and \\m\\ are key molecules in astrochemical networks leading to the formation of more complex N- and O-bearing molecules, such as \\mc\\ and \\dime. Despite a number of recent studies, little is known about their abundances in the solid state. This is particularly the case for low-mass protostars, for which only the launch of the \\spitzer\\ Space Telescope has permitted high sensitivity observations of the ices around these objects. In this work, we investigate the $\\sim 8-10$~\\micron\\ region in the \\spitzer\\ IRS (InfraRed Spectrograph) spectra of 41 low-mass young stellar objects (YSOs). These data are part of a survey of interstellar ices in a sample of low-mass YSOs studied in earlier papers in this series. We used both an empirical and a local continuum method to correct for the contribution from the 10~\\micron\\ silicate absorption in the recorded spectra. In addition, we conducted a systematic laboratory study of \\ammonia- and \\m-containing ices to help interpret the astronomical spectra. We clearly detect a feature at $\\sim$9~\\micron\\ in 24 low-mass YSOs. Within the uncertainty in continuum determination, we identify this feature with the \\ammonia\\ $\\nu_2$ umbrella mode, and derive abundances with respect to water between $\\sim$2 and 15\\%. Simultaneously, we also revisited the case of \\m\\ ice by studying the $\\nu_4$ C--O stretch mode of this molecule at $\\sim$9.7~\\micron\\ in 16 objects, yielding abundances consistent with those derived by \\citet{boogert-etal08} (hereafter paper I) based on a simultaneous 9.75 and 3.53~$\\mu$m data analysis. Our study indicates that \\ammonia\\ is present primarily in \\water-rich ices, but that in some cases, such ices are insufficient to explain the observed narrow FWHM. The laboratory data point to \\m\\ being in an almost pure methanol ice, or mixed mainly with CO or CO$_2$, consistent with its formation through hydrogenation on grains. Finally, we use our derived \\ammonia\\ abundances in combination with previously published abundances of other solid N-bearing species to find that up to 10--20~\\% of nitrogen is locked up in known ices. ", "introduction": "Ammonia and methanol are among the most ubiquitous and abundant (after H$_2$ and CO) molecules in space. Gaseous \\ammonia\\ and \\m\\ are found in a variety of environments such as infrared dark clouds, dense gas surrounding ultra-compact \\hii\\ regions, massive hot cores, hot corinos, and comets. Solid \\m\\ has been observed in the ices surrounding massive YSOs (e.g. \\citealt{schutte-etal91,dartois-etal99,gibb-etal04}) and more recently toward low-mass protostars \\citep{pontoppidan-etal03}. The presence of solid \\ammonia\\ has been claimed toward massive YSOs only \\citep{lacy-etal98,dartois-etal02,gibb-etal04,guertler02}, with the exception of a possible detection in the low-mass object IRAS 03445+3242 % \\citep{guertler02}. However, these detections are still controversial and ambiguous \\citep{taban-etal03}. Both molecules are key participants in gas-grain chemical networks resulting in the formation of more complex N- and O-bearing molecules, such as \\mc\\ and \\dime\\ (e.g. \\citealt{rodgers+charnley01}). Moreover, UV processing of \\ammonia- and \\m-containing ices has been proposed as a way to produce amino-acids and other complex organic molecules (e.g., \\citealt{munozcaro+schutte03,bernstein-etal02,oberg-etal09}). In addition, the amount of \\ammonia\\ in ices has a direct impact on the content of ions such as NH$_4^+$ and OCN$^-$, which form reactive intermediates in solid-state chemical networks. A better knowledge of the \\ammonia\\ and \\m\\ content in interstellar ices will thus help to constrain chemical models and to gain a better understanding of the formation of more complex, prebiotic, molecules. During the pre-stellar phase, \\ammonia\\ is known to freeze out on grains (if the core remains starless long enough -- \\citealt{lee-etal04}). Moreover, \\m\\ is known to have gas-phase abundances with respect to H$_2$ in hot cores/corinos that are much larger than in cold dense clouds: $\\sim (1-10)\\times 10^{-6}$ vs $\\leq 10^{-7}$, with the former values most likely representing evaporated ices in warm regions \\citep[e.g.][]{genzel-etal82,blake-etal87,federman-etal90}. Together, these findings suggest that ices are an important reservoir of \\ammonia\\ and \\m\\ and that prominent features should be seen in the absorption spectra toward high- and low-mass protostars. Unfortunately, as summarized in Table~\\ref{tab:features}, \\ammonia\\ and \\m\\ bands, with the exception of the 3.53~\\micron\\ \\m\\ feature, are often blended with deep water and/or silicate absorptions, complicating unambiguous identifications and column density measurements. This is particularly true for \\ammonia\\ whose abundance determination based on the presence of an ammonium hydrate feature at 3.47~\\micron\\ remains controversial (e.g. \\citealt{dartois+dhendecourt01}). Nonetheless, it is important to use all available constraints to accurately determine the abundances of these two molecules. Despite the overlap with the 10~\\micron\\ silicate (Si--O stretch) feature, the \\ammonia\\ $\\nu_2$ umbrella mode at $\\sim$9~\\micron\\ ($\\sim$1110~\\wn) offers a strong intrinsic absorption cross section and appears as the most promising feature to determine the abundance of this species in the solid phase. Moreover, the \\m\\ $\\nu_4$ C--O stretch at $\\sim$9.7~\\micron\\ ($\\sim$1030~\\wn) provides a good check on the validity of the different methods we will use to subtract the 10-\\micron\\ silicate absorption, since the abundance of this molecule has been accurately determined previously from both the 3.53 and 9.75~\\micron\\ features (see Paper I). \\begin{deluxetable}{lccp{6cm}} \\tablewidth{0pt} \\tablecaption{Selected near- and mid-infrared features of \\ammonia\\ and \\m. \\label{tab:features}} \\tablehead{Mode\t& $\\lambda$ (\\micron)\t& $\\bar{\\nu}$ (cm$^{-1}$) & Problem} \\startdata \\sidehead{\\ammonia\\ features:} $\\nu_3$ N--H stretch\t& \\phn 2.96\t& 3375\t& Blended with \\water\\ (O--H stretch, 3.05~\\micron/3275~\\wn) \\\\ $\\nu_4$ H--N--H bend\t& \\phn 6.16\t& 1624\t& Blended with \\water\\ (H--O--H bend, 5.99~\\micron/1670~\\wn), HCOOH\\\\ {\\bf \\boldmath{$\\nu_2$} umbrella}\t\t& {\\bf \\phn 9.00}\t& {\\bf 1110}\t& {\\bf Blended with silicate} \\\\ \\hline \\sidehead{\\m\\ features:} $\\nu_2$ C--H stretch\t& \\phn 3.53\t& 2827\t& --\\\\ $\\nu_6$ \\& $\\nu_3$ --CH$_3$ deformation\t& \\phn 6.85\t& 1460\t& Blended (e.g. with NH$_4^+$)\\\\ $\\nu_7$ --CH$_3$ rock\t& \\phn 8.87\t& 1128\t& Weak; blended with silicate\\\\ {\\bf \\boldmath{$\\nu_4$} C--O stretch}\t& {\\bf \\phn9.75}\t& {\\bf 1026}\t& {\\bf Blended with silicate} \\\\ Torsion\t\t& 14.39\t\t& \\phn695\t& Blended with \\water\\ libration mode\\\\ \\enddata \\tablecomments{The bold-faced lines indicate the features studied here.} \\tablecomments{The nomenclature for the NH$_3$ and CH$_3$OH vibrational modes are adopted from \\citet{herzberg45}. } \\end{deluxetable} More detailed spectroscopic information is particularly interesting for low-mass protostars as the ice composition reflects the conditions during the formation of Sun-like stars. Such detections have only become possible with \\spitzer, whose sensitivity is necessary to observe low luminosity objects even in the nearest star-forming clouds. The spectral resolution of the \\spitzer\\ Infrared Spectrograph (IRS, \\citealt{houck-etal04}) of $\\Delta \\lambda/\\lambda \\sim 100$ in this wavelength range is comparable to that of the Infrared Space Observatory (ISO) PHOT-S instrument but lower than that of the ISO-SWS and other instruments used to identify solid \\ammonia\\ toward high luminosity sources. The spectral appearance of ice absorption features, such as band shape, band position and integrated band strength, is rather sensitive to the molecular environment. Thus, the interpretation of the astronomical spectra should be supported by a systematic laboratory study of interstellar ice analogues containing \\ammonia\\ and \\m. Changes in the lattice geometry and physical conditions of an ice are directly reflected by variations in these spectral properties. In the laboratory, it is possible to record dependencies over a wide range of astrophysically relevant parameters, most obviously ice composition, mixing ratios, and temperature. Such laboratory data exist for pure and some H$_2$O-rich NH$_3$- and CH$_3$OH-containing ices \\citep[e.g.,][]{dhendecourt+allamandola86,hudgins-etal93,kerkhof-etal99,taban-etal03}, but a systematic study and comparison with observational spectra is lacking. In principle, the molecular environment also provides information on the formation pathway of the molecule. For example, NH$_3$ ice is expected to form simultaneously with H$_2$O and CH$_4$ ice in the early, low-density molecular cloud phase from hydrogenation of N atoms \\citep[e.g.,][]{tielens+hagen82}. In contrast, solid CH$_3$OH is thought to result primarily from hydrogenation of solid CO, a process which has been confirmed to be rapid at low temperatures in several laboratory experiments \\citep[e.g.][]{watanabe+kouchi02,hidaka-etal04,fuchs-etal09}. A separate, water-poor layer of CO ice is often found on top of the water-rich ice layer in low-mass star-forming regions due to the `catastrophic' freeze-out of gas-phase CO at high densities \\citep{pontoppidan-etal03,pontoppidan06}. Hydrogenation of this CO layer should lead to a nearly pure CH$_3$OH ice layer \\citep[e.g.,][]{cuppen-etal09}, which will have a different spectroscopic signature from that of CH$_3$OH embedded in a water-rich matrix. The latter signature would be expected if CH$_3$OH ice were formed by hydrogenation of CO in a water-rich environment or by photoprocessing of H$_2$O:CO ice mixtures, another proposed route \\citep[e.g.,][]{moore+hudson98}. Here, we present \\spitzer\\ spectra between 5 and 35~\\micron\\ of ices surrounding 41 low-mass protostars, focusing on the $\\sim 8-10$~\\micron\\ region that contains the $\\nu_2$ umbrella and $\\nu_4$ C--O stretch modes of \\ammonia\\ and \\m, respectively. This work is the fourth paper in a series of ice studies \\citep{boogert-etal08,pontoppidan-etal08-co2,oberg-etal08-ch4} carried out in the context of the \\spitzer\\ Legacy Program ``From Molecular Cores to Planet-Forming Disks'' (``c2d''; \\citealt{evans-etal03}). In Section~\\ref{sec:astro}, we carry out the analysis of the \\spitzer\\ data in $8-10~\\micron$ range. In Section~\\ref{sec:lab}, we present the laboratory data specifically obtained to help interpret the data that are discussed in Section~\\ref{sec:comparison}. Finally, we conclude in Section~\\ref{sec:ccl} with a short discussion of the joint astronomy-laboratory work (including the overall continuum determination). ", "conclusions": "\\label{sec:ccl} We have analyzed in detail the 8-10~\\micron\\ range of the spectra of 41 low-mass YSOs obtained with \\spitzer\\ and presented in \\citet{boogert-etal08}. The sources are categorized into three types: straight, curved and rising 8~\\micron\\ silicate wings, and for each category template sources with little or no absorption from ices around 9-10~\\micron\\ have been determined. This has led to two ways of subtracting the contribution from the 10~\\micron\\ silicate absorption: first, by determining a local continuum, and second, by scaling the templates to the optical depth at 9.7~\\micron. The two methods give consistent band positions of the NH$_3$ features, but the resulting widths can be up to a factor of two larger using the template continuum method. Taking into account the uncertainty in continuum determination, NH$_3$ ice is most likely detected in 24 of the 41 sources with abundances of $\\sim$2 to 15~\\% w.r.t. \\water, with an average abundance of 5.5$\\pm$2.0~\\%. These abundances have estimated uncertainties up to a factor of two and are not inconsistent with other features in the 3 and 6~\\micron\\ ranges. CH$_3$OH is often detected as well, but the NH$_3$/CH$_3$OH abundance ratio changes strongly from source to source. Our inferred CH$_3$OH column densities are consistent with the values derived in paper I. Targeted laboratory experiments have been carried out to characterize the NH$_3$ and CH$_3$OH profiles (position, FWHM, integrated absorbance). Comparison with the observational data shows reasonable agreement (within $\\sim$1\\%) for the position of the NH$_3$ feature in H$_2$O-rich ices, but the observed widths are systematically smaller than the laboratory ones for nearly all sources. The silicate template continuum method gives widths that come closest to the laboratory values. This difference in width (i.e. widths derived from astronomical spectra smaller than those in the laboratory spectra) suggests that the NH$_3$ abundances determined here may be on the low side. The CH$_3$OH profile is most consistent with a significant fraction of the CH$_3$OH in a relatively pure or CO-rich phase, consistent with its formation by the hydrogenation of CO ice. In contrast, the most likely formation route of NH$_3$ ice remains hydrogenation of atomic N together with water ice formation in a relatively low density molecular phase. Finally, the nitrogen budget indicates that up to 10 to 20~\\% of nitrogen is locked up in known ices." }, "1005/1005.5319_arXiv.txt": { "abstract": "{Recently, wavelets and $R / S$ analysis have been used as statistical tools to characterize the optical flickering of cataclysmic variables.} {Here we present the first comprehensive study of the statistical properties of X-ray flickering of cataclysmic variables in order to link them with physical parameters.} {We analyzed a sample of 97 X-ray light curves of 75 objects of all classes observed with the \\textit{XMM}-Newton space telescope. By using the wavelets analysis, each light curve has been characterized by two parameters, $\\alpha$ and $\\Sigma$, that describe the energy distribution of flickering on different timescales and the strength at a given timescale, respectively. We also used the $R / S$ analysis to determine the Hurst exponent of each light curve and define their degree of stochastic memory in time.} {The X-ray flickering is typically composed of long time scale events $(1.5 \\la \\alpha \\la 3)$, with very similar strengths in all the subtypes of cataclysmic variables $(-3 \\la \\Sigma \\la -1.5)$. The X-ray data are distributed in a much smaller area of the $\\alpha -\\Sigma$ parameter space with respect to those obtained with optical light curves. The tendency of the optical flickering in magnetic systems to show higher $\\Sigma$ values than the non-magnetic systems is not encountered in the X-rays. The Hurst exponents estimated for all light curves of the sample are larger than those found in the visible, with a peak at 0.82. In particular, we do not obtain values lower than 0.5. The X-ray flickering presents a persistent memory in time, which seems to be stronger in objects containing magnetic white dwarf primaries.} {The similarity of the X-ray flickering in objects of different classes together with the predominance of a persistent stochastic behavior can be explained it terms of magnetically-driven accretion processes acting in a considerable fraction of the analyzed objects.} ", "introduction": "Cataclysmic variables (CVs) are binary systems in which a late-type secondary star transfers matter onto a white dwarf (WD) primary. The accretion configuration is different depending on the magnetic field strength of the WD. In non-magnetic systems the accretion disk extends down to the WD surface, while a truncated disk may form in moderately magnetic CVs or may not be present at all in strongly magnetized systems (see \\citealt{warner95} for a review). Owing to the gravitational potential of the compact primary, the accretion of matter produces a non-negligible flux of X-rays. The mechanism responsible for this emission in non-magnetic CVs during quiescence is the shock heating that acts in the boundary layer between the accretion disk and the WD surface \\citep{patray85b,patray85a}, while in magnetic systems X-rays are emitted by a standing shock above the magnetic poles of the WD \\citep{aizu73}. In the latter case the X-ray flux is higher, which makes the magnetic systems the brightest X-ray CVs. Cataclysmic variables are subdivided into three main classes: novae (classical and recurrent), dwarf novae (DNe) and nova-like (NL) systems. In addition, there is a class of objects closely related to novae, the so-called super-soft X-ray sources (SSSs), whose members are characterized by a prominent soft spectral component due to thermonuclear burning at the WD surface \\citep[see e.g.][]{orio95}. A parallel classification can be made considering the strength of the magnetic field $B$ of the primary. In this case we have the non-magnetic systems, the intermediate polars (IPs, $B \\sim 5-10$~MG) and the polars ($B \\ga 10$~MG). The X-ray light curves of CVs may exhibit a variety of modulations. Periodic coherent modulations due to occultations of the emitting regions can be produced by the orbital motion or by the rotation of the WD. For magnetic systems periodic modulations can also be ascribed to absorption effects produced in the magnetically-confined accretion flow [``columns'' in polars \\citep{cropper90}, ``curtains'' in IPs \\citep{rosetal88}]. Besides this persistent variability, CVs may also show quasi-periodic oscillations (QPOs) in both the soft and hard X-ray bands as well as completely stochastic brightness variations, which are usually identified with the term ``flickering'' \\citep[see][and references therein]{kuuetal06}. Flickering is generally constituted by a sequence of random flares with typical timescales ranging from a few seconds to a few minutes. This phenomenon has been observed in the X-ray light curves of CVs of all classes. For example, observations with the \\textit{HEAO}-1 satellite allowed the discovery of soft X-ray flickering in the polar prototype \\object{AM~Her} \\citep{szketal80}, as well as hard X-ray aperiodic variability in the DN \\object{SS~Cyg} during quiescence \\citep{coretal84}. Flickering at timescales of $\\sim 10$~s was detected in the NL \\object{TT~Ari} using both \\textit{Einstein} \\citep{jenetal83} and \\textit{ASCA} \\citep{baykiz96} data. Rapid aperiodic fluctuations were discovered in the light curves of a bunch of CVs also using \\textit{ROSAT} observations \\citep{holetal94,roselta95,bucetal98}. A common observational feature is that X-ray flickering shows a continuous power law frequency spectrum, as it does in the visible region. A correlation between the time scale of the X-ray flickering activity and that detected in simultaneous optical observations was found in a number of CVs of different classes, e.g. the IP system \\object{V795~Her} \\citep{roselta95}, the polar \\object{EF~Eri} \\citep{watetal87}, the \\object{VY~Scl} star TT~Ari \\citep{jenetal83}, the old nova \\object{V603~Aql} \\citep{dreetal83} and in the DNe SS~Cyg and \\object{U~Gem} \\citep{coretal84}. This observational evidence has been commonly explained as the result of reprocessing of the X-ray flickering energy. The properties of flickering have been studied for a long time, especially in the visible region of the electromagnetic spectrum. However, the origin of optical flickering is still uncertain, although there is plenty of evidence that it has to be related to the accretion process. It is very likely that the location of its source should be restricted either to regions very close to the WD, like the innermost part of the disk, or the hot spot (see \\citealt{bruch92} for a thorough discussion). The connection between X-ray flickering and the accretion onto the WD, instead, is evident. For highly magnetic CVs, the rapid soft X-ray flares often detected in their light curves could been explained as the result of a bombardment of the WD surface by random inhomogeneous structures (``blobs'') present in the accretion streams \\citep{kuipri82}. For instance, the features observed in the light curves of the polar \\object{V1309~Ori} \\citep{demetal98,schetal05} and of the asynchronous polar \\object{BY~Cam} \\citep{ramcro02}, although they are quite peculiar objects, could be explained with this hypothesis. Also the flares observed in the X-ray light curve of the old nova \\object{GK~Per}, which is an IP, have been interpreted as an indirect indication of blobby accretion \\citep{vrieetal05}. In general, it is often possible to obtain some information about the flickering from the observations of a single object. However, as this is a stochastic process in time, it can be better studied by means of its statistical properties. \\citet[][hereafter FB98]{fribru98} carried out a statistical analysis of the optical flickering of a large sample of CVs based on the wavelet analysis \\citep{daubechies92} of their light curves. They represented the properties of flickering in a two-dimensional parameter space and showed that CVs of different classes tend to occupy different regions of this space. Later \\citet[][hereafter TDB09]{tametal09} used the $R / S$ rescaled range analysis \\citep{hurst51,huretal65} for the first time as a complementary tool to characterize the degree of persistence/anti-persistence of the white light flickering of the IP-class CV \\object{V709~Cas}. Motivated by the interesting results presented in these works, we have therefore used the same statistical techniques to study the X-ray flickering properties of CVs as a whole as well as a function of their individual classes to link them with physical parameters. ", "conclusions": "We have used the wavelets and the $R / S$ analysis as statistical tools to characterize the flickering in a sample of 97 X-ray light curves of CVs of all classes. In general, the distribution of the X-ray data points in the $\\alpha -\\Sigma$ parameter space is almost independent of the class and, therefore, quite different from that in the visible. The $\\alpha$ parameters are all contained in the range $1.5 \\la \\alpha \\la 3$. This implies that the dissipation of the flickering energy occurs typically in long time scale flares. Moreover, the strength of the X-ray flickering is found to vary in the region $-3 \\la \\Sigma \\la -1.5$, which is much smaller than that found in the optical region. There is no evidence for magnetic NL systems to have higher $\\Sigma$ values than the non-magnetic systems. The different behavior in the visible region and in the X-rays could be explained assuming that part of the optical flickering in system harboring an accretion disk could be originated by other slower components such as trailing waves. All objects in our sample show values of the Hurst exponent higher than 0.5, with a peak of the distribution at $H = 0.82$. The predominance of a persistent stochastic behavior can be ascribed to a magnetically-driven accretion process acting in the majority of the CVs considered here. In particular, the result obtained for magnetic CVs clearly depicts the behavior of a plasma trapped by the magnetic field in the accretion streams. However, it must be pointed out that the sample of the X-ray light curves analyzed here is still not large enough to allow us to draw firm conclusions. Therefore, the analysis of further data collected in future X-ray observations will permit us to overcome possible selection effects and to better define the statistical properties of the flickering. On the other hand, a larger sample of polar systems analyzed in the optical region would allow us to verify our suggestion that the observed properties of X-ray flickering are intimately related to magnetically-driven accretion processes. As a final remark, it would have been interesting to use the method proposed by TBD09 to calculate the Hurst exponent from the $\\alpha$ parameter (or vice-versa) and quantitatively demonstrate the impact of that procedure. However, it was not possible to do it with the X-ray data because for the vast majority the of the objects in the sample the $R / S$ analysis has been typically stopped at a timescale comparable to the second or third ``octave'' of the corresponding scalegrams. For this reason, the $H$ values obtained with the two methods are not comparable because either they refer to different timescales maybe including multiple Hurst exponents, or the $\\alpha$ parameter is obtained by linearly fitting just two or three points, which is not reliable." }, "1005/1005.3824_arXiv.txt": { "abstract": "{The Sunyaev--Zel'dovich (SZ) effect is a spectral distortion of the cosmic microwave background as observed through the hot plasma in galaxy clusters. This distortion is a decrement in the CMB intensity for $\\lambda > 1.3 \\,$mm, an increment at shorter wavelengths, and small again by $\\lambda \\sim 250 \\, \\mu$m. As part of the \\textit{Herschel} Lensing Survey (HLS) we have mapped \\bulletc\\ (the Bullet cluster) with SPIRE with bands centered at $250$, $350$ and $500 \\, \\mu$m and have detected the SZ effect at the two longest wavelengths. The measured SZ effect increment central intensities are $\\Delta I_{0} = 0.097 \\pm 0.019 \\,$MJy sr$^{-1}$ at $350 \\, \\mu$m and $\\Delta I_{0} = 0.268 \\pm 0.031 \\,$MJy sr$^{-1}$ at $500 \\, \\mu$m, consistent with the SZ effect spectrum derived from previous measurements at $2 \\,$mm. No other diffuse emission is detected. The presence of the finite temperature SZ effect correction is preferred by the SPIRE data at a significance of $2.1 \\sigma$, opening the possibility that the relativistic SZ effect correction can be constrained by SPIRE in a sample of clusters. The results presented here have important ramifications for both sub-mm measurements of galaxy clusters and blank field surveys with SPIRE.} ", "introduction": "The Sunyaev-Zel'dovich (SZ) effect is a distortion of the spectral shape of the cosmic microwave background (CMB) due to inverse Compton scattering in the ubiquitous, hot ($T_{\\mathrm{e}} \\sim 10^{7} \\,$K) intracluster medium (ICM) of galaxy clusters \\citep{Sunyaev1972}. The canonical thermal SZ spectrum is a decrement in the brightness of the CMB as measured through galaxy clusters at mm wavelengths and an increment at sub-mm wavelengths which passes though a null at $\\lambda \\approx 1.3 \\,$mm. To correctly describe the SZ spectral distortion when relativistic electrons are present or the cluster is moving with respect to the CMB additional correction terms, usually termed ``relativistic'' (or ``finite temperature'') corrections and ``kinetic'' SZ effect, are required. Measurement of these corrections is only possible using observations at multiple wavelengths, and is expedited by measurement at wavelengths where the expected modifications to the thermal SZ effect are largest. In the case of the finite temperature corrections, the largest changes expected in the SZ increment are at wavelengths shorter than $1 \\, \\mu$m. Decrements in emission are rare astrophysically and can be ascribed to the SZ effect with little ambiguity; this has lead to measurements of the SZ effect at $\\lambda \\gtrsim 2 \\,$mm becoming almost routine. In comparison, measurements of the SZ effect increment are complicated by the presence of the dusty, high redshift galaxies which constitute the sub-mm cosmic background. Additionally, the individual sources comprising the sub-mm background are gravitationally lensed by galaxy clusters, the effect of which is to preferentially correlate increases in sub-mm emission with clustering. This correlation makes unambiguous detection of the SZ increment difficult, though successful measurements do exist (\\citealt{Lamarre1998}, \\citealt{Komatsu1999}, \\citealt{Zemcov2003}, \\citealt{Zemcov2007}, \\citealt{Nord2009}). Moreover, the presence of the sub-mm background may contaminate measurements of the SZ effect for $\\lambda > 1 \\,$mm in less massive galaxy clusters \\citep{Aghanim2005}. A better understanding of the sub-mm emission associated with galaxy clusters is required. Heretofore, systematic far infrared (FIR) surveys of many galaxy clusters to large radii have been technically challenging so a complete census of sub-mm emission from clusters has been difficult to obtain. The advent of SPIRE \\citep{Griffin2010} on \\textit{Herschel} \\citep{Pilbratt2010} has, for the first time, provided the capability to make deep maps of clusters to large angles on the sky and to use colour information to separate the different sources of sub-mm emission present in galaxy clusters. In addition to gravitationally lensed background sources \\citep{Rex2010}, emission in clusters above the confused sub-mm background may also comprise emission from galaxies in the cluster itself \\citep{Rawle2010}, as well as truly diffuse emission from the SZ effect and possibly even cold dust in the ICM. SPIRE's ability to separate sources of emission based both on spatial and spectral information allows the demographics of the sub-mm emission to be measured. In this paper, we use deep SPIRE maps of the $z=0.3$ Bullet cluster (\\bulletc) taken as part of the \\textit{Herschel} Lensing Survey (HLS, P.I.~Egami) at $250$, $350$, and $500 \\, \\mu$m with $18$, $25$ and $36$ arcsec resolution to measure the SZ effect and constrain other diffuse emission associated with the cluster. ", "conclusions": "\\label{S:conclusion} The possibility of detecting the SZ effect significantly shortward of its positive peak is a testament to the extraordinary capabilities of SPIRE. However, works like \\citet{Zemcov2007} show that in large galaxy cluster survey samples, significant contamination to the SZ signal from bright, gravitationally lensed background sources is common. Though the properties of this particular cluster have not precluded measurement of the SZ effect -- \\bulletc\\ has fewer bright, lensed sources close to its SZ effect center than typical clusters which have been observed in the sub-mm and is relatively bright and broad in the SZ effect -- based on this single example it is difficult to determine whether gravitational lensing of the sub-mm background will make measurements similar to this one more challenging in more typical clusters at these and other wavelengths. Data from surveys like the HLS will allow us to understand whether typical clusters are suitable for SZ effect increment detection, and how the lensed sub-mm background will affect measurements at other wavelengths. The possibility of diffuse emission from cold dust in the ICM has been discussed in the past (\\textit{e.g.} \\citealt{Stickel2002}). It is expected that, due to sputtering by energetic photons, such dust would have a very short lifetime in the ICM environment \\citep{Draine1979}. Based on the radial averages of the $250 \\, \\mu$m source subtracted maps, where for ICM dust with reasonable temperatures the brightest thermal emission would occur, we find no evidence for this type of diffuse emission in this cluster. Because in any reasonable scenario such dust emission would be faint in a $z > 0.1$ cluster, we expect that targeted searches of local clusters will be more successful for this science, though the HLS and similar surveys can provide useful constraints. Given the statistical uncertainties, the $350 \\, \\mu$m SZ effect increment measured here is slightly less than $2 \\sigma$ larger than would be expected from the $2 \\,$mm measurements. This may point to residual problems with the source subtraction procedure. More data will allow tuning of the confused sub-mm background removal algorithm and checks on whether biases arising from poor background removal are endemic in a large cluster sample. Given the presence of the SZ effect in clusters, it seems that care must be taken with photometry of $500 \\, \\mu$m sources within an arcminute or so of the cluster center; such sources will be positioned on a diffuse background so their fluxes will be biased by a small amount \\citep{Rex2010}. As the peak of the FIR emission of dusty sources is redshifted to progressively longer wavelengths, high redshift sources are expected to have exceptionally red spectral energy distributions. The results of this work show that care must be taken when searching for sources based on their presence in the $500 \\, \\mu$m band alone; galaxy clusters whose SZ effects are relatively bright and compact could well masquerade as such sources. Due to the well known problem of determining counterparts to sub-mm sources at other wavelengths, such SZ effect contamination may not be immediately obvious, particularly as cluster fields are crowded and many possible counterparts may be present. As a corollary, searching for very red sources in confusion limited SPIRE surveys may turn up compact clusters based on the presence of strong $500 \\, \\mu$m emission. The search for such extremely red SPIRE sources in the HLS and other programmes is underway now." }, "1005/1005.3223_arXiv.txt": { "abstract": "{In a previous paper we showed that the radio sources selected by combining large areas radio and optical surveys, present a strong deficit of radio emission with respect to 3CR radio-galaxies matched in line emission luminosity. We argued that the prevalence of sources with luminous extended radio structures in high flux limited samples is due to a selection bias. Sources with low radio power form the bulk of the radio-loud AGN population but are still virtually unexplored. We here analyze their photometric and spectroscopic properties. From the point of view of their emission lines, the majority of the sample are Low Excitation Galaxies (LEG), similar to the 3CR objects at the same level of line luminosity. The hosts of LEG are red, massive ($10.5 \\lesssim {\\rm log} \\, M_*/M_{\\odot} \\lesssim 12$) Early-Type Galaxies (ETG) with large black holes masses ($7.7 \\lesssim {\\rm log} \\, M_{\\rm{BH}}/M_{\\odot} \\lesssim 9$), statistically indistinguishable from the hosts of low redshift 3CR/LEG sources. No genuine radio-loud LEG could be found associated with black holes with a mass substantially lower than $10^8 M_{\\odot}$ or with a late type host. The fraction of galaxies with signs of star formation ($\\sim 5\\%$) is similar to what is found in both the quiescent ETG and 3CR/LEG hosts. We conclude that the deficit in radio emission cannot be ascribed to differences in the properties of their hosts. We argue that instead this could be due to a temporal evolution of the radio luminosity. A minority ($\\sim 10\\%$) of the sample show rather different properties, being associated with low black hole masses, with spiral galaxies, or showing a high excitation spectrum. In general these outliers are the result of the contamination from Seyfert and from galaxies where the radio emission is powered by star formation. For the objects with high excitation spectra there is no a clear discontinuity in either the host or nuclear properties as they span from radio-quiet and radio-loud AGN. ", "introduction": "The advent of large area surveys opened the possibility for the scientific community to explore large samples of extragalactic sources and to set the results on several key issues on strong statistical foundations. In particular, the cross-match of astronomical data from radio and optical surveys provides a unique tool in the analysis of the radio emitting galaxies, since it allows us to identify optically large numbers of radio sources, to obtain spectroscopic redshifts, to determine the properties of their hosts, and to build up their spectral energy distributions. In recent years several studies have been indeed carried out on large samples of radio galaxies in order to investigate the links between the radio structures, the central engine associated with an Active Galactic Nucleus (AGN), and the host galaxies. In particular, \\citet{best05b} selected a sample of 2215 low-luminosity radio galaxies by cross-correlating SDSS (Data Release 2), NVSS, and FIRST\\footnote{Sloan Digital Sky Survey, \\citep{york00}, National Radio Astronomy Observatory (NRAO) Very Large Array (VLA) Sky Survey \\citep{condon98}, and the Faint Images of the Radio Sky at Twenty centimeters survey \\citep{becker95} respectively.}. The resulting catalogue is by far larger than any previously studied sample of fully optically identified radio-sources. This sample (hereafter we refer to it as SDSS/NVSS sample) is highly (95 \\%) complete down to the flux threshold of 5 mJy and provides a very good representation of radio-galaxies in the local Universe, up to a redshift of $\\lesssim 0.3$, covering the range $10^{38} - 10^{42}$ erg s$^{-1}$ in radio power. In a subsequent paper \\citep{best05a} studied the properties of the host galaxies of these radio-loud AGN (hereafter RLAGN) finding that these are massive galaxies, usually of early Hubble type, with stellar masses in the range log $(M_*/M_{\\odot}) \\sim 10^{10}-10^{12}$, and located in richer environments than normal galaxies. The SDSS/NVSS sample was then considered by \\citet{baldi09} in the context of the properties of low luminosity radio-galaxies. We show that they display a strong deficit of radio emission with respect to their nuclear emission-line luminosity when compared to sources, part of other samples of radio-galaxies, matched in line luminosity. This result is particularly intriguing considering that the presence of a strong correlation between line and radio luminosity is instead well established (see, e.g., \\citealt{baum89b,rawlings89,morganti97,tadhunter98,willott99}) from the study of different samples of radio-galaxies. Both Low and High Excitation Galaxies (LEG and HEG respectively) obey separately to such a correlation, although the normalizations for the two classes differ \\citep{buttiglione10}. This suggests that the energy source of the narrow lines is closely linked to the source of the radio emission. It can be envisaged that a constant fraction of the available accretion power is channeled into radiative emission (that powers the emission line regions) and into jet kinetic energy (of which the radio emission is the electromagnetic manifestation). The study of the SDSS/NVSS sample suggests instead that total radio luminosity and line emission are independent from each other. Nonetheless, considering the properties of miniature radio-galaxies \\citep{balmaverde06b} we found that a link exists between radio {\\sl core} and line emission, extending to low luminosities the analogous relation present in high power sources. By assuming that the [O~III] luminosity gives an appropriate bolometric estimate of AGN power, and that radio cores are a good proxy for the jet power, radio galaxies with similar nuclear properties are able to produce an extremely wide range of total radio power, of a factor $\\geq$100. Indeed, these radio-galaxies of very low power are all highly core-dominated, with only feeble extended emission. A similar sample, studied by \\citet{prandoni09} at higher frequency (1.4-15 GHz) and selected at lower flux cutoff ($>1$ mJy), also shows a substantial fraction of very compact radio morphologies (d$<$10 kpc) in the same range of radio power covered by 3CR and B2 sample. We argued that the prevalence of sources with luminous extended radio structures in high flux limited samples is due to a selection bias, since the inclusion of such objects is highly favored. Core dominated sources with a low ratio between radio and emission line power form the bulk of the local radio-loud AGN population but they are still virtually unexplored. Unfortunately, the images available for the SDSS/NVSS sample are not sufficient to properly isolate their radio cores. The aim of this paper is a better understanding of the physical properties of these RLAGN. In particular we will explore the spectro-photometric properties of their hosts looking for possible differences that might explain their low level of radio emission with respect to the classical powerful radio galaxies matched in line luminosity. Operatively, with the information provided by SDSS we classify the optical spectra of the AGN on the basis of the excitation level. Then, we analyze the key properties of the hosts and the AGN, like the black hole mass, the mass and colors of the hosts, drawing a quantitative comparison with the 3CR sample. The main result is that the sample shows indistinguishable properties from those of 3CR sample. This implies that the reasons for the different radio properties between the bulk of the RLAGN population and the most studied sample of radio-galaxies must be ascribed to other mechanisms. The paper is organized as follows. In Sect.~\\ref{sample} we briefly present the sample selection carried out by \\citet{best05a}. In Sect.~\\ref{results} and ~\\ref{highz} we analyze the spectro-photometric properties of the SDSS/NVSS sample considering separately the sub-samples in the redshift ranges $0.0310~keV$. Such excess is ascribed to X-ray radiation piercing a partial, Compton thick ($\\rm N_H \\sim 4\\times 10^{24}~cm^{-2}$) absorber. The high energy excess appears stable over long time scales \\citep{risaliti00,risaliti09c}. The emerging scenario is that three distinct absorbers are present: (1) a distant absorber with $\\rm N_H < 10^{23}~cm^{-2}$, probably associated with gas in the host galaxy; (2) an absorber made of broad line region (BLR) clouds with $\\rm N_H\\sim 10^{23-24} cm^{-2}$ rapidly orbiting around the black hole; (3) an absorber with $\\rm N_H \\sim 4\\times 10^{24}~cm^{-2}$ partially covering the source, consisting either of the outer region of a warped accretion disk, or of a large number of small ($\\rm <10^{12}cm$) and dense ($\\rm n>10^{12}cm^{-3}$) clouds located in the vicinity of the accretion disk. In this paper we focus on the absorption variations in the spectral range 2-10~keV, ascribed to BLR clouds passing along the line of sight, through a detailed temporal analysis of the spectra provided by the {\\it Suzaku} low energy (XIS) data of NGC~1365. We identify two eclipsing events that we can model in detail in terms of variations of both covering factor and column density. The time variations of these quantities are highly asymmetric and strongly suggest that the absorbing clouds have a cometary shape, i.e. a well defined head and a more diffuse elongated, conical, tail. ", "conclusions": "We have presented a detailed spectral analysis of a long ($\\rm \\sim 300~ksec$), continuous {\\it Suzaku} observation of the Seyfert nucleus in the galaxy NGC~1365. The spectrum shows evidence for an absorption component that is variable in time, both in terms of column density and in terms of covering factor, which is ascribed to clouds eclipsing the X-ray source. This is the first time that a temporally resolved X-ray spectral analysis is able to break the degeneracy between the evolution of the column density ($\\rm N_H$) and the covering factor (CF) of the X-ray absorber. We identify two main eclipses. The temporal evolution of each eclipse is far from being symmetrical. The initial occultation is very rapid (within $\\rm \\sim 1~ksec$) and covering about 65\\% of the source. Subsequently the covering factor increases, but more slowly, reaching unity in about 50~ksec. The absorbing column density of the cloud is about $\\rm 10^{23}~cm^{-2}$ at the beginning of each eclipse and decreases afterwards. These results are inconsistent with a spherical geometry for the absorbing clouds. The most likely geometry compatible with the observations is an elongated, ``cometary'' shape, with a dense ``head'' ($\\rm n\\sim 10^{11}~cm^{-3}$, consistent with that expected for the clouds of the BLR) and with a dissolving and expanding tail. The data allow us to quantitatively constrain the geometry, dynamics and location of such cometary clouds. The cometary clouds are probably located at a distance of about $\\rm 2~10^{15}~cm$ from the nuclear black hole, well within the estimated BLR radius ($\\rm \\sim 10^{16}~cm$), strongly supporting the association of these absorbing systems with BLR clouds (at least with the inner, high-ionization ones). The cometary clouds ``head'' must have a size comparable to the X-ray source and must be moving with velocity higher than about 1000~km/s (consistent with the velocity expected for the BLR clouds). The cometary tail must be longer than a few times $\\rm 10^{13}~cm$. The tail opening angle must be very narrow, less than a few degrees, and consistent with the opening angle of the Mach cone expected from the supersonic motion of the cloud into the hot intracloud medium. We suggest that such cometary clouds may be common to most AGNs, but have been difficult to recognize in previous X-ray observations. We estimate that the cloud ``head'' loses a significant fraction of its mass through the cometary tail, which is expected to cause the total cloud destruction within a few months. If these clouds are representative of most BLR clouds (or at least the high-ionization ones), our result implies that the BLR region must be continuously replenished with gas clouds, possibly from the accretion disk. We briefly discussed the possible nature of such cometary clouds. The most likely scenario is that the tail is made of gas lost by the cloud head through hydrodynamical instabilities generated by its supersonic motion through the hot intracloud medium. We can estimate the mass of the absorbing clouds ($\\rm M_{cloud}\\sim 10^{-10}~M_{\\odot}$) and their total number within the central region ($\\rm \\mathcal{N_C}\\sim 3~10^7$). The inferred total mass of the BLR is about $\\rm 4~10^{-3}~M_{\\odot}$, which is two orders of magnitude lower than the BLR mass inferred from photoionization models. The discrepancy may originate from a population of large, massive BLR clouds not identified in our eclipsing studies. Alternatively, photoionization models may overestimate the BLR mass. In particular, UV and X-ray radiation produced by the shocks generated by the supersonic motion of the clouds may provide a local source of ionizing photons, not accounted for by classical photoionization models that assume a central point-like radiation source." }, "1005/1005.5363_arXiv.txt": { "abstract": "A recent observation of the nearby (z=0.042) narrow-line Seyfert 1 galaxy \\src\\ on 2007 May 31 showed strong quasi-periodic oscillations (QPOs) in the 0.3--10 keV X-ray flux. We present phase-resolved spectroscopy of this observation, using data obtained by the EPIC PN detector onboard \\xmm. The ``low'' phase spectrum, associated with the troughs in the light curve, shows (at $>4\\sigma$ confidence level) an absorption edge at $0.86\\pm0.05$ keV with an absorption depth of $ 0.3\\pm0.1$. Ionized oxygen edges are hallmarks of X-ray warm absorbers in Seyfert active galactic nuclei (AGN); the observed edge is consistent with H-like O VIII and implies a column density of $N_{\\rm OVIII}\\sim3\\times10^{18}$ cm$^{-2}$. The edge is not seen in the ``high'' phase spectrum associated with the crests in the light curve, suggesting the presence of a warm absorber in the immediate vicinity of the supermassive black hole which periodically obscures the continuum emission. If the QPO arises due to Keplerian orbital motion around the central black hole, the periodic appearance of the O VIII edge would imply a radius of $\\sim9.4(M/[4\\times10^6 \\msun])^{-2/3}(P/[1 hr])^{2/3}\\;r_{\\rm g}$ for the size of the warm absorber. ", "introduction": "\\src\\ is a narrow line Seyfert 1 galaxy (NLS1; see \\citealt{op1985} for definition of NLS1s). It is well known for its high soft X-ray excess compared not only to other AGNs but also among the NLS1s \\citep{pu1995,boller1996,middleton2007}. While the origin of this soft excess is still not clear, the similarity of its soft X-ray spectrum with that of stellar black hole binaries in their `high state' has been used to postulate that \\src\\ harbors a comparatively low mass black hole which is currently in a state of high mass accretion rate \\citep{pounds1995}. A recent X-ray observation of \\src\\ using the \\xmm\\ satellite made between 2007 May 31 and 2007 June 1 showed a strong signature of quasi-periodic variability in the 0.3--10 keV light curve \\citep{g08}. The oscillations were transient in nature since they have never been seen in any prior observations of this source, nor did we see strong QPO-like variability in the 0.3--10 keV light curve obtained during a subsequent \\xmm\\ observation made on 2009 May 31. It was noted by \\citet{g08} that the frequency of the observed oscillations was similar to what would be expected if the frequencies of the quasi-periodic oscillations often seen in Galactic black hole binaries \\citep[see e.g.][for a review on observations of QPOs in compact stellar X-ray binary systems]{vdk2006} were to be scaled by the same factor as the mass ratio between the mass of the supermassive black hole in \\src\\ and the mass of typical stellar black holes. Recently \\citet{MiddletonDone2009} have performed detailed comparison of the timing properties of this \\xmm\\ observation with that of stellar mass black hole binaries and concluded that the observed QPO in \\src\\ is similar to the 67 Hz QPO seen in GRS~1915+105 (a black hole binary system which, like \\src, also boasts a high L/L$_{\\rm Edd}$). This scaling of oscillation frequencies with mass would support the hypothesis that accretion physics scales with mass and that the supermassive black holes at the center of galaxies are scaled-up versions of stellar-mass black holes. The origin of QPOs in stellar black hole (and neutron star) binaries is not yet well understood. Various models have been proposed, but none can fully explain the wealth of observed phenomenology. QPO frequencies in stellar black hole binaries range from Hertz to kHz, and it is extremely difficult to carry out phase-resolved spectroscopy in these systems. In the sparse cases \\citep[e.g. in the source GRS~1915+105, reported by][]{Morganetal1997} where it is possible to ``see'' the QPO in the light curves, the phase of the oscillations appear to perform a random walk. \\citet{MillerHoman2005} showed that for GRS~1915+105 the flux in the Fe K$\\alpha$ line (created by hard X-ray photons irradiating a relatively cold accretion disk) varies with the QPO oscillation phase of the 1 and 2 Hz QPOs, thus suggesting that the 1 Hz and 2 Hz QPOs may be linked to variable reflection. The strong variability amplitude and comparatively large period of the oscillations in \\src\\ make it an ideal source to carry out phase-resolved spectroscopy and probe changes in the spectral energy distribution between different phases. The origin of the soft excess in \\src\\ is not clear, and its time-averaged spectral energy distribution can be well modeled by a wide variety of models including reflection from a partially ionized accretion disc \\citep{crummy2006}, Comptonized disc emission from a low temperature disc, ionized partial covering, or a smeared disc wind seen in absorption \\citep[see e.g.][]{middletonetal2009}. Phase-resolved spectra could give important insights for breaking this degeneracy. Therefore one of our major goals in this work is to search for signatures of variable reflection (in the continuum and reflection lines) and/or variable absorption features (e.g. changes in the properties of O VII, O VIII edges which are the hallmark of the presence of warm absorber in active galactic nuclei). \\citet{middletonetal2009} analyzed the energy dependence of the observed variability in the EPIC data and showed that the variability primarily originated in the high-energy photons. They also presented a spectral decomposition of the EPIC MOS spectrum, averaged over the entire observation. Here we analyze the EPIC PN data taken during this observation to test if there is any signature of spectral variations between the different phases. Due to limitations in photon statistics (largely caused by the high pileup on the CCD as described in \\S\\ref{prepare_data}; also see \\citealt{middletonetal2009}) we could extract spectra with good signal-to-noise ratio from two phases only: the ``high'' phase spectrum by accumulating spectra near the crests, and the ``low'' phase spectrum by accumulating spectra near the troughs. In \\S\\ref{prepare_data} we describe in detail the algorithm we used to create the phase-resolved ``high'' and ``low'' spectra. The spectral analysis is described in \\S\\ref{analysis}, and conclusions are summarized in \\S\\ref{conclusion}. ", "conclusions": "\\label{conclusion} In this work we have presented phase-resolved spectroscopy of the narrow line Seyfert 1 galaxy \\src. We have used the \\xmm\\ observation of this source between 2007 May 31 and 2007 June 1, when the X-ray light curve showed strong quasi-periodic oscillations. We have extracted spectra during the high (crest) and low (trough) phases. Simple continuum models assuming either a blackbody plus broken power-law, or two blackbodies plus a power-law model, fit the high phase spectrum quite well. The required low-energy power-law index is very steep ($\\Gamma\\sim3.6\\pm0.1$) for a single blackbody plus broken power-law model, which is typical for narrow line AGNs with a high soft X-ray excess \\citep{boller1996,pounds1995,middleton2007}. The spectrum can also be well modeled using two blackbodies and a power-law continuum. Due to the steep decline of the spectrum, and high pileup on the PN detector which severely limits the photon statistics, we do not have good quality spectral data beyond $\\sim$2--3 keV. Therefore for these spectra we cannot very well constrain the relatively flatter 2--10 keV continuum usually seen in this source \\citep{crummy2006,middletonetal2009}, although fits to a two-blackbody plus power-law model show that the $>2$ keV spectrum is comparatively flatter with a photon index of $\\sim2$. The low phase spectrum cannot be adequately described by scaling the overall normalization of either of the best fit high phase models, suggeting that the shape of the low phase spectrum is different from that of the high phase. Analyzing the r.m.s. variability at different energies \\citet{middletonetal2009} also concluded that the variability is mainly from the hard photons. The right panel of Fig.~\\ref{f:lo_renorm} shows that the differences between the low and high phases are most prominent at higher energies. In particular, the residuals in Fig.~\\ref{f:lo_renorm} show a sharp drop in the continuum flux between 0.8--0.9 keV, which then slowly increases toward the expected continuum flux at higher energies, reminiscent of an absorption edge. The inclusion of an edge improves the fits significantly (at $>$4$\\sigma$ level) for both single-blackbody+broken power-law model and double-blackbody+power-law model. The best fit values for the edge threshold energy and absorption depth are $0.87\\pm0.04$ keV and $0.29\\pm0.11$ respectively for the single-blackbody+broken power-law model, and $0.86^{+0.05}_{-0.06}$ keV and $0.23^{+0.12}_{-0.11}$ respectively for the double-blackbody+power-law model. The edge is most likely associated to the 0.87 keV (rest frame) K-edge of (H-like) oxygen VIII which is quite commonly seen in Seyferts \\citep[see e.g.,][]{reynolds1997}, and is indicative of the presence of optically thin, photoionized matter along the line of sight. \\src\\ however is at a redshift of 0.042; therefore the 0.87 keV edge (in rest frame) should appear at 0.84 keV (easily within the 90\\% confidence error bar for both models). The presence of a warm absorber should create other features in the spectrum, e.g. the 0.74 keV K edge from He-like O VII. While the residuals in Fig.~\\ref{fig:lo_edge} do show a drop in the flux, it is not very significant statistically. To estimate the column density ($N_{\\rm OVIII}$) of the O VIII ions responsible for the edge we used Eq.(1) of \\citet{verneretal1996} and calculated the absorption cross section $\\sigma$ (for O VIII, $\\sigma[{\\rm E}=0.87\\;{\\rm keV}]=0.098$ Mb). Since $\\tau=N_{\\rm OVIII}\\cdot\\sigma$, we obtain $N_{\\rm OVIII}=3\\times10^{18}$ cm$^{-2}$. Estimating the number density ($n_{\\rm OVIII}$) requires a size-scale for the warm absorber. Assuming that the QPO originates in the accretion disk, the warm absorber is could be confined within the orbit of the QPO, or in case of an outflow it originates inside the QPO orbit. The mass of the supermassive black hole in \\src\\ is in the range of $1-7\\times10^{6}$ \\msun\\ \\citep{BianHuang2010,zhouetal2010}. A circular Keplerian orbit would have a radius of $9.4(M/[4\\times10^6 \\msun])^{-2/3}(P/[1 hr])^{2/3}\\; r_{\\rm g}$, which would be an upper limit on the size of the absorber in this case. This would give $n_{\\rm OVIII}=3\\times10^5$ cm$^{-3}$ for the number density of OVIII ions. Note however that this assumes a Keplerian orbit origin for the QPO, and therefore depends on the mass of the central supermassive black hole. Thus in \\src\\ we may be seeing an infalling ``blob'', the emission from which is periodically absorbed as it passes behind a warm absorber in the immediate vicinity of the central engine. If the ``blob'' is at $\\sim$9 $r_{\\rm g}$, that would imply the ionizing flux must be generated within 9 $r_{\\rm g}$ (or else we would not see absorption), and this would be easier to accomplish if the black hole is spinning since the disk can get down to a smaller radius (e.g. as low as 1.25 $r_{\\rm g}$ for a maximally spinning black hole). Recent work by \\citet{crummy2006} has shown that NLS1 spectra can be modeled using reflection from a disk around a spinning black holes \\citep[see][for a review]{Miller2007}. We also created a multidimensional grid of models using the XSTAR photoionization code, to create a more physical model for the warm absorber and explore a wide range of column density, ionization parameter, oxygen and neon abundance. Fits to the low-phase data using this model suggests a warm absorber column density of $\\sim4\\times10^{21}$ cm$^{-2}$, log($\\xi$)=$3.2 _{-0.1}^{+0.2}$, and an oxygen abundance of $2.1 _{-0.5}^{+0.8}$ relative to solar. The neon abundance could not be well constrained from the data, although the best-fit solution suggests a sub-solar value. Another possible physical scenario for the difference between high and low phase spectrum could be that the warm absorber is along the line of sight, between us and the source (where the QPO originates), and the warm absorber responds to flux variations in the source itself, i.e. it is more ionized and transparent at high fluxes and less at lower fluxes. However, based on our modeling of the spectra using physically motivated models generated using XSTAR this scenario appears unlikely because the change in luminosity of the ionizing flux between low and high phases is too small to create the observed difference in the depth of the OVIII edge (also see \\S\\ref{analysis}). It is interesting to note in this context that a somewhat complementary situation has been observed in the Seyfert galaxy NGC~1365, where \\citet{Risalitietal2009} have reported a possible transit of an obscuring cloud (with $N_H\\sim3.5\\times10^{23}$ cm$^{-2}$ and other inferred properties similar to that of a broad-line region cloud) in front of the central X-ray source. Signatures of variable absorption on short timescales have been previously observed in high resolution {\\em Chandra} spectra of the stellar black hole candidate H1743--22 \\citep{miller2006}. For H1743--22, the lines were slightly blueshifted, suggesting an outflow origin. For \\src, we do not have definitive evidence for an outflow, but the best fit edge energy for both models is higher (though within 90\\% confidence interval) than the rest frame energy of the oxygen edge. If the warm absorber in \\src\\ is indeed in an outflow, then the observed oscillations could originate due to an instability in the inner accretion disk threaded by a poloidal magnetic field, giving rise to an outflowing wind or jet \\citep{BlandfordPayne1982,tagger1999}. The observed oscillations in this case would correspond to scaled up versions of low-frequency QPOs \\citep[see e.g.,][]{vdk2006} seen in stellar mass X-ray binaries. Since the presence of a warm absorber is bound to have noticeable imprints on other regions of the electromagnetic spectrum, especially in UV and soft X-rays, future deep UV and X-ray observations of \\src\\ will be useful not only in constraining the origin of the oscillations, but also in gaining a better understanding of the physical emission mechanism of sources like \\src\\ which show a strong soft-X-ray excess. Future X-ray missions with improved sensitivity and larger collecting area like the {\\em International X-ray Observatory} \\citep[IXO;][]{Milleretal2009,Whiteetal2010} would play key role in understanding accretion geometry as well as emission/absorption mechanism in these sources." }, "1005/1005.0293_arXiv.txt": { "abstract": "Stellar activity has a particularly strong influence on planets at small orbital distances, such as close-in exoplanets. For such planets, we present two extreme cases of stellar variability, namely stellar coronal mass ejections and stellar wind, which both result in the planetary environment being variable on a timescale of billions of years. For both cases, direct interaction of the streaming plasma with the planetary atmosphere would entail servere consequences. In certain cases, however, the planetary atmosphere can be effectively shielded by a strong planetary magnetic field. The efficiency of this shielding is determined by the planetary magnetic dipole moment, which is difficult to constrain by either models or observations. We present different factors which influence the strength of the planetary magnetic dipole moment. Implications are discussed, including nonthermal atmospheric loss, atmospheric biomarkers, and planetary habitability. ", "introduction": "One of the many fascinating questions in the field of exoplanet studies is the search for habitable worlds. Because of their relatively small mass, low luminosity, long lifetime and large abundance in the galaxy, M dwarfs are sometimes suggested as prime targets in searches for terrestrial habitable planets \\citep{Tarter07, Scalo07}. Interestingly, M dwarfs also seem to have a larger number of Super-Earth planets (i.e.~planets with a mass smaller than 10 terrestrial masses, i.e.~$M \\lesssim 10 M\\E$) than more massive stars \\citep{Forveille09}. Excluding planets orbiting around pulsars, the first super-Earth detected was GJ 876d, a planet with $\\sim\\!\\!7.5 M\\E$ or\\nolinebreak[4]biting an M star \\citep{Rivera05}. Thereafter, other super-Earth planets have been discovered \\citep{Beaulieu06,Lovis06,Udry07,Ribas08,Mayor09,Forveille09,Howard09,Bouchy09}. In total, 12 planets with masses $\\lesssim 10 M\\E$ are reported today, and more detections are expected for the near future \\citep{Howard09,Mayor09}\\footnote{Up-to-date numbers can be found at the Extrasolar Planets Encyclopaedia: {\\tt http://www.exoplanet.eu}}. The least massive planet has a mass of 4.2 $M\\E$. For M stars the habitable zone \\citep[HZ, defined by the range of orbital distances over which liquid water is possible on the planetary surface, see e.g.][]{Kasting93} is much closer to the star (depending on stellar mass, but typically $\\le$0.3 AU). Such close-in distances pose unique problems and constraints to habitability. For example, such planets are exposed to strong stellar winds in the early phases of the host star's evolution. Also, one expects a large number of coronal mass ejections (CMEs) on active M stars \\citep{Houdebine90,Scalo07}. The related interaction of the dense plasma flux (either from the stellar wind or a CME) with the atmosphere/magnetosphere environment of the exposed planets during the active stage of the stellar evolution could be strong enough to erode the atmosphere or the planets' water inventory via non-thermal atmospheric loss processes \\citep{Khodachenko07AB, Lammer07AB}. Weak magnetic shielding of tidally locked planets will also lead to an increased influx of galactic cosmic rays and stellar energetic particles \\citep{Griessmeier05a,Griessmeier09}, which can have important consequences for biological systems, both directly and indirectly via the modification of the planetary atmosphere \\citep[e.g.~destruction of atmospheric ozone, see][]{Grenfell07AB,Grenfell09}. This paper is organised as follows: The evolution of the conditions in the stellar vicinity over the timescale of the stellar evolution is presented in Section \\ref{sec:stellarvar}. Two cases are considered: stellar CMEs (Section \\ref{sec:stellarvar:CME}), and the average stellar wind (Section \\ref{sec:stellarvar:sw}). Planetary magnetic shielding against the influence of the star is discussed in Section \\ref{sec:magneticshield}. Consquences are discussed in Section \\ref{sec:consequences}, with special focus on the potential evaporation of the planetary atmosphere (Section \\ref{sec:consequences:atm}) and the flux of cosmic rays to the planetary atmosphere (Section \\ref{sec:consequences:cr}). Section \\ref{sec:conclusions} closes with a few concluding remarks. ", "conclusions": "\\label{sec:conclusions} Because of tidal locking, extrasolar planets with orbits in the habitable zone of K/M are likely to have much smaller magnetic moments than more distant planets. Combined with the strong ram pressure of dense and fast stellar winds and the high CME activity expected around young stars, this leads to the conclusion that such planets will have small magnetospheres which offer only limited magnetic protection. The weak magnetic protection may lead to strong atmospheric erosion during the early stages of stellar evolution, which could be strong enough to erode the atmosphere or the planets' water inventory via non-thermal atmospheric loss processes. Another consequence of the weak magnetic protection is a high flux of galactic cosmic rays and stellar energetic particles to the planetary atmosphere. This has implications for potential habitability, but also for the atmospheric chemistry and composition (e.g.~destruction of atmospheric ozone). These effects have also to be considered for missions studying biosignatures in the observed spectra of Earth-like exoplanets. Overall, the magnetic protection of exoplanets in the habitable zone of K/M stars can be weak, so that these planets may only be weakly shielded against stellar activity. {\\footnotesize" }, "1005/1005.0770_arXiv.txt": { "abstract": "We model the detectability of exoplanets around stars in the Beta Pic Moving Group (BPMG) using the Gemini Planet Imager (GPI), a coronagraphic instrument designed to detect companions by imaging. Members of the BPMG are considered promising targets for exoplanet searches because of their youth ($\\sim 12$ MY) and proximity (median distance $\\sim 35$ pc). We wrote a modeling procedure to generate hypothetical companions of given mass, age, eccentricity, and semi-major axis, and place them around BPMG members that fall within the V-band range of the GPI. We count as possible detections companions lying within the GPI's field of view and H-band fluxes that have a host-companion flux ratio placing them within its sensitivity. The fraction of companions that could be detected depends on their brightness at 12 Myr, and hence formation mechanism, and on their distribution of semi-major axes. We used brightness models for formation by disk instability and core-accretion. We considered the two extreme cases of the semi-major axis distribution - the log-normal distribution of the nearby F and G type stars and a power-law distribution indicated by the exoplanets detected by the radial velocity technique. We find that the GPI could detect exoplanets of all the F and G spectral type stars in the BPMG sample with a probability that depends on the brightness model and semi-major axis distribution. At spectral type K to M1, exoplanet detectability depends on brightness and hence distance of the host star. GPI will be able to detect the companions of M stars later than M1 only if they are closer than 10 pc. Of the four A stars in BPMG sample, only one has V-band brightness in the range of GPI; the others are too bright. ", "introduction": "The primary goal of the Gemini Planet Imager (GPI), a coronagraphic instrument under construction for the Gemini Observatory (Graham et al. 2007; see also {\\it Future Instrumentation} at www.gemini.edu), is the detection of exoplanets directly by imaging. Young stars will be important targets for GPI science because their exoplanets are expected to dim as they age and cool. Over the past 20 years astronomers have identified groups of relatively nearby young stars distinguished by their space motion through the galaxy, hence the name \u201cNearby Young Moving Groups\u201d (NYMGs). The TWA group at median distance 56 pc is the youngest, $\\sim 8$ MY, and the $\\beta$ Pic and AB Dor groups are the nearest with median distances 35 and 30 pc, respectively (Zuckerman and Song, 2004; Torres et al. 2006; Fern\\'andez et al. 2008). Their ages are about 12 MY for $\\beta$ Pic and 30-100 MY for AB Dor (Fern\\'andez et al. 2008). By analyzing the kinematics of the NYMGs, Fern\\'andez et al. (2008) showed that they share a common origin at the edges of the giant molecular cloud that formed the Sco-Cen star forming region. On this scenario, the supernova that heated the Local Bubble also triggered the formation of the stars in the NYMGs. Owing to their youth and proximity stars in the NYMGs are promising targets for exoplanet searches by high resolution and high contrast imaging. With detectable contrast ratios of $10^7$, the GPI is designed to image brown dwarfs and exoplanets that may surround a star. Astronomers can then spectrally analyze the composition of their atmospheres, which has further implications in astrobiology and stellar and planetary evolution (eg. Oppenheimer and Hinkley, 2009). L\\'epine and Simon (2009) described a pilot program to identify new low-mass probable members of the $\\beta$ Pic moving group (BPMG). Schlieder et al. (2009) are continuing the program and expanding it to include the AB Dor moving group. Our goals in the work described here were to evaluate the detectability of exoplanets in the BPMG and hence to provide guidance for observing programs of the BPMG, and other NYMGs, as new members are discovered. ", "conclusions": "We have found that for stars in the BPMG with $5 \\sim 80 M_{J}$." }, "1005/1005.5013_arXiv.txt": { "abstract": "{The iron lines at 630.15 and 630.25~nm are often used to determine the physical conditions of the solar photosphere. A common approach is to invert them simultaneously under the Milne-Eddington approximation. The same thermodynamic parameters are employed for the two lines, except for their opacities, which are assumed to have a constant ratio. } {We aim at investigating the validity of this assumption, since the two lines are not exactly the same.} {We use magnetohydrodynamic simulations of the quiet Sun to examine the behavior of the ME thermodynamic parameters and their influence on the retrieval of vector magnetic fields and flow velocities.} {Our analysis shows that the two lines can be coupled and inverted simultaneously using the same thermodynamic parameters and a constant opacity ratio. The inversion of two lines is significantly more accurate than single-line inversions because of the larger number of observables. } {} ", "introduction": "\\label{sec:intro} Both solar physics and radiative transfer owe many things to the Milne-Eddington (ME) approximation. Being an incredibly simplistic description of spectral line formation, it provides useful hints to understand the behavior of spectral lines as formed in solar and stellar atmospheres. It also offers a key diagnostics to infer the physical conditions of the plasma. This is particularly true when magnetic fields are present. The first solution of the radiative transfer equation in the presence of a magnetic field was derived adopting the ME assumption that all physical quantities relevant to line formation are constant with depth \\citep{1956PASJ....8..108U,1962aIzvKrAO...27.148R,1962aIzvKrAO...28.259R}. Under these conditions, the solution is analytic and, therefore, by simply varying the model parameters, one gets a handle on the behavior of Stokes line profiles. Similarly, the analytic character of the ME solution allows perturbative analyses like those performed by \\cite{1983SoPh...87..221L}, who studied the influence of velocity gradients, or \\cite{orozco2007}, who calculated the sensitivities of ME Stokes profiles to the various model parameters. Inversion codes of the radiative transfer equation are diagnostic tools that become simpler under the ME hypothesis. Since the pioneering work by \\cite{1972lfpm.conf..227H} and \\cite{1977SoPh...55...47A}, and after improvements by \\cite{1982SoPh...78..355L} and \\cite{1984SoPh...93..269L}, a number of ME inversion codes have been developed. These include the HAO code \\citep{lites2,1990ApJ...348..747L}, MELANIE \\citep{socas2001}, HELIX \\citep{2004A&A...414.1109L}, MILOS \\citep{orozco2007}, and VFISV \\citep{2010SoPh..tmp...35B}. Strictly speaking, the ME model is applicable to just one spectral line. The reason is that the so-called thermodynamic parameters of the model are meant to characterize the behavior of the specific line under consideration. The line-to-continuum opacity ratio, $\\eta_{0}$, the Doppler width of the line, $\\Delta \\lambda_{\\rm D}$, and the damping parameter, $a$, govern the shape of the Stokes profiles (i.e., they are the parameters of the Voigt and Faraday-Voigt functions). In turn, the source-fuction terms, $S_{0}$ and $S_{1}$, control the continuum level, the line depression, and the Stokes amplitudes. However, many investigations are based upon the simultaneous ME inversion of spectral line pairs, like the well-known \\ion{Fe}{i} doublet at 630~nm (e.g., Lites et al. 1993) or the \\ion{Mg}{i}~{\\em b} lines at 517.2 and 518.3~nm \\citep{1988ApJ...330..493L}. These inversions are reported to use no extra free parameters as compared to single-line inversions, based on the similarities between the two lines belonging to the same multiplet: the opacities for each line are specified in the ratio of their respective oscillator strengths, while the Doppler widths and the damping parameters are assumed to be identical for the two lines. $S_{0}$ and $S_{1}$ are also the same for both lines since they lie very close in wavelength. This strategy has been widely used for more than 20 years to analyze the \\ion{Fe}{i} 630 nm measurements taken with instruments such as the Advanced Stokes Polarimeter \\citep{1992SPIE.1746...22E} or the spectropolarimeter aboard the {\\em Hinode} satellite \\citep{2007ASPC..369...55L, 2007SoPh..243....3K, 2008SoPh..249..167T}. The simultaneous inversion has been shown to provide better results than single-line inversions \\citep[e.g.,][]{1994SoPh..155....1L}. The better performance is easy to understand: if the observables reproduced by the inversion are doubled, the results can be expected to be more accurate (at least by a factor $\\sqrt{2}$). However, the two lines do not have exactly the same atomic parameters (see Table\\ \\ref{cap5:tablelines}). This makes them to be formed at slightly different heights, where different physical conditions may exit \\citep[e.g.][]{2006A&A...456.1159M}. In view of these differences and of the large excursions of $\\eta_{0}$, $a$, and $\\Delta \\lambda_\\mathrm{D}$ in the real solar photosphere, both horizontally and vertically, one may wonder whether the simultaneous ME inversion of the two lines is valid and what the limitations are. General inversion codes not relying on the ME approximation perform exact line transfer calculations, so they are able to invert the two lines without inconsistencies. The purpose of the present Research Note is to check the ME case using the excellent test bench offered by modern magnetohydrodynamic (MHD) simulations. \\begin{table} \\caption{\\label{cap5:tablelines}Atomic data for the \\ion{Fe}{i} 630~nm lines. } \\centering \\begin{tabular}{@{}c c c c c c c@{}} \\hline $\\lambda_{0}$ (nm) & $\\chi$ (eV) & $\\log gf$ & {\\scshape Transition }& $\\alpha/a^2_0$ & $\\sigma$ & $g_\\mathrm{eff}$\\\\ \\hline \\hline 630.1501 & 3.654 & $-0.75$ & $5P_2-5D_2$ & 0.243 & 840.5 & 1.67 \\\\ 630.2494 & 3.686 & $-1.236$ & $5P_1-5D_0$ & 0.240 & 856.8 & 2.5 \\\\ \\hline \\end{tabular} \\tablefoot{ Shown are the central wavelength of the transition, $\\lambda_{0}$, the excitation potential of the lower atomic level, $\\chi$, the multiplicity of the lower level times the oscillator strength, $\\log gf$, the collisional broadening parameters, $\\alpha$ and $\\sigma$ (in units of Bohr's radius $a_0$), and the effective Land\\'e factor of the line, $g_\\mathrm{eff}$. The $\\log gf$ values have been derived from a fit to the solar spectrum using a two-component model of the quiet Sun \\citep{borrero2002}. } \\end{table} ", "conclusions": "" }, "1005/1005.3918_arXiv.txt": { "abstract": "The work presented in this paper aims at restricting the input parameter values of the semi-analytical model used in \\galics and \\momaf, so as to derive which parameters influence the most the results, \\eg star formation, feedback and halo recycling efficiencies, \\etc Our approach is to proceed empirically: we run lots of simulations and derive the correct ranges of values. The computation time needed is so large, that we need to run on a grid of computers. Hence, we model \\galics and \\momaf execution time and output files size, and run the simulation using a grid middleware: \\diet. All the complexity of accessing resources, scheduling simulations and managing data is harnessed by \\diet and hidden behind a web portal accessible to the users. ", "introduction": "\\label{sec:introduction} Cosmological simulations in this context are used to simulate the evolution of dark matter through cosmic time in various universes. A classical simulation begins with an N-body computation using for example \\ramses~\\cite{2002AA_385_337T} or \\gadget~\\cite{2005MNRAS_364_1105S}. The output of the simulation is then post-processed using semi-analytical models such as \\galics~\\cite{2003MNRAS_343_75H} (\\textbf{GAL}axies \\textbf{I}n \\textbf{C}osmological \\textbf{S}imulations), then mock catalogs of observed galaxies are produced using \\momaf~\\cite{2005MNRAS_360_159B} (\\textbf{Mo}ck \\textbf{Ma}p \\textbf{F}acility). Those models use as input a set of parameters, which influence the results, as for example: the galaxy luminosity function, the number of galaxies per observed cones, and more globally the history of star formation rate in the galaxy evolution process. Those parameters have a large range of values, the experiment here aims at reducing for each parameter the intervals of values to a subset of ranges within which the output simulations would be coherent with observed galaxies distributions. In order to stress how realistic the post-processing results are, a first important step is to identify the \\galics and \\momaf parameters which have the largest impact on the astrophysical results, such as star formation efficiency, feedback efficiency, halo recycling efficiency. As the parameters have a large range of possible values, exploring ``all'' combinations is really time consuming and requires lots of computing power. To harness the difficulty of running these analysis, one need firstly to have access to many computing resources, and secondly an efficient way to access those latter. Hence, we propose a client/server implementation for running post-processings on a distributed platform composed of heterogeneous machines: a \\emph{Grid}. A transparent access to these machines is provided by a \\emph{grid middleware}: \\diet (\\textbf{D}istribued \\textbf{I}nteractive \\textbf{E}ngineering \\textbf{T}oolbox). \\diet handles in one common and effective way the deployment of the computations on a grid of heterogeneous and distributed computers, the management of the different components of the post-processing, and also provides monitoring, communications and computation scheduling, and data and workflows management. Running efficiently the post-processings on a set of distributed and heterogeneous machines requires estimations on both the execution time and the amount of data to be transfered for each applications. Hence, we benchmarked \\galics and \\momaf and derived their execution time and output file size. Those models are used within \\diet to select which computer should run the post-processing. We first present in Section~\\ref{sec:cosmo} the cosmological simulation post-processing workflow, then we give the execution time and output file size models in Section~\\ref{sec:model}. In Section~\\ref{sec:ArchDIET} we give an overview of the \\diet middleware: its architecture, and the various features used within this project. Finally, and before concluding the paper, we present in Section~\\ref{sec:implem} the client/server implementation that allows the transparent execution of the post-processing workflow. ", "conclusions": "\\label{sec:conclusion} In this paper, we presented models for \\galaxymaker's and \\momaf's execution time, and for \\galaxymaker's output files's size. Execution time models give an estimation that is quite close to the real execution time for input files having a reasonable size. For small input files, the model returns an overestimate of the execution time, but this is easily explained by cache mechanisms at the hardware level, and can be taken into account in the model. We also modeled the post-processing workflow, and provided efficient means of executed it in a parameter sweep manner on a grid of computers. Our solution relies on the \\diet middleware that provides transparent access to resources, and data and workflows management. All the complexity of running the post-processing on a grid, and the creation of the workflows are hidden by a web interface, that provides an easy and user-friendly way of submitting such cosmological simulations post-processing on many heterogeneous and distributed resources. The next step is of course to analyze the results produced by the post-processing, and derive which parameters are the most influent on the results. Finally, we aim at providing ranges of values for the different parameters, that would provide correct results, \\ie comparable to observational data. \\begin{theacknowledgments} This work was developed with financial support from the ANR (Agence Nationale de la Recherche) through the LEGO project referenced ANR-05-CIGC-11. \\end{theacknowledgments}" }, "1005/1005.1406_arXiv.txt": { "abstract": "{ We perform spectral simulations of dynamo for magnetic Prandtl number of one with Taylor-Green forcing. We observe dynamo transition through a supercritical pitchfork bifurcation. Beyond the transition, the numerical simulations reveal complex dynamo states with windows of constant, periodic, quasiperiodic, and chaotic magnetic field configurations. For some forcing amplitudes, multiple attractors were obtained for different initial conditions. We show that one of the chaotic windows follows the period-doubling route to chaos.} \\begin{document} ", "introduction": " ", "conclusions": "" }, "1005/1005.4479_arXiv.txt": { "abstract": "For finite chemical potential effective models of QCD predict a first order phase transition. In favour for the search of such a phase transition in nature, we construct an equation of state for strange quark matter based on the MIT bag model. We apply this equation of state to highly asymmetric core collapse supernova matter with finite temperatures and large baryon densities. The phase transition is constructed using the general Gibbs conditions, which results in an extended coexistence region between the pure hadronic and pure quark phases in the phase diagram, i.e. the mixed phase. The supernovae are simulated via general relativistic radiation hydrodynamics based on three flavor Boltzmann neutrino transport in spherical symmetry. During the dynamical evolution temperatures above \\(10\\) MeV, baryon densities above nuclear saturation density and a proton-to-baryon ratio below \\(0.2\\) are obtained. At these conditions the phase transition is triggered which leads to a significant softening of the EoS for matter in the mixed phase. As a direct consequence of the stiffening of the EoS again for matter in the pure quark phase, a shock wave forms at the boundary between the mixed and the pure hadronic phases. This shock is accelerated and propagates outward which releases a burst of neutrinos dominated by electron anti-neutrinos due to the lifted degeneracy of the shock-heated hadronic material. We discuss the radiation-hydrodynamic evolution of the phase transition at the example of several low and intermediate mass Fe-core progenitor stars and illustrate the expected neutrino signal from the phase transition. ", "introduction": "The investigation of the QCD phasediagram via heavy-ion experiments at BNL's RHIC, CERN's LHC and the FAIR facility at GSI poised to explore the properties of QCD matter under extreme conditions. Three of the most important aspects are the behaviour and the position of the critical points in the phase diagram, the phase transition from hadronic matter to quark matter at finite chemical potentials and the properties of the predicted quark phases. In search for these aspects, observations of astronomical objects and astrophysical processes that are assumed to contain quark matter could be helpful. In such favour, Cold and isolated or binary neutrons stars (NS) have long been served as powerful objects to probe the equation of state (EoS) for hadronic as well as for quark matter. In the latter case the NS is called a hybrid star if in addition to the quark core a hadronic envelope is present. The astrophysical processes that leave an isolated NS as the final remnant are core collapse supernovae of low and intermediate mass Fe-core progenitor stars. Naturally the question rises at which stage during the dynamical evolution from a collapsing Fe-core to an isolated NS the thermodynamic conditions are such that the appearance of quark matter is favoured? Even further in the case of a phase transition from hadronic matter to quark matter, which hydrodynamic evolution can be expected and is there a relation to observations? The two intrinsically different scenarios are either during the early post bounce phase when the temperatures are moderately high or during the cooling of the remnant, where deleptonisation causes a temperature decrease and a density increase. The present article discusses the first case due to the critical conditions given by the quark matter EoS, where a relation to the explosion mechanism is explored. Therefore, a detailed study of core collapse supernovae including radiation-hydrodynamics with spectral neutrino transport and a sophisticated EoS for hot and dense asymmetric matter is required to simulate the matter conditions accurately. The first study of the QCD phase transition in core collapse supernovae was published by \\citet{TakaharaSato:1988}, suggesting a relation of the multi-peak neutrino signal from supernova 1987A (see \\citet{Hirata:etal:1988}) to the appearance of quark matter. Using general relativistic hydrodynamics, they modelled the phase transition via a polytropic EoS. Due to the absence of neutrino transport they could neither confirm nor exclude the suggested prediction of a neutrino signal from the phase transition. Additional microphysics was included in the study by \\citet{Gentile:etal:1993}, where general relativistic hydrodynamics is coupled to a description of deleptonisation during the collapse phase of a progenitor star. Applying the MIT-bag model for the description of the quark phase, they find the formation of a (second) shock wave as a direct consequence of an extended co-existence region between the hadronic phase and the quark phase with a significantly smaller adiabatic index. The second shock wave follows and merges with the first shock from the Fe-core bounce after a few milliseconds. However, due to the lack of neutrino transport in the post bounce phase, they were also not able to confirm the predictions made for a possible neutrino signal from the phase transition. The recent investigation by \\citet{Nakazato:etal:2008} is based on general relativistic radiation hydrodynamics with spectral three flavour Boltzmann neutrino transport. They investigate very massive progenitors of \\(\\sim 100\\) M\\(_\\odot\\) which collapse to a black hole. Applying the MIT-bag model for quark matter, the time after bounce for black hole formation is shortened and corresponds to the appearance of quark matter, where the central mass exceeds the maximum stable mass given by the quark EoS. We follow a similar approach and apply the MIT-bag model for the description of quark matter in general relativistic radiation hydrodynamics simulations, based on spectral three flavor Boltzmann neutrino transport. Our simulations are launched from low and intermediate mass progenitors, where no explosions could be obtained in spherical symmetry. We investigate the dynamical effects and discuss the possibility of an observable in the neutrino signal related to the QCD phase transition. The manuscript is organised as follows. In \\S 2 we will present the standard core collapse supernova scenario and in \\S 3 we will lay down our neutrino radiation hydrodynamics model including both the hadron and quark EoSs. In \\S4 we will discuss the appearance of quark matter during the early post bounce phase of core collapse supernovae of intermediate mass Fe-core progenitors and summarise the results in \\S 5. ", "conclusions": "\\begin{table}[htpb] \\begin{center} \\caption{Selected properties of the PNSs for the different progenitor models under investigation. } \\begin{tabular}[b]{ c c c c c c c c } \\hline \\hline Prog. & bag constant & \\(t_{pb}\\) & \\(M_\\text{PNS}\\) & \\(E_\\text{expl}\\) & \\(\\rho_\\text{c}\\) & \\(T_\\text{c}\\) & \\(Y_e\\) \\\\ \\([\\)M\\(_\\odot]\\)& \\([\\)MeV\\(]\\) & \\([\\)ms\\(]\\) & \\([\\)M\\(_\\odot]\\) & \\([10^{51}\\) erg\\(]\\) & \\([10^{14}\\) g/cm\\(^3]\\) & \\([\\)MeV\\(]\\) & \\\\ \\hline 10 & 162 & 248 & 1.434 & 0.361 & 6.6069 & 13.14 & 0.2343 \\\\ 10 & 165 & 429 & 1.482 & 1.080 & 5.8884 & 15.33 & 0.2488 \\\\ 13 & 162 & 241 & 1.467 & 0.146 & 6.4565 & 13.32 & 0.2335 \\\\ 13 & 165 & 431 & 1.498 & 0.323 & 5.6234 & 15.55 & 0.2462 \\\\ 15 & 162 & 209 & 1.608 & 0.420 & 6.7608 & 14.10 & 0.2262 \\\\ 15 & 165 & 330\\(^1\\) & 1.700 & unknown\\(^2\\) & 5.4954 & 15.33 & 0.2479 \\\\ \\hline \\end{tabular} \\end{center} \\small \\(^1\\) time of black hole formation\\\\ \\(^2\\) black hole formation before positive explosion energies are achieved \\label{table:quark-runs} \\end{table} The results of this first investigation of the QCD phase transition in core collapse supernova simulations of low and intermediate mass Fe-core progenitor stars is summarised in Table~\\ref{table:quark-runs}. The simulations are launched from progenitor stars of \\(10\\), \\(13\\) and \\(15\\) M\\(_\\odot\\) from \\citet{Woosley:etal:2002}. The post bounce times \\(t_{pb}\\) correspond to the appearance of the second neutrino burst in the spectra and the PNS masses \\(M_\\text{PNS}\\) are taken at the electron-neutrinospheres at some late post bounce times when the simulations are stopped. The central thermodynamic conditions, density \\(\\rho_\\text{c}\\), temperature \\(T_\\text{c}\\) and electron fraction \\(Y_e\\), correspond to the initial PNS collapse. A special model is the \\(15\\) M\\(_\\odot\\) progenitor using \\(B^{1/4}=165\\) MeV, where the maximum mass is reached shortly after the QCD phase transition. Hence, the PNS collapses to a black hole before the already formed second shock is accelerated to positive velocities. Due to our co-moving coordinate choice, no stable solution for the equations of energy and momentum conservation can be obtained and \\(t_{pb}\\) determines the time of black hole formation after bounce. The significant softening of the EoS for matter which is in the mixed phase causes the PNS to collapse. As a direct consequence of the softening of the EoS for matter which reaches the pure quark phase, a second shock wave forms. The second shock accelerates and finally merges with the first shock and launches an explosion. The explosion energies \\(E_\\text{expl}\\) are give in Table~\\ref{table:quark-runs}. This mechanism was explored and found to even produce explosions in spherical symmetry using general relativistic radiation hydrodynamics based on spectral three-flavour Boltzmann neutrino transport, where otherwise no explosion could be obtained. The explosion energies in Table~\\ref{table:quark-runs} are evaluated at some late post bounce times when the simulations are stopped and might have not yet converged to the final value of the explosion energy. The simulations will have to be carried out longer. However, moderate explosion energies of \\(\\simeq 10^{51}\\) erg could be found for the \\(10\\) M\\(_\\odot\\) model using \\(B^{1/4}=165\\) MeV (i.e. the later phase transition), otherwise lower explosion energies are obtained. The explosion energies might be shifted to larger values when multi-dimensional effects such as convection and rotation are taken into account. The second shock forms in the high density regime where neutrinos are fully trapped. Hence, no direct observational identification of the QCD phase transition can be found in the emitted neutrino spectra. It would be different if analysing the emitted gravitational waves directly from the phase transition (\\citet{Abdikamolov:etal:2009}). Unfortunately, gravitational waves have proven difficult to detect. Nevertheless, an indirect observation can be found in the emitted neutrino spectra. A second neutrino burst is released when the second shock, formed during the PNS collapse, propagates over the neutrinospheres. This second burst is, due to the lifted degeneracy of the shock heated hadronic material, dominated by electron-antineutrinos. The burst is accompanied by a significant increase of the mean neutrino energies which might become resolvable by neutrino detector facilities such as Super-Kamiokande for a future Galactic core collapse supernova explosion if quark matter appears. One of the most important observations from supernova explosions is the composition of the ejecta, which is determined by explosive nucleosynthesis investigations and are model dependent (see for example Fr\\\"ohlich et al. (2006a-c)). Of special interest is the production of heavy elements for which rapid neutron captures (\\(r\\)-process) in supernova explosion models have long been investigated (see for example \\citet{WoosleyBaron:1992}, \\citet{Woosley:etal:1994}, \\citet{Witti:etal:1994}, \\citet{Otsuki:etal:2000}, \\citet{Thompson:etal:2001}, Wanajo et~al.(2006a,b), \\citet{Arcones:etal:2007}, \\citet{PanovJanka:2009}). The abundances are calculated via post processing of mass trajectories from explosion models and compared with the well known solar abundances. Unfortunately, the very recent neutrino driven explosion models fail to provide conditions favourable for the \\(r\\)-process, which are high entropies per baryon, a fast expansion timescales and generally neutron rich material. Especially the latter aspect is a subject of research. Although the explosion models achieved via the QCD phase transition of the present article have to be analysed consistently with respect to explosive nucleosynthesis, a reasonable amount of ejected matter is found to be neutron-rich where \\(Y_e\\simeq0.35-0.45\\). In addition, moderate entropies per baryon and a fast expansion timescale are obtained. The conditions found may indeed be favourable for the \\(r\\)-process." }, "1005/1005.1892_arXiv.txt": { "abstract": "A large number of spectroscopic studies have provided evidence of the presence of multiple populations in globular clusters by revealing patterns in the stellar chemical abundances. This paper is aimed at studying the origin of these abundance patterns. We explore a model in which second generation (SG) stars form out of a mix of pristine gas and ejecta of the first generation of asymptotic giant branch stars. We first study the constraints imposed by the spectroscopic data of SG stars in globular clusters on the chemical properties of the asymptotic and super asymptotic giant branch ejecta. With a simple one-zone chemical model, we then explore the formation of the SG population abundance patterns focussing our attention on the Na-O, Al-Mg anticorrelations and on the helium distribution function. We carry out a survey of models and explore the dependence of the final SG chemical properties on the key parameters affecting the gas dynamics and the SG formation process. Finally, we use our chemical evolution framework to build specific models for NGC 2808 and M4, two Galactic globular clusters which show different patterns in the Na--O and Mg--Al anticorrelation and have different helium distributions. We find that the amount of pristine gas involved in the formation of SG stars is a key parameter to fit the observed O--Na and Mg--Al patterns. The helium distribution function for these models is in general good agreement with the observed one. Our models, by shedding light on the role of different parameters and their interplay in determining the final SG chemical properties, illustrate the basic ingredients, constraints and problems encountered in this self-enrichment scenario which must be addressed by more sophisticated chemical and hydrodynamic simulations. ", "introduction": "\\label{sec:introduction} In the last decade, a large number of spectroscopic and photometric studies have provided strong indication of the presence of multiple stellar populations in globular clusters. The spectroscopic evidence comes from the observed spreads and anticorrelations between the abundances of light elements, not shown by the halo stars of similar metallicities. Specifically, all the clusters studied have shown an anticorrelation between Na and O and an anticorrelation between Mg and Al has been found in several massive clusters \\citep[cf.][]{grasneca04}. The fact that these chemical anomalies have been found also in turnoff and subgiant stars \\citep{gratton2001, ramirezcohen2002} indicates that they must be already present in the gas from which these stars formed. The anticorrelation between CN and CH bands observed in giant \\citep[e.g.][]{norrisfd1984} and turnoff stars indicates a difference in the nitrogen abundance and is considered another fingerprint of the presence of multiple populations in globular clusters. An additional important constraint on the possible sources of gas from which second generation stars formed is provided by the observed constancy of the total C+N+O. The only exception, so far, is NGC 1851 for which a difference of ~0.6 dex is found among four red giants \\citep{yong2009} and for which the analysis of the subgiant branch suggests that the total C+N+O increases by about a factor three between the stars in the brighter and those in the dimmer branch \\citep{ventura18512009}. Photometric studies have also provided a number of important results indicating the presence of significant differences in the helium abundances of stars within individual clusters. Indeed, the presence of helium differences seems to be the only explanation for the multiple main sequences or spread in the main sequence (MS) observed in some clusters. These observations confirmed the prediction of the presence of a helium spread based on the complex morphology of the horizontal branches of some clusters made by \\citep{daca04,dc2008}. In extreme cases (the blue MS of \\ocen\\ and of NGC 2808) a helium content Y=0.38--0.40 is inferred \\citep{norris2004,dantona2005,pio07}. A recent spectroscopic survey by \\citet{carretta2009a} has found that all the 15 clusters studied show the spectroscopic evidence of the presence of multiple populations and that in all cases second generation stars represent a significant fraction (50-80 \\%) of a cluster population. This implies that matter forming more than half of each cluster stars has been processed through the hot CNO cycle and by other proton-capture reactions on light nuclei. In spite of the light elements variation, most GCs are mono--metallic objects, as far as abundances of heavier elements are concerned \\citep[see][for a recent review on this subject]{grasneca04}. Their heavy (Z $>$13) element metallicity, usually represented by the ratio [Fe/H], is found to be extremely homogeneous from star to star in each cluster. \\cite{carretta2009c} find that the upper limit to the scatter of iron is less than 0.05~dex in the 19 clusters they examine. This is consistent with previous determinations, e.g. for NGC~6752 \\citep[$\\sigma$=0.02][]{yong2005}, and for the clusters examined by \\cite{kraftivans2003} (0.03 -- 0.10~dex). The width in color of the main sequence and/or of the red giant branch also agree with the spectroscopic determination \\citep{suntzeff1993}. As the site of production of heavy elements ($\\alpha$--capture and Fe--group elements) are stars exploding as core-collapse or thermonuclear supernovae \\citep[e.g.][]{wheeler1989}, the star to star iron homogeneity means that supernova ejecta do not affect the chemistry of the gas processed through the hot CNO cycle. On the other hand, significant iron and s--process elements spread is well known to be present in \\ocen\\ \\citep[e.g., for a recent analysis,][and references therein]{johnson2009} and more recently it has been confirmed in other massive clusters like M22 \\citep[][and references therein]{Marino09, dacosta2009}, M54 \\citep[][and references therein]{sl1995, bellazzini2008}, and Terzan~5 \\citep{ferraro2009}. In smaller clusters like NGC~1851, the SG might be enriched in calcium, according to \\cite{Lee09}, but see the discussion by \\cite{carretta2010}. In \\citet{der08} (hereafter Paper I), we presented a model for the formation and dynamical evolution of multiple populations in globular clusters (GCs). In particular, we explored a model in which second generation (hereafter SG) stars form out of the ejecta of the first generation (hereafter FG) stars. By means of hydrodynamic simulations we have shown that the ejecta of AGB (asymptotic giant branch) stars collect in a cooling flow into the cluster core, where they form a subsystem of SG stars initially strongly concentrated in the cluster innermost regions. By means of N-body simulations we have then explored the subsequent stellar dynamical evolution of the cluster focusing our attention on the early loss of FG stars, on the evolution of the ratio of the number of FG and SG stars and on the evolution of their relative spatial distribution and mixing. Paper I also included a preliminary discussion on the chemical abundances of FG AGB stars and on the possibility of reproducing the helium excess and other chemical anomalies observed in SG stars. Several theoretical and observational papers have shown that dilution of the FG ejecta with gas having the pristine composition is needed to explain the general shape of the anticorrelations \\citep{prantzos2007, bekki07,dantonaventura2007,carretta2009a}. If the SG anomalies are inputed to processing by Hot Bottom Burning (HBB) at the bottom of the convective envelope of massive AGBs, these models produce a direct correlation between sodium and oxygen abundance in the processed matter. This direct correlation is unavoidable, as found in all the relevant AGB computations \\citep{dh2003,herwig04,karakas2007,ventura2008a} and as explained in more detail in Section 2. While the total sodium yield is dependent on many uncertain factors, it is certain that a small oxygen depletion (lower burning temperature) goes together with a larger sodium abundance and a large oxygen depletion (higher burning temperature) is accompanied by a smaller sodium abundance. Any attempt to reverse this direct correlation, in the limits provided by available cross sections, has been unfruitful \\citep[e.g.][]{ventura2008a}. An observational hint in favour of the above arguments is given by \\citet{carretta2009a} who find a direct correlation between the minimum O and the maximum Na abundances of the 15 clusters they studied; such a result could not be understood if O and Na of the AGB ejecta were anticorrelated. From all the above, it is apparent that, in order to explain the GC chemical patterns, it is necessary to introduce a form of dilution of the AGB ejecta, either with pristine gas, or with gas not showing the peculiar abundance patterns of the hot--CNO processed matter. The source of this pristine matter still requires a detailed understanding. \\cite{der08} proposed that initial asymmetries in the gas distribution allow to vent out the Type II supernovae (SNe~II) ejecta along preferential directions, creating an ``hour-- glass\" cavity and leaving some pristine gas in a torus at the outskirts of the cluster; at the end of the SN II activity this torus eventually collapses back into the cluster. In this scenario, the torus is not affected by contamination from supernova ejecta, and the iron content of the diluting matter is still the pristine one, consistently with the absence of iron differences between the FG and SG summarized above. While the evolution of the larger clusters is probably also affected by more prolonged star formation and supernovae type Ia contamination \\citep[e.g.][]{mar07}, the presence of slight metal increase in smaller clusters might be a possible indication that the ``pristine\" gas in the falling back torus has been partially contaminated by the SN~II ejecta. This paper is aimed at significantly expanding the initial study of SG stars chemical abundances presented in Paper I and at exploring in detail the SG chemical anomalies emerging from a model in which SG stars form from a mix of AGB ejecta and gas with pristine chemical composition. We will refer to the scenario presented in Paper I. Consequently, we need to associate to the temporal evolution of the FG clusters the chemical composition of the ejecta of super--AGB and then massive AGB stars (hereinafter simply referred to as AGB pollutors) that successively provide the matter used for building up second generation stars. The chemical composition of the H-rich matter in the envelopes of these stars is affected by the hot CNO cycle processing at the basis of the convective envelope of these giants (hot bottom burning, hereafter HBB) and possibly also by the third dredge up (3DU) occurring after each thermal pulse. Stellar winds and planetary nebula ejection deposit the matter of these envelopes into the interstellar medium (ISM), where it can contribute to the formation of the second generation. Despite the significant effort devoted to the determination of the yields of the massive AGBs and the numerous attempts to quantify the uncertainties in these calculations \\citep{ventura2005a,ventura2005b,karakas2007}, there are still large differences in the results obtained by different groups and even by the same groups at different times and no general consensus has been reached. Our aim is to focus on the chemical properties of the stellar envelopes and explore the constraints on these properties imposed by the spectroscopic data of SG stars in globular clusters. We have addressed this problem by means of a simple one-zone model in which the ISM is supplied by the FG AGB ejecta and by the possible accretion of pristine gas, and is depleted by the SG star formation (hereafter SF). This simplified approach allows to easily control the evolution of the amount of the AGB ejecta and pristine gas involved in the SG formation process and to fully explore the chemical properties of the ISM and of the SG stars. The structure of the paper is the following. We start in sect. \\ref{sec:yelds} with an overview of the chemical abundances of the massive AGB ejecta. We then illustrate our general assumptions in sect. \\ref{sec:assump} and the model details in sect. \\ref{sec:comp}. We then analyze the model in sect. \\ref{sec:moana}, and compare our results with the data of NGC 2808 and M4 in sect. \\ref{sec:ngc2808} and sect. \\ref{sec:m4}, respectively. Finally, we summarize our conclusions in sect. \\ref{sec:conc}. ", "conclusions": "\\label{sec:conc} In this paper we have studied the origin of the chemical patterns which have been observed in many globular clusters and which are considered the spectroscopic fingerprints of the presence of multiple stellar populations. Specifically, in our investigation we have focussed our attention on the O-Na and Mg-Al anticorrelations and the helium distribution function. In our model we have assumed that AGB stars are the polluting source of gas with the anomalous abundances of light elements observed in globular cluster second-generation stars. Our chemical framework is based on a one-zone model following the formation of SG stars from a mix of ejecta of AGB stars and pristine gas. We have carried out a large number of simulations to explore the dependence of the SG chemical properties on the parameters characterizing the star formation process and the dynamics of the involved pristine gas. Finally, we have used our framework to model the observed chemical patterns in the Galactic clusters NGC 2808 and M4. The main results of our study are the following: \\begin{enumerate} \\item The current stellar models provide AGB ejecta in which both Na and O decrease with increasing stellar mass. In order to reproduce the observed O-Na anticorrelation, the gas from which SG stars form must be diluted with O-rich, Na-poor pristine gas (see e.g. Figs. \\ref{fig:refmod} and \\ref{fig:refchm}). \\item The amount of pristine gas involved in the SG formation process, the timescales driving the dynamics of such gas, and the star formation efficiency play a key role in determining the extension of the O-Na anticorrelation and the fraction of extreme Na-rich/O-poor stars. \\item The helium abundance distribution function is correlated with the distribution of stars in the O-Na plane. Extreme Na-rich/O-poor stars are also those with an extreme He enrichement. Our models predict that all the clusters with a very extended O-Na anticorrelation should also host a population of He-rich stars. \\item Our models show that the extension of the O-Na anticorrelation is closely correlated with that in the Mg-Al plane; Na-rich/O-poor stars have also high Al and low Mg abundances. \\item We have used our framework to build specific models for two prototypical Galactic globular clusters: NGC 2808, a massive cluster that hosts a SG population characterized by a very extended O-Na anticorrelation and includes a very He-rich population, and M4, a low-mass cluster with a significantly less extended O-Na anticorrelation that does not include extremely Na-rich O-poor stars and for which there is no photometric evidence of a He-rich population. Despite the significant differences in their chemical patterns, in both clusters a significant fraction of stars belong to the SG \\citep[50 per cent in NGC 2808,] [and 65 per cent in M4 \\cite{carretta2009a,mar08})]{carretta2009a}. Our models successfully reproduce the differences in the O-Na anticorrelation observed in these two clusters, the distribution of stars in the Mg-Al plane in M4 and predict an extended Mg-Al anticorrelation for NGC 2808. Al and Mg abundances for NGC 2808 have been determined with UVES observations only for a small number of either Na-rich/O-poor or Na-poor/O-rich stars. As predicted by our model, these stars populate only the Al-rich/Mg-poor and the Al-poor/Mg-rich regions of the Mg-Al plane. According to our model, future determinations of Al and Mg abundances of targeting stars with intermediate Na and O should lead to intermediate Al and Mg abundances filling the extended Al-Mg anticorrelation. \\item The helium distribution, including the extreme population formed directly from super--AGB ejecta is well reproduced by our model for NGC~2808 (see Fig. \\ref{fig:2808}). In clusters like M4, in which a larger dilution with pristine matter is necessary to model the O-Na and Mg-Al patterns, a large helium dispersion is not required, but a small helium spread should generally be present (see Fig. \\ref{fig:m4}). \\end{enumerate} Our investigation has shed light on the key chemical and hydrodynamical ingredients determining the formation of the chemical patterns observed in globular clusters. The results presented here are to be the starting point informing further study based on full hydrodynamical simulations. Several issues will require further investigation. Specifically, as for the stellar evolution models, we have shown that our conservative choice of an educated linear and monotonic extrapolation for the sodium and oxygen abundances in the mass range from 6.5 to 9 $\\msun$ reproduces the general trend of the O--Na anticorrelations of different clusters, when also the differences in neon abundances in the FG of different clusters are taken into account. In NGC~2808, however, we can not reproduce the (few) very large sodium values of some stars for which only upper limits on the oxygen abundance are available (see Fig. \\ref{fig:2808}). Another parameter is probably at work here, and only further investigation into the super--AGB phase will shed light on this issue. The lack of any evidence of a significant metal enrichment in most clusters hosting multiple stellar populations implies that neither ejecta from FG nor SG supernovae is involved in the chemical enrichment of matter from which SG stars form. While in our model SG formation starts after the end of the FG SN II epoch, it is to be further clarified whether the lack of metal enrichment from SG supernovae is due to SG forming with IMF truncated at $M<9$ $\\msun$ or whether this is a consequence of a more complex gas dynamics to be further explored with full hydrodynamical simulations. Finally, additional full hydrodynamical simulations will be needed to clarify the source of pristine gas and the accretion mechanism. Different processes have been considered in the literature, such as accretion from a diffuse surrounding medium \\citep[e.g.][]{limu07,pflkro09} or interaction with molecular clouds \\citep{bekmac09}. Conditions for an effective gas collection within the cluster turn out to be rather specific. Instead, our model requires a mechanism general enough to work in all the clusters and able to take into account the cluster-to-cluster differences in the dynamics and the amount of pristine gas involved in the SG formation process." }, "1005/1005.2497_arXiv.txt": { "abstract": "{The properties of the dust grains (e.g., temperature and mass) can be derived from fitting far-IR SEDs ($\\geq$100~\\micron). Only with SPIRE on {{\\it Herschel}} has it been possible to get high spatial resolution at 200 to 500~\\micron\\ that is beyond the peak ($\\sim$160~\\micron) of dust emission in most galaxies.} {We investigate the differences in the fitted dust temperatures and masses determined using only $<$200~\\micron\\ data and then also including $>$200~\\micron\\ data (new SPIRE observations) to determine how important having $>$200~\\micron\\ data is for deriving these dust properties.} {We fit the 100 to 350~\\micron\\ observations of the Large Magellanic Cloud (LMC) point-by-point with a model that consists of a single temperature and fixed emissivity law. The data used are existing observations at 100 and 160~\\micron\\ (from IRAS and {{\\it Spitzer}}) and new SPIRE observations of 1/4 of the LMC observed for the HERITAGE Key Project as part of the {{\\it Herschel}} Science Demonstration phase.} {The dust temperatures and masses computed using only 100 and 160~\\micron\\ data can differ by up to 10\\% and 36\\%, respectively, from those that also include the SPIRE 250 \\& 350~\\micron\\ data. We find that an emissivity law proportional to $\\lambda^{-1.5}$ minimizes the 100--350~\\micron\\ fractional residuals. We find that the emission at 500~\\micron\\ is $\\sim$10\\% higher than expected from extrapolating the fits made at shorter wavelengths. We find the fractional 500~\\micron\\ excess is weakly anti-correlated with MIPS 24~\\micron\\ flux and the total gas surface density. This argues against a flux calibration error as the origin of the 500~\\micron\\ excess. Our results do not allow us to distinguish between a systematic variation in the wavelength dependent emissivity law or a population of very cold dust only detectable at $\\lambda \\geq 500~\\micron$ for the origin of the 500~\\micron\\ excess.} {} ", "introduction": "Among nearby galaxies, the Large Magellanic Cloud (LMC) and Small Magellanic Cloud (SMC) represent unique astrophysical laboratories for interstellar medium (ISM) studies. Both Clouds are relatively nearby, the LMC at $\\sim$50 kpc \\citep{Schaefer08} and the SMC at $\\sim$60 kpc \\citep{Hilditch05}, and provide ISM measurements that are relatively unconfused by line-of-sight uncertainties when compared to the Milky Way. The two Clouds span an interesting metallicity range with the LMC at $\\sim$1/2 Z$_{\\sun}$ \\citep{Russell92} being above the threshold of 1/3--1/4 Z$_{\\sun}$ where the properties of the ISM change as traced by the rapid reduction in the PAH dust mass fractions and possible dust-to-gas ratios \\citep{Draine07} and the SMC at $\\sim$1/5Z$_{\\sun}$ \\citep{Russell92} below this threshold. Finally, the dust in the LMC and SMC shows strong variations in its ultraviolet characteristics \\citep{Gordon03}. The HERschel Inventory of The Agents of Galaxy Evolution (HERITAGE) in the Magellanic Clouds {\\it Herschel} Key Project will map both Clouds using the PACS/SPIRE Parallel observing mode providing observations at 100, 160, 250, 350, and 500~\\micron\\ \\citep{Meixner10}. The HERITAGE wavelength coverage (100--500~\\micron) and spatial resolution ($\\sim$10~pc at 500~\\micron) is well suited to measuring the spatial variations of dust temperatures and masses. The infrared dust emission in most galaxies peaks between 100--200~\\micron\\ \\citep{Dale05} and observations $>$200~\\micron\\ are important for accurate dust temperature and masses \\citep{Willmer09}. Ground-based submilimeter observations do provide the needed $>$200~\\micron\\ observations, but they have been seen to be in excess of that expected from extrapolating fits to the $<$200~\\micron\\ data for sub-solar metallicity galaxies \\citep{Galliano05}. This excess could be due to very cold dust that only emits at submilimeter wavelengths or variations in the wavelength dependent dust emissivity law \\citep{Reach95, Paradis09}. As part of the Science Demonstration Program (SDP), two HERITAGE AORs centered on the LMC were executed. These observations are used in this paper to explore the impact SPIRE observations have on the measurement of dust temperatures and masses including the behavior of any submilimeter excess. ", "conclusions": "We investigate the importance of $>$200~\\micron\\ data in determining dust temperatures and masses using new {\\it Herschel} SPIRE observations of the LMC (taken for the HERITAGE Key Project as part of the {\\it Herschel} Science Demonstration phase) combined with existing IRAS 100~\\micron\\ and {\\it Spitzer} MIPS 160~\\micron\\ images. We fit the observations with a model consisting of dust emitting as a single temperature blackbody modified with an emissivity law proportional to $\\lambda^{-\\beta}$. For fixed values of $\\beta$, fits using only the 100--160~\\micron\\ data give dust temperatures and masses that are on average up to 8\\% and 23\\% different from fits using the same $\\beta$ and the 100--350~\\micron\\ data. The new SPIRE observations allowed us to determine that $\\beta = 1.5$ minimizes the residuals from 100 to 350~\\micron. Using a $\\beta = 2.0$ for the 100--160~\\micron\\ and a $\\beta = 1.5$ for the 100--350~\\micron\\ fits results in an increase of 10\\% for the dust temperature and a decrease in the dust mass by 36\\%. On average, there is a fractional excess at 500~\\micron\\ of $\\sim$10\\%. The origin of the fractional excess is unlikely to be due to our fitting algorithm or a flux calibration error, but it could be due to either very cold dust that emits only $\\geq$500~\\micron\\ or a variation in the wavelength dependent change in the dust emissivity. Planned HERITAGE observations of the LMC and SMC will allow for a more detailed investigation of including $> 200~\\micron$ data (mainly the 500~\\micron\\ excess) due to better quality PACS and SPIRE images (optimized observations and cross-scans)." }, "1005/1005.2342_arXiv.txt": { "abstract": "{Cyg OB2 \\#5 is a contact binary system with variable radio continuum emission. This emission has a low-flux state where it is dominated by thermal emission from the ionized stellar wind and a high-flux state where an additional non-thermal component appears. The variations are now known to have a period of 6.7$\\pm$0.2 yr. The non-thermal component has been attributed to different agents: an expanding envelope ejected periodically from the binary, emission from a wind-collision region, or a star with non-thermal emission in an eccentric orbit around the binary. The determination of the angular size of the non-thermal component is crucial to discriminate between these alternatives. We present the analysis of VLA archive observations made at 8.46 GHz in 1994 (low state) and 1996 (high state), that allow us to subtract the effect of the persistent thermal emission and to estimate an angular size of $\\leq 0\\rlap.{''}02$ for the non-thermal component. This compact size favors the explanation in terms of a star with non-thermal emission or of a wind-collision region. } ", "introduction": "Cyg OB2 \\#5 (V729 Cyg, BD +40 4220) is an eclipsing, contact binary system consisting of two O-type supergiants with a 6.6-day period (Hall 1974; Leung \\& Schneider 1978; Rauw et al. 1999; Linder et al. 2009). As several other luminous O-star systems in the Cyg OB2 association, this source was found to show evidence of variable radio emission (Persi et al. 1985, 1990; Bieging et al. 1989). The radio emission appears to have two states: a low-flux state of $\\sim$2 mJy at 4.8 GHz where the spectral index is consistent with thermal emission from an ionized stellar wind, and a high-flux state of $\\sim$8 mJy at 4.8 GHz where the spectral index is flatter than in the low state. The variations were proposed to have a 7-year period (Miralles et al. 1994) and have been attributed to variable non-thermal emission from an expanding envelope arising periodically in the binary (Persi et al. 1990; Bieging et al. 1989; Miralles et al. 1994). In addition, Abbott et al. (1981) and Miralles et al. (1994) reported on the existence of a radio ``companion'' 0.8 arcsec to the NE of the main radio source (which is associated with the eclipsing binary). Observations by Contreras et al. (1997) revealed that this radio source has an elongated shape and lies in-between the short-period binary and a third star, which was first reported by Herbig (1967). Contreras et al. (1997) suggested that the proposed NE radio companion actually corresponds to the wind interaction zone between the binary system and the tertiary component. Recently, Kennedy et al. (2010) reanalysed all VLA observations of Cyg OB2 \\#5 and showed that the primary radio source, associated with the eclipsing binary, varies with a period of $6.7 \\pm 0.2$ yr while the flux from the secondary NE source remains constant in time. These authors proposed that the variations in the main radio component can be represented by a simple model in which a fourth star (with constant non-thermal emission) moves around the eclipsing binary in an eccentric orbit and the varying radio emission results from the variable free-free opacity of the wind between the star and the observer (Kennedy et al. 2010). These authors estimate a major axis of 14 AU for the star with non-thermal emission. In this scenario, we have a quadruple star system: the contact binary, the star with non-thermal emission in a 6.7 yr orbit around the contact binary, and the NE component. To advance our understanding of the nature of the time-variable non-thermal emission, a determination of the angular size of the source producing it is needed. Using VLA data taken at 8.46 GHz in 1991 October 3 (when Cyg OB2 \\#5 was in the high radio state), Miralles et al. (1994) estimated an angular size of $0\\rlap.{''}046 \\pm 0\\rlap.{''}006$ for the whole (non-thermal plus thermal components) emission. Using MERLIN data taken at 4.8 GHz in 1996 November 14 (when Cyg OB2 \\#5 was again in the high radio state), Kennedy et al. (2010) estimated an angular size of $\\sim0\\rlap.{''}077$ for the whole (non-thermal plus thermal components) emission associated with the primary radio component. Since the thermal contribution comes from the ionized wind that is known to be extended, these size estimates can be considered as upper limits to the size of the non-thermal emission. In this paper we present the analysis of archive 8.46 GHz continuum observations made with the Very Large Array (VLA) of the NRAO\\footnote{The National Radio Astronomy Observatory is operated by Associated Universities Inc. under cooperative agreement with the National Science Foundation.} in the A configuration toward Cyg OB2 \\#5 in two epochs. During 1994 April 9 (1994.27), the source was in the low radio state and only the thermal emission from the ionized stellar wind was present. These observations were used to estimate the characteristics of the persistent thermal component. During 1996 December 28 (1996.99), the source was in the high radio state, with both the thermal and non-thermal components present. Using the information obtained from the 1994 data, it is possible to subtract in the \\sl(u,v) \\rm plane the contribution from the thermal component and obtain a more stringent estimate of the angular size of the non-thermal emission. ", "conclusions": "We present the analysis of VLA archive data of the Cyg OB2 \\#5 system taken during the high (1996) and low (1994) radio states. This analysis allows the subtraction, in the \\sl (u,v) \\rm plane of the persistent thermal component and a better estimate of the angular dimensions of the non-thermal component. We obtain an upper limit, $\\theta('') \\leq 0\\rlap.{''}02$ for the angular size of the non-thermal emission. Cyg OB2 \\#5 will be again in high state during 2010 and 2011 and a very long baseline interferometry study is needed to determine the angular size and morphology of the compact non-thermal emission." }, "1005/1005.1207_arXiv.txt": { "abstract": "{The question of how much gas cools in the cores of clusters of galaxies has been the focus of many, multiwavelength studies in the past 30 years. In this letter we present the first detections of the strongest atomic cooling lines, [C{\\sc ii}], [O{\\sc i}] and [N{\\sc ii}] in two strong cooling flow clusters, A1068 and A2597, using {\\it Herschel} \\rm PACS. These spectra indicate that the substantial mass of cold molecular gas ($>10^{9}$M$_\\odot$) known to be present in these systems is being irradiated by intense UV radiation, most probably from young stars. The line widths of these FIR lines indicate that they share dynamics similar but not identical to other ionised and molecular gas traced by optical, near-infrared and CO lines. The relative brightness of the FIR lines compared to CO and FIR luminosity is consistent with other star-forming galaxies indicating that the properties of the molecular gas clouds in cluster cores and the stars they form are not unusual. These results provide additional evidence for a reservoir of cold gas that is fed by the cooling of gas in the cores of the most compact clusters and provide important diagnostics of the temperature and density of the dense clouds this gas resides in.} \\date{Received 30 March 2010/Accepted} ", "introduction": "The cooling process at the cores of galaxy clusters is highly complex: recent {\\it XMM-Newton} and {\\it Chandra} observations indicate that the cooling rates are reduced by an order of magnitude below the simple cooling flow models at temperatures below $\\sim2\\times10^7$K (Peterson \\& Fabian 2006). These X-ray observations, when linked with the detection of radio jet inflated bubbles in the cores of many of the strongest cooling flows (see McNamara \\& Nulsen 2007 for a review), suggest that the strong suppression of gas cooling is related to energy injection into the intracluster medium by the action of jets and related AGN activity. The detection of substantial masses of molecular gas in the cores of the most rapidly cooling clusters through CO lines (Edge 2001, Salom\\'e \\& Combes 2003) and warm H$_2$ molecular lines in the NIR and MIR (Jaffe \\& Bremer 1997, Egami et al. 2006) indicates that not all cooling is suppressed and this cooled gas may provide the fuel for future AGN activity. These tracers of molecular gas appear to correlate with the strength of optical lines from ionised gas (Crawford et al. 1999, Edge 2001) and the dust continuum at MIR and sub-mm wavelengths (O'Dea et al. 2008). However, the excitation of these various emission lines and the relative importance of energy input from star formation, AGN, cosmic rays and/or the intracluster medium is poorly constrained (Ferland et al. 2009). One as yet unexplored diagnostic of the properties of the cold gas are the atomic cooling lines found in the FIR, [C{\\sc ii}], [O{\\sc i}] and [N{\\sc ii}]. The unprecedented sensitivity of {\\it Herschel} \\rm (Pilbratt et al. 2010) to FIR line emission offers the opportunity to assess the ionisation and density of the colder gas for the first time with the [C{\\sc ii}] line and two principle [O{\\sc i}] lines. The authors were awarded 140~hours of time in an Open Time Key Project (PI Edge) to investigate the FIR line and continuum properties of a sample of 11 brightest cluster galaxies (BCGs) in well-studied cooling flow clusters selected on the basis of optical emission line and X-ray properties. The full goals of the project are to observe at least five atomic cooling lines for each object that cover a range in density and temperature behaviour and obtain a fully sampled FIR spectral energy distribution for systems where significant star formation is expected. In this paper we present the Photodetector Array Camera \\& Spectrometer (PACS, Poglitsch et al. 2010) spectroscopy for the two targets observed in the Science Demonstration Phase (SDP), Abell~1068 ($z=0.1386$) and Abell~2597 ($z=0.0821$). In a parallel paper (Edge et al. 2010), we present the FIR photometry for these clusters. The two clusters observed have quite contrasting multiwavelength properties. A1068 is a strong MIR source (O'Dea et al 2008) with a bright CO detection (Edge 2001) but a weak radio source (McNamara et al 2004). A1068 lies just below the luminosity threshold of a ULIRG (10$^{12}$~L$_\\odot$) and exhibits some contribution from an AGN (Crawford et al. 1999, O'Dea et al. 2008). On the other hand, A2597 is a relatively weak MIR source (Donahue et al. 2007) with a weak CO detection (Salom\\'e, priv. comm.) but a powerful radio source (Sarazin et al. 1995). The implied FIR luminosity of A2597 of 8.8$\\times 10^{9}$~L$_\\odot$ is a factor of around 30 below that of A1068 (3.5$\\times 10^{11}$~L$_\\odot$) and, in addition, the fractional contribution from an AGN in the MIR is also lower. ", "conclusions": "These initial results from {\\it Herschel} \\rm indicate that atomic cooling lines are present in the brightest cluster galaxies in cooling flow clusters. The intensity and velocity width of these lines is consistent with all the other observed tracers of cold gas in these systems implying they originate from the same population of clouds. The only apparent exception to this in our current observations is that the FIR lines appear to be systematically broader than the CO lines impling that the {\\it relative} intensity of these lines varies with position within the BCG. The results that will come from our Open Time Key Project for 11 BCGs will expand greatly on those presented here with more lines and a greater range of BCG properties. Beyond this, the potential for {\\it Herschel} \\rm to illuminate the properties of the cold gas that may fuel cold nuclear accretion in more distant clusters and local groups is vast." }, "1005/1005.1031_arXiv.txt": { "abstract": "{During an X-ray survey of the Small Magellanic Cloud, carried out with the XMM-Newton satellite, we detected significant soft X-ray emission from the central star of the high-excitation planetary nebula SMP SMC 22. Its very soft spectrum is well fit with a non local thermodynamical equilibrium model atmosphere composed of H, He, C, N, and O, with abundances equal to those inferred from studies of its nebular lines. The derived effective temperature of 1.5$\\times10^5$ K is in good agreement with that found from the optical/UV data. The unabsorbed flux in the 0.1--0.5 keV range is $\\sim3\\times10^{-11}$ erg cm$^{-2}$ s$^{-1}$, corresponding to a luminosity of $\\sim1.2\\times10^{37}$ erg s$^{-1}$ at the distance of 60 kpc. We also searched for X-ray emission from a large number of SMC planetary nebulae, confirming the previous detection of SMP SMC 25 with a luminosity of (0.2--6)$\\times10^{35}$ erg s$^{-1}$ (0.1-1 keV). For the remaining objects that were not detected, we derived flux upper limits corresponding to luminosity values from several tens to hundreds times smaller than that of SMP SMC 22. The exceptionally high X-ray luminosity of SMP SMC 22 is probably due to the high mass of its central star, quickly evolving toward the white dwarf's cooling branch, and to a small intrinsic absorption in the nebula itself.} \\keywords {planetary nebulae: individual: SMP SMC 22, SMP SMC 25 - Magellanic Clouds - X-rays: general} ", "introduction": "Planetary nebulae (PNe) are a common stage in the evolution of low and intermediate mass stars, leading to the formation of white dwarfs. They appear when the fast wind from the central star interacts with the matter of the denser wind that was previously ejected during the asymptotic giant branch (AGB) phase, and are characterized by H$_{\\alpha}$ line emission. Since the advent of imaging X-ray telescopes a number of PNe have been detected in the soft X-ray range (see, e.g., \\citealt{chu03}), but due to their relatively low fluxes, detailed studies with the \\xmm\\ and Chandra satellites have been carried out only for a few objects \\citep{gru06,kas07,mon09}. X-ray emission from PNe can originate from their central star or from the hot gas shocked in the interaction between the two stellar winds. Usually one of these two processes, characterized by different spectral and spatial signatures, is dominant, but there are also PNe in which both components have been detected. There is also the possibility that some X-ray emission observed from PNe is actually due to coronal emission from a companion star (see, e.g., \\citealt{sok02}). SMP SMC 22 (hereinafter \\sm ) is a high-excitation planetary nebula located in the Small Magellanic Cloud (SMC) \\citep{san78,all87} and characterized by a very high X-ray luminosity \\citep{wan91}. Its large X-ray flux led to an early detection of this source in the soft X-ray range with the Einstein Observatory \\citep{sew81} and to its inclusion in the class of super-soft X-ray sources (SSS), a rather heterogeneous group of luminous (10$^{36}$--10$^{38}$ erg s$^{-1}$) sources characterized by thermal-like emission corresponding to blackbody temperatures of $10^5$--10$^6$ K (see, e.g., \\citealt{kah06}). Most SSS are binary systems containing accreting white dwarfs, but a few of them have been identified with the nuclei of PNe. Here we report on recent X-ray observations of \\sm\\ obtained with the \\xmm\\ satellite, as well as on a systematic search for X-ray emission from a large sample of PNe in the SMC. ", "conclusions": "The first X-ray observations of the SMC planetary nebula \\sm\\ obtained with a modern high-throughput satellite have allowed us to study its X-ray emission with unprecedented statistics. No evidence for a binary nature, such as long or short term variability, as seen in other SSS was found. It is remarkable that, despite different spectral models can fit the data, a self-consistent picture in terms of temperature, mass and radius of the central star can be obtained with a NLTE model atmosphere with the same elemental abundances seen in the nebula. The inferred mass for the central star, of the order of 1 $\\msun$, implies that \\sm\\ is the descendent of a relatively massive progenitor (see Fig. \\ref{HR}). This may explain its exceptional luminosity, as well as the apparent rarity of such objects that evolve very quickly toward the cooling white dwarfs sequence." }, "1005/1005.4422_arXiv.txt": { "abstract": "With the aim of determining if Milky Way (MW) progenitors could be identified as high redshift Lyman Alpha Emitters (LAEs) we have derived the intrinsic properties of $z\\approx 5.7$ MW progenitors, which are then used to compute their observed Ly$\\alpha$ luminosity, $L_\\alpha$, and equivalent width, EW. MW progenitors visible as LAEs are selected according to the canonical observational criterion, $L_\\alpha>10^{42}$~erg~s$^{-1}$ and $EW>20$ \\AA. Progenitors of MW-like galaxies have $L_{\\alpha} =10^{39-43.25}$~erg~s$^{-1}$, making some of them visible as LAEs. In any single MW merger tree realization, typically only 1 (out of $\\approx 50$) progenitor meets the LAE selection criterion, but the probability to have {\\it at least} one LAE is very high, $P=68\\%$. The identified LAE stars have ages, $t_* \\approx 150-400$~Myr at $z\\approx 5.7$ with the exception of five small progenitors with $t_* < 5$~Myr and large $EW=60-130$~\\AA. LAE MW progenitors provide $> 10\\%$ of the halo very metal-poor stars [Fe/H]$<-2$, thus establishing a potentially fruitful link between high-$z$ galaxies and the Local Universe. ", "introduction": "\\label{intro} Lyman Alpha Emitters (LAEs) are galaxies identified by means of a very strong Ly$\\alpha$ line (1216 \\AA). Advances in instrument sensitivity and specific spectral signatures (strength, width and continuum break bluewards of the line) have enabled the confirmation of hundreds of LAEs in a wide redshift range, at $z\\approx 2.25$ (Nilsson et al. 2008), $z\\approx3$ (Cowie \\& Hu 1998; Steidel et al. 2000; Matsuda et al. 2005; Venemans et al. 2007; Ouchi et al. 2008), $z\\approx 4.5$ (Finkelstein et al. 2007), $z\\approx5.7$ (Malhotra et al. 2005; Shimasaku et al. 2006) and $z\\approx 6.6$ (Taniguchi et al. 2005; Kashikawa set al. 2006). LAEs have by now been used extensively as probes of both the ionization state of the intergalactic medium (IGM) and probes of high redshift galaxy evolution (Santos 2004; Dayal, Ferrara \\& Gallerani 2008; Nagamine et al. 2008; Dayal et al. 2009; Dayal, Ferrara \\& Saro 2010; Dayal, Maselli \\& Ferrara 2010). However, in spite of the growing data sets, there has been no effort to establish a link between the properties of these early galaxies to observations of the local Universe, {\\it in primis} the Milky Way (MW). Our aim in this work is to investigate the possible connection between the Galactic building blocks and LAEs at a time when the Universe was $\\approx 1$ Gyr old. This will allow us to answer to questions as: are the progenitors of MW-like galaxies visible as LAEs at high redshifts? How can we discriminate amongst LAEs which are possible MW progenitors and those that are not? What are the physical properties of these Galactic building blocks? To this end, we adopt a novel approach of coupling the semi-analytical code {\\tt GAMETE} (Salvadori, Schneider \\& Ferrara 2007; Salvadori, Ferrara \\& Schneider 2008, Salvadori \\& Ferrara 2009), which traces the hierarchical build-up of the Galaxy, successfully reproducing most of the observed MW and dwarf satellite properties at $z=0$, to a previously developed LAE model (Dayal, Ferrara \\& Gallerani 2008; Dayal et al. 2009; Dayal, Ferrara \\& Saro 2010), that reproduces a number of important observational data sets accumulated for high-z LAEs. ", "conclusions": "\\label{discussion} We have linked the properties of high-$z$ LAEs to the Local Universe by coupling the semi-analytical code {\\tt GAMETE} to a previously developed LAE model. According to our results {\\it} the progenitors of MW-like galaxies cover a wide range of observed Ly$\\alpha$ luminosity, $L_{\\alpha} =10^{39-43.25}$ erg~s$^{-1}$, with $L_{\\alpha}$ increasing with $M_*$ (or, equivalently, $M_h$); hence some of them can be observed as LAEs. In each hierarchical merger history we find that, on average, {\\it only one} star-forming progenitor (among $\\approx 50$) is a LAE, usually corresponding to the major branch of the tree ($M_h\\approx 10^{10}\\Msun$). Nevertheless, the probability to have {\\it at least} one visible progenitor in any merger history is very high ($P=68\\%$). Interestingly, we found that the LAE candidates can be also observed as dropout galaxies since their UV magnitudes are always $M_{UV} < -18$. On average the identified LAE stars have intermediate ages, $t_* \\approx 150-400$~Myr, and metallicities $Z\\approx 0.3-1 \\Zsun$; an exception is represented by five newly formed galaxies, which are hosted by small DM halos, $M_h\\approx 10^9\\Msun$, which are yet visible as (faint) LAEs ($L_{\\alpha}\\approx 10^{42.05}$~erg~s$^{-1}$) by virtue of their high star formation rate and extremely young stellar population, $t_* < 5$~Myr. The low metallicity of these young galaxies, $Z\\approx 0.016-0.044 \\Zsun$, reflects that of the MW environment at their formation epoch. Although rare (5 out of 80 LAEs), these Galactic building blocks could be even more unambiguously identified among the least luminous LAEs, due to their larger Ly$\\alpha$ equivalent widths, $EW = 60-130$~\\AA\\ with respect to those ($\\approx 40$~\\AA) shown by older LAEs\\footnote{Due to the low dust mass, $f_c= f_\\alpha \\approx1$ for all the progenitors we identify as LAEs; the observed EW is solely governed by $T_\\alpha$, which leads $EW \\approx 40$ \\AA\\, for almost all LAEs, except those with $t_*\\leq 10$ Myr.}. These small and recently virialized halos ($z_{vir}\\lsim 6$) could be the progenitors of Fornax-like dwarf spheroidal galaxies (see Fig.~1 of Salvadori \\& Ferrara 2009). By identifying these LAEs, therefore, it would be possible to observe the most massive dSphs of the MW system just at the time of their birth. Uncertainties remain on the treatment of dust in calculating the LF and observed properties of the MW progenitor LAEs identified here, especially at the low luminosity end of the LF. Several aspects require additional study. As gas, metal and dust are preferentially lost from low mass halos, pushing the mass resolution of simulations to even lower masses would be important. Also, the amount of dust lost in SN-driven winds remains very poorly understood, an uncertainty that propagates in the evaluation of the continuum and Ly$\\alpha$ radiation escaping from the galaxy. Finally, Ly$\\alpha$ photons could be affected considerably by the level of dust clumping. It is unclear to what extent these effects may impact the visibility of LAEs, as discussed in e.g. Dayal, Ferrara \\& Saro (2010). Progress on these issues is expected when high-resolution FIR/sub-mm observations of LAEs with ALMA will become available in the near future (Finkelstein et al. 2009; Dayal, Hirashita \\& Ferrara 2010)." }, "1005/1005.2562_arXiv.txt": { "abstract": "{Sensitive \\herschel\\ far-infrared observations can break degeneracies that were inherent to previous studies of star formation in high-z AGN hosts. Combining PACS 100 and 160$\\mu$m observations of the GOODS-N field with 2Msec \\chandra\\ data, we detect $\\sim$20\\% of X-ray AGN individually at $>3\\sigma$. The host far-infrared luminosity of AGN with $\\lhard\\approx 10^{43}\\ergs$ increases with redshift by an order of magnitude from z=0 to z$\\sim$1. In contrast, there is little dependence of far-infrared luminosity on AGN luminosity, for $\\lhard\\lesssim 10^{44}\\ergs$ AGN at z$\\gtrsim$1. We do not find a dependence of far-infrared luminosity on X-ray obscuring column, for our sample which is dominated by $\\lhard<10^{44}\\ergs$ AGN. In conjunction with properties of local and luminous high-z AGN, we interpret these results as reflecting the interplay between two paths of AGN/host coevolution. A correlation of AGN luminosity and host star formation is traced locally over a wide range of luminosities and also extends to luminous high z AGN. This correlation reflects an evolutionary connection, likely via merging. For lower AGN luminosities, star formation is similar to that in non-active massive galaxies and shows little dependence on AGN luminosity. The level of this secular, non-merger driven star formation increasingly dominates over the correlation at increasing redshift. } ", "introduction": "Measuring the star formation rate of the host galaxy is important for studying the co-evolution of active galactic nuclei (AGN) and their hosts. It is often difficult because the AGN can outshine the host at many wavelengths. However, the rest frame far-infrared/submm emission appears dominated by the host for AGN with $\\rm\\nu L_\\nu$(60$\\mu$m)$\\approx$0.1\\ldots 0.2 $\\rm L_{Bol,AGN}$ and higher (e.g. Netzer et al. \\cite{netzer07} and introduction to Lutz et al. \\cite{lutz10}, see also Wang et al. \\cite{wang08} for luminous high-z QSOs) and has been used as a host star formation rate diagnostic of high-z AGN (e.g. Serjeant \\& Hatziminaoglou \\cite{serjeant09}, Mullaney et al. \\cite{mullaney10}, and Lutz et al. \\cite{lutz10}). \\herschel\\ can detect much lower star formation rates than previous far-infrared and submm studies. It further improves such work by measuring the rest frame far-infrared SED peak without extrapolation from longer wavelengths. We here present a first \\herschel\\ study of rest frame far-infrared emission and host star formation in X-ray selected AGN in the GOODS-N field. Throughout the paper, we adopt an $\\Omega_m =0.3$, $\\Omega_\\Lambda =0.7$ and $H_0=70$ km\\,s$^{-1}$\\,Mpc$^{-1}$ cosmology. ", "conclusions": "" }, "1005/1005.2612_arXiv.txt": { "abstract": "We compute new chemical profiles for the core and envelope of white dwarfs appropriate for pulsational studies of ZZ Ceti stars. These profiles are extracted from the complete evolution of progenitor stars, evolved through the main sequence and the thermally-pulsing asymptotic giant branch (AGB) stages, and from time-dependent element diffusion during white dwarf evolution. We discuss the importance of the initial-final mass relationship for the white dwarf carbon-oxygen composition. In particular, we find that the central oxygen abundance may be underestimated by about 15\\% if the white dwarf mass is assumed to be the hydrogen-free core mass before the first thermal pulse. We also discuss the importance for the chemical profiles expected in the outermost layers of ZZ Ceti stars of the computation of the thermally-pulsing AGB phase and of the phase in which element diffusion is relevant. We find a strong dependence of the outer layer chemical stratification on the stellar mass. In particular, in the less massive models, the double-layered structure in the helium layer built up during the thermally-pulsing AGB phase is not removed by diffusion by the time the ZZ Ceti stage is reached. Finally, we perform adiabatic pulsation calculations and discuss the implications of our new chemical profiles for the pulsational properties of ZZ Ceti stars. We find that the whole $g-$mode period spectrum and the mode-trapping properties of these pulsating white dwarfs as derived from our new chemical profiles are substantially different from those based on chemical profiles widely used in existing asteroseismological studies. Thus, we expect the asteroseismological models derived from our chemical profiles to be significantly different from those found thus far. ", "introduction": "\\label{intro} Pulsating DA (H-rich atmospheres) white dwarfs, also called ZZ Ceti or DAV stars, are the most numerous class of degenerate pulsators, with over 143 members known today (Winget \\& Kepler 2008). Since the discovery of the first ZZ Ceti star, HL Tau 76, by Landolt (1968), there has been a continuous effort to model the interior of these variable stars. ZZ Ceti stars are found within a very narrow strip of effective temperatures ($10\\,500$ K $\\lesssim T_{\\rm eff} \\lesssim 12\\,500$ K). They are characterized by multiperiodic brightness variations of up to $0.30$ mag caused by spheroidal, non-radial $g$-modes of low degree ($\\ell \\leq 2$) with periods between 100 and 1200 s. The driving mechanism thought to excite the pulsations near the blue edge of the instability strip is the $\\kappa-\\gamma$ mechanism that takes place in the hydrogen partial ionization zone (Dolez \\& Vauclair 1981; Dziembowski \\& Koester 1981; Winget et al. 1982). Also, the ``convective driving'' mechanism has been proposed --- first by Brickhill (1991) and later re-examined by Goldreich \\& Wu (1999). It appears to be the responsible of mode driving once a thick convection zone has developed at the stellar surface. The comparison of the observed pulsation periods in white dwarfs and the periods computed for appropriate theoretical models (white dwarf asteroseismology) allows to infer details of their origin, internal structure and evolution (Winget \\& Kepler 2008; Fontaine \\& Brassard 2008). In particular, the stellar mass, the thickness of the outer envelopes, the core chemical composition, magnetic fields and rotation rates can be determined from the observed periods. In addition, the asteroseismology of ZZ Ceti stars is a valuable tool for studying axions (Isern et al. 1992; C\\'orsico et al. 2001; Bischoff-Kim et al. 2008; Isern et al. 2010) and crystallization (Montgomery et al. 1999; C\\'orsico et al. 2004, 2005; Metcalfe et al. 2004; Kanaan et al. 2005). Finally, the temporal changes in the observed periods can help detect planets orbiting around white dwarfs (Mullally et al. 2008). The first published complete set of DA white dwarf models suitable for asteroseismology was that of Tassoul et al. (1990). A large parameter space was explored in such a monumental study, and for a long time (since the early eighties) this set of models represented the state-of-the-art in the area. The pulsation properties of these models were thoroughly explored in a series of important papers by Brassard et al. (1991, 1992a, 1992b). As important as these models were at that time, they suffer from a number of shortcomings. For instance, the core of the models is made of pure carbon, while stellar evolution calculations indicate that cores of typical white dwarfs are made of a mixture of carbon and oxygen. Also, the carbon/helium (C/He) and helium/hydrogen (He/H) chemical interfaces are modeled on the basis of the assumption of the diffusive equilibrium in the ``trace element approximation'', an approach that involves a quasi-discontinuity in the chemical profiles at the transition regions which, in turn, leads to peaked features in the Brunt-V\\\"ais\\\"al\\\"a frequency and exaggerated mode-trapping effects (C\\'orsico et al. 2002a, 2002b). These models were employed for asteroseismological inferences of the DAVs G 226$-$29 (Fontaine et al. 1992) and GD 154 (Pfeiffer et al. 1996). More recently, Pech et al. (2006) and Pech \\& Vauclair (2006) have presented asteroseismological analysis on HL Tau 76 and G 185$-$32, respectively, by employing similar DA white dwarf models, although with updated input physics. The models of Bradley (1996) constituted a substantial improvement in the field. These models have carbon-oxygen cores in varying proportions, and the C/He and He/H chemical interfaces are more realistic. Perhaps the most severe shortcoming of these models is the (unrealistic) ramp-like shape of the core carbon-oxygen chemical profiles. These DA models were the basis of the very important asteroseismological studies on the DAVs G 29-38 (Bradley \\& Kleinman 1997), G 117$-$B15A and R 548 (Bradley 1998), GD 165 and L 19$-$2 Bradley (2001), and G 185$-$32 (Bradley 2006). The next step in improving the modeling of DAVs was given by C\\'orsico et al. (2002b) and Benvenuto et al. (2002a), who employed evolutionary models characterized by He/H chemical interfaces resulting from a time-dependent element diffusion treatment (Althaus \\& Benvenuto 2000), and the carbon-oxygen core chemical structure extracted from the evolutionary computations of Salaris et al. (1997). The use of very smooth outer chemical interfaces, as shaped by chemical diffusion, revealed that the use of the trace element approximation turns out to be inappropriate to model the shape of the chemical interfaces in a DA white dwarf. This grid of models was employed in an asteroseismological study of G 117$-$B15A (Benvenuto et al. 2002a). For these sequences, the starting configurations for the white dwarf evolution were obtained through an artificial procedure, and not as result of evolutionary computations of the progenitor stars. Recently, Castanheira \\& Kepler (2008, 2009) have carried out an extensive asteroseismological study of DAVs by employing DA white dwarf models similar to those of Bradley (1996), but with a simplified treatment of the core chemical structure, by somewhat arbitrarily fixing the central abundances to 50 \\% oxygen and 50 \\% carbon. The He/H chemical interfaces adopted for these models are a parametrization of the realistic chemical profiles resulting from time-dependent element diffusion (Althaus et al. 2003). The study includes the ``classical'' DAVs and also the recently discovered SDSS DAVs. In total, 83 ZZ Ceti stars are analyzed. An important result of these studies is that the thickness of the H envelopes inferred from asteroseismology is in the range $10^{-4} \\gtrsim M_{\\rm H}/M_* \\gtrsim 10^{-10}$, with a mean value of $M_{\\rm H}/M_*= 5 \\times 10^{-7}$. This suggests that an important fraction of DAs characterized by envelopes substantially thinner than predicted by the standard evolution theory could exist, with the consequent important implications for the theory of white dwarf formation. However, these results are preliminary and do not include the possible effects of realistic carbon-oxygen profiles on the asteroseismological fits. Almost simultaneously with the study of Castanheira \\& Kepler (2008, 2009), Bischoff-Kim et al. (2008) performed a new asteroseismological study on G 117$-$B15A and R 548 by employing DA white dwarf models similar to those employed by Castanheira \\& Kepler (2008, 2009), but incorporating realistic core chemical profiles according to Salaris et al. (1997). The results of this work are in reasonable agreement with previous studies on these ZZ Ceti stars. However, the mass and effective temperatures found by Bischoff-Kim et al. (2008) for G 117$-$B15A are rather high (especially the mass, at $0.66 \\, M_{\\sun}$). Recently, Bogn\\'ar et al. (2009) have employed the same asteroseismological modeling to study the pulsations of the ZZ Ceti star KUV 02464+3239. Finally, Bischoff-Kim (2009) presented the results of an asteroseismological analysis of two DAVs with rich pulsation spectrum, G 38$-$29 and R 808 based on similar models, with parametrized, smooth ramp-like core profiles. These models are able to reproduce the observed period spectra reasonably well, though some assumptions about the $m$ and $\\ell$ identification of modes were made. White dwarf stellar models with realistic chemical profiles are crucial to correctly assess the adiabatic period spectrum and mode-trapping properties of the DAVs, which lies at the core of white dwarf asteroseismology (Brassard et al. 1992a; Bradley 1996; C\\'orsico et al. 2002a). In this paper, we compute for the first time consistent chemical profiles for both the core {\\sl and} the envelope of white dwarfs with various stellar masses appropriate for detailed asteroseismological fits of ZZ Ceti stars. These chemical profiles are extracted from the full and complete evolution of progenitor stars from the zero age main sequence, to the thermally-pulsing and mass-loss phases on the asymptotic giant branch (AGB), and from time-dependent element diffusion predictions during the white dwarf stage\\footnote{The chemical profiles for the core and envelope of our models appropriate for ZZ Ceti stars are available at {\\tt http://www.fcaglp.unlp.edu.ar/evolgroup}}. These profiles will be valuable in conducting future asteroseismological studies of ZZ Ceti stars that intend to include realistic chemical profiles throughout the interior of white dwarfs. To assess the impact of these new profiles on the theoretical pulsational inferences, we perfom adiabatic pulsation computations and compare the resulting periods with the pulsational inferences based on the most widely used chemical profiles in existing asteroseismological fits. The paper is organized as follows. In Sect. \\ref{physics}, we provide a description of the input physics assumed in the evolutionary calculations of relevance for the chemical composition. In Sect. \\ref{mimf} we discuss the importance of the initial-final mass relationship for the expected white dwarf carbon-oxygen composition. The resulting chemical profiles are described at some length in Sect. \\ref{profiles}. The implications of our new chemical profiles for the pulsational properties of ZZ Ceti stars are discussed in Sect. \\ref{pulsation}. We conclude in Sect. \\ref{conclusions} by summarizing our findings. ", "conclusions": "\\label{conclusions} In this paper we computed new chemical profiles for the core and envelope of white dwarfs appropriate for pulsational studies of ZZ Ceti stars. These profiles were derived from the full and complete evolution of progenitor stars from the zero age main sequence, through the thermally-pulsing and mass-loss phases on the asymptotic giant branch (AGB). These new profiles are intented for asteroseismological studies of ZZ Ceti stars that require realistic chemical profiles throughout the white dwarf interiors. In deriving the new chemical profiles, we employed the {\\tt LPCODE} evolutionary code, based on detailed and updated constitutive physics. Extra-mixing episodes during central hydrogen and helium burning, time-dependent element diffusion during the white dwarf stage and chemical rehomogenization of the inner carbon-oxygen composition by Rayleigh-Taylor instabilities were considered. The metallicity of progenitor stars is $Z=0.01$. We discussed at some length the importance of the initial-final mass relationship for the white dwarf carbon-oxygen composition. A reduction of the efficiency of extra-mixing episodes during the thermally-pulsing AGB phase, supported by different pieces of theoretical and observational evidence, yields a gradual increase of the hydrogen-free core mass as evolution proceeds during this phase. As a result, the initial-final mass relationship by the end of the thermally-pulsing AGB is markedly different from that resulting from considering the mass of the hydrogen free core right before the first thermal pulse. We found that this issue has implications for the carbon-oxygen composition expected in a white dwarf. In particular, the central oxygen abundance may be underestimated by about 15\\% if we assume the white dwarf mass to be the hydrogen-free core mass before the first thermal pulse. We also discuss the importance of the computation of the thermally-pulsing AGB and element diffusion for the chemical profiles expected in the outermost layers of ZZ Ceti stars. In this sense, we found a strong dependence of the outer layer chemical stratification on the stellar mass. In less massive models, the intershell region rich in helium and carbon --- which is built during the mixing episode at the last AGB thermal pulse --- is not removed by diffusion by the time the ZZ Ceti stage is reached. Finally, we performed adiabatic pulsation computations and discussed the implications of our new chemical profiles for the pulsational properties of ZZ Ceti stars. We found that the whole $g-$mode period spectrum and the mode-trapping properties of these pulsating white dwarfs as derived from our new chemical profiles are substantially different from those based on the most widely used chemical profiles in existing asteroseismological studies. We expect the best fit parameters of asteroseismological studies using the {\\tt LPCODE} chemical profiles to differ significantly from those found in studies made so far. Further studies will show in what way. Will we solve the high mass problem with G117$-$B15A and Salaris-like core profiles (Bischoff-Kim et al. 2008) or find thicker hydrogen layers in asteroseismological fits, more in line with stellar evolution calculations (Castanheira \\& Kepler 2008)?" }, "1005/1005.0617_arXiv.txt": { "abstract": "We investigate the dark matter and the cosmological baryon asymmetry in a simple theory where baryon ($B$) and lepton ($L$) number are local gauge symmetries that are spontaneously broken. In this model, the cold dark matter candidate is the lightest new field with baryon number and its stability is an automatic consequence of the gauge symmetry. Dark matter annihilation is either through a leptophobic gauge boson whose mass must be below a TeV or through the Higgs boson. Since the mass of the leptophobic gauge boson has to be below the TeV scale one finds that in the first scenario there is a lower bound on the elastic cross section of about $5 \\times 10^{-46} \\ \\rm{cm}^2$. Even though baryon number is gauged and not spontaneously broken until the weak scale, a cosmologically acceptable baryon excess is possible. There is tension between achieving both the measured baryon excess and the dark matter density. ", "introduction": "In the LHC era, we hope to either verify the standard model or discover the theory that describes the physics of the weak scale. One of the open issues in the standard model (SM) is the origin of the accidental global symmetries, $U(1)_{B}$ and $U(1)_{L}$, where $B$ stands for baryon number and $L$ for the total lepton number. At the non-renormalizable level in the SM one can find operators that violate baryon number and lepton number. For example, $QQQl/\\Lambda_{B}^2$ and $llHH/\\Lambda_L$, where $\\Lambda_B$ and $\\Lambda_L$ are the scales where $B$ and $L$ are respectively broken~\\cite{Weinberg:1979sa}. Since the $QQQl/\\Lambda_{B}^2$ operator gives rise to proton decay~\\cite{Nath:2006ut} the cutoff of the theory has to be very large, $\\Lambda_{B} > 10^{15}$ GeV. There is no other reason that the cutoff of the SM has to be that large and so it is worth thinking about the possibility that both $B$ and $L$ are local gauge symmetries that are spontaneously broken~\\cite{FileviezPerez:2010gw} at a much lower scale (e.g., the weak scale) and it is these gauge symmetries that prevent proton decay. Recently, two simple models (denoted model (1) and model (2)) where $B$ and $L$ are local gauge symmetries have been proposed~\\cite{FileviezPerez:2010gw}. In these models all anomalies are cancelled by adding a single new fermionic generation. One of the theories (model (1)) has an interesting realization of the seesaw mechanism~\\cite{Minkowski:1977sc, Gell-Mann1979, Mohapatra:1979ia} for neutrino masses and they both have a natural suppression of tree-level flavor changing neutral currents in the quark and leptonic sectors due to the gauge symmetries and particle content. In model (2), the neutrinos have Dirac masses. In addition, for model (2), the lightest new field with baryon number is a candidate for the cold dark matter and its stability is an automatic consequence of the gauge symmetry. It has been shown in Ref.~\\cite{FileviezPerez:2010gw} that $B$ and $L$ can be broken at the weak scale and one does not generate dangerous operators mediating proton decay. We show how a dark matter candidate can arise in model (1). In this article we investigate the properties of the cold dark matter candidates in the models proposed in Ref.~\\cite{FileviezPerez:2010gw} and study the implications of spontaneous $B$ and $L$ breaking at the weak scale for the baryon asymmetry in the Universe. In model (2), the dark matter candidate, $X$, which has baryon number $-2/3$ can either annihilate through the leptophobic $Z_B$ present in the theory or through the Higgs boson. We study the constraints from the relic density and the predictions for the elastic cross section relevant for direct detection experiments. We discuss the implications of the gauging of $B$ and $L$ for baryogenesis. There is a potential conflict between the measured baryon excess and dark matter density. For model (1), we discuss the generation of a baryon excess. We introduce a limit of the theory where $L$ is broken at a high scale but $B$ is spontaneously broken at the weak scale. In this limit standard leptogenesis plus a primordial excess in the field responsible for baryon number breaking can give rise to an acceptable baryon excess and dark matter density even though the baryon number gauge symmetry is not broken until the weak scale. This paper is organized as follows: In Section \\ref{Section2} we discuss the main features of the model. In Section \\ref{Section3} we discuss, for model (2), the properties of the dark matter candidate in the theory, constraints from the relic density and the predictions for the elastic cross section relevant for direct detection experiments. The properties of the dark matter candidate in model (1) are similar to cases already discussed in the literature (see for example \\cite{LopezHonorez:2006gr} and \\cite{Dolle:2009fn}). In Section \\ref{Section4} we discuss the implications of the breaking of $B$ and $L$ at the weak scale for baryogenesis. We summarize the main results in Section \\ref{Section5}. ", "conclusions": "\\label{Section5} We have investigated the cosmological aspects of two simple models, denoted (1) and (2), in which baryon number ($B$) and lepton number ($L$) are local gauge symmetries that are spontaneously broken around the weak scale. In these models, the stability of our scalar dark matter candidate is a consequence of the gauge symmetry. In model (2), we studied the possible dark matter annihilation channels and found what values of the masses and couplings lead to the observed relic abundance of dark matter. In the case where the s-wave annihilation through an intermediate Higgs dominates, we find that, for $M_H = 120$ GeV, in order to evade the direct detection bounds the coupling between the Higgs and the dark matter must be less than $10^{-1.5}$ and $ 51 \\ \\rm{GeV}\\ \\lesssim M_X \\lesssim 63 \\ \\rm{GeV} $. In the case where the p-wave annihilation through an intermediate leptophobic gauge boson dominates, we find that the coupling between the leptophobic $Z_B$ and the dark matter must be less than $0.1$ and $ 235 \\ \\rm{GeV}\\ \\lesssim M_X \\lesssim 250 \\ \\rm{GeV} $ when $M_{Z_B}=500$ GeV. In this case the leptophobic gauge boson has to be below the TeV scale and one finds a lower bound on the elastic cross section $\\sigma_{SI}^B \\gtrsim 5 \\times 10^{-46} \\ \\rm{cm}^2$. In both cases, direct detection experiments constrain the annihilation to proceed close to resonance in order to evade direct detection and to produce the observed relic abundance of dark matter. We have shown that even though baryon number is gauged and spontaneously broken at the weak scale it is possible to generate a cosmological baryon excess. A modest fine-tuning is needed to achieve both the measured dark matter relic abundance and baryon excess. In model (1), we introduced a simple mechanism to split the masses of the real of the imaginary part of the neutral component of the new scalar doublet to evade direct detection limits. We showed that one can simultaneously achieve both the observed baryon asymmetry of the Universe and the dark matter relic abundance. In particular, when $L$ is broken at the high scale but $B$ is spontaneously broken at the weak scale, standard leptogenesis can be applied. \\subsection*" }, "1005/1005.4728_arXiv.txt": { "abstract": "It is found that \\feii emission contributes significantly to the optical and ultraviolet spectra of most active galactic nuclei. The origin of the optical/UV \\feii emission is still a question open to debate. The variability of \\feii would give clues to this origin. Using 7.5 yr spectroscopic monitoring data of one Palomer-Green (PG) quasi-stellar object (QSO), PG 1700+518, with strong optical \\feii emission, we obtain the light curves of the continuum \\lv, \\feii, the broad component of \\hb, and the narrow component of \\hb by the spectral decomposition. Through the interpolation cross-correlation method, we calculate the time lags for light curves of \\feii, the total \\hb, the broad component of \\hb, and the narrow component of \\hb with respect to the continuum light curve. We find that the \\feii time lag in PG1700+518 is $209^{+100}_{-147}$ days, and the \\hb time lag cannot be determined. Assuming that \\feii and \\hb emission regions follow the virial relation between the time lag and the FWHM for the \\hb and \\feii emission lines, we can derive that the \\hb time lag is $148^{+72}_{-104}$ days. The \\hb time lag calculated from the empirical luminosity--size relation is 222 days, which is consistent with our measured \\feii time lag. Considering the optical \\feii contribution, PG 1700+518 shares the same characteristic on the spectral slope variability as other 15 PG QSOs in our previous work, i.e., harder spectrum during brighter phase. ", "introduction": "The variability is a common phenomenon in quasi-stellar objects (QSOs) and provides a powerful constraint on their central engines. In the past two decades, the optical variability research focused on the spectral monitoring instead of the pure photometric monitoring. With the active galactic nuclei (AGNs) watch and the Palomer-Green (PG) QSOs spectrophotometrical monitoring projects, the reverberation mapping method, i.e., exploring the correlation between the emission lines and the continuum variations, is used to investigate the inner structure in AGNs \\citep[e.g.,][]{Blandford82, Peterson93}. It is found that motions of clouds in the broad line regions (BLRs) are virialized \\citep[e.g.,][]{Kaspi00, Kaspi05, Peterson04}. With the line width of \\hb, \\mgii, \\civ from BLRs, the empirical size-luminosity relation derived from the mapping method is used to calculate the masses of their central supermassive black holes \\citep[SMBHs; e.g.,][]{Kaspi00, McLure04, Bian04, Peterson04, Greene05}. It is found that the \\feii emission contributes significantly to the optical and ultraviolet spectra of most AGNs. Thousands of UV \\feii emission lines blend together to form a pseudocontinuum, resulting in the ``small blue bump'' around 3000\\AA\\ when they are combined with Balmer continuum emission \\citep[e.g.,][]{Wills85}. The optical \\feii would lead to two bumps in two sides around the \\hb $\\lambda 4861$\\AA\\ \\citep[e.g.,][]{Boroson92}. It is found that the flux ratio of \\feii to \\hb, $R_{\\rm Fe}$, where the optical \\feii flux is the flux of the \\feii\\ emission between $\\lambda$4434 and $\\lambda$4684, strongly correlates with the so-called Eigenvector 1, which is suggested to be driven by the accretion rate \\citep[e.g.,][]{Boroson92, Marziani03a}. The origin of the optical/UV \\feii emission is still an open question. It is found that photoionized BLRs cannot produce the observed shape and strength of the optical \\feii emission and that the strength of UV \\feii cannot be explained unless considering the micro-turbulence of hundreds of \\kms or the collisional excitation in warm, dense gas \\citep{Baldwin04}. However, \\citet{Vestergaard05} found the correlation between the optical \\feii variance and the continuum variance and suggested that the optical \\feii is due to the line fluorescent in a photoionized plasma. It suggests that the optical \\feii line do not come from the same region as the UV \\feii emission \\citep[e.g.,][]{Kuehn08}. \\citet{Maoz93} found that the reverberation time lag of UV \\feii in NGC 5548 is about 10 days, similar to \\civ time lag, smaller than the \\hb time lag. The reverberation measurement for the optical \\feii emission has not fared so well. Some suggested that the optical emission is produced in the same region as the other broad emission lines, and some suggested that it is in the outer portion of the BLRs because of narrower FWHM of \\feii with respect to \\hb \\citep[e.g.,][]{Laor97, Marziani03a, Vestergaard05, Kuehn08}. Recently, \\citet{Hu2008a, Hu2008b} did a systematic analysis of \\feii emission in QSOs from the Sloan Digital Sky Survey (SDSS), and found that the \\feii emission is redshifted with respect to the rest frame defined by the \\oiii narrow emission line and \\hb intermediate-width component is correlated with \\feii which locates at the outer portion of the BLRs. \\citet{Kaspi00} gave the 7.5 yr spectroscopic monitoring data for 17 PG QSOs. There is one PG QSO, PG 1700+518, with strongest optical \\feii emission and $R_{\\rm Fe}=1.42$ \\citep{Turnshek85, Boroson92}. Its \\hb FWHM is $1846\\pm 682 \\kms$ \\citep{Peterson04}, and it is also called as a narrow line Seyfert 1 galaxy (NLS1). Using the \\feii template from one NLS1, I ZW 1, we model the \\feii emission to investigate the \\feii variability and the relation to the continuum variability in PG 1700+518. The data and analysis are described in Section 2, the results are given in Section 3, the discussion is given in Section 4, and the conclusions are presented in Section 5. All of the cosmological calculations in this paper assume $H_{\\rm 0}=70 \\kms \\rm ~Mpc^{-1}$, $\\Omega_{\\rm M}=0.3$, $\\Omega_{\\Lambda} = 0.7$. ", "conclusions": "With the spectral decomposition of 39 spectra of PG 1700+518 with the strong \\feii emission, we investigate the \\feii variability and its time lag. The main conclusions can be summarized as follows: (1) we give light curves of \\lv, \\feii, \\hbb, \\hbn, and \\hb$^{n+b}$, as well as the mean and rms spectra for PG 1700+518. With the normalized variability measure, $\\sigma_N$, we find that all components are variable. (2) With the code of Peterson et al. (2004), we find that \\feii time lag in PG1700+518 is $209^{+100}_{-147}$ days, and \\hb time lag cannot be determined. (3) Considering the uncertainties of time lags, the expected \\hb time lag from the empirical luminosity--size relation is 221.6 lt-days, consistent with our measured \\feii time lag. If we take FWHM/time lag uncertainties into consideration, \\feii emission region is located near the \\hb emission region, not conclusively located outside of the \\hb emission region. (4) Assuming that \\feii and \\hb emission regions follow the virial relation between the time lag and the FWHM for the \\hb and \\feii emission lines, we can derive that the \\hb time lag is $148^{+72}_{-104}$ days. With respect to the \\hb time lag of 252 days suggested by Peterson et al. (2004), smaller \\hb time lag, which leads to the black hole mass estimation in the logarithm, is decreased by 0.23 dex. (5) After considering \\feii contribution, PG 1700+518 shares the same characteristic on spectral slope variability to other 15 PG QSOs in our previous work \\citep{Pu06}, i.e., harder spectrum during brighter phase." }, "1005/1005.3422_arXiv.txt": { "abstract": "{We present an analysis of the first space-based far-IR-submm observations of M~33, which measure the emission from the cool dust and resolve the giant molecular cloud complexes. With roughly half-solar abundances, M33 is a first step towards young low-metallicity galaxies where the submm may be able to provide an alternative to CO mapping to measure their H$_2$ content. In this Letter, we measure the dust emission cross-section $\\sigma$ using SPIRE and recent CO and \\HI\\ observations; a variation in $\\sigma$ is present from a near-solar neighborhood cross-section to about half-solar with the maximum being south of the nucleus. Calculating the total H column density from the measured dust temperature and cross-section, and then subtracting the \\HI\\ column, yields a morphology similar to that observed in CO. The H$_2$/\\HI\\ mass ratio decreases from about unity to well below 10\\% and is about 15\\% averaged over the optical disk. The single most important observation to reduce the potentially large systematic errors is to complete the CO mapping of M~33. } ", "introduction": "Understanding star formation requires studying the interplay between the phases of the interstellar medium (ISM). Dust processes most of the energy transiting the ISM, but the cool dust component, although representing the vast majority of the dust mass, is difficult to observe from the ground. {\\it Herschel} SPIRE observations \\citep{Pilbratt10,Griffin10} are the first space-based 250-500 \\micron\\ data and as such provide a unique occasion to put together a global picture of the cool gas and dust in M~33. In particular, we compare the morphology of the Far-IR emission and that of the gas as determined from CO and \\HI\\ measurements and attempt to measure how the dust cross-section varies in M~33. A longer term goal is to be able to use the dust emission to constrain the variation of the $\\ratioo$ factor within M~33 and elsewhere. This Letter is one of a series on the {\\tt HERM33ES} project on the ISM of the Local Group galaxy M~33, an overview of which is given in \\citet{Kramer10}, hereafter K10. For consistency with the other M\\,33 papers in this volume, we adopt a distance of $D$ = 840 kpc for M\\,33 (i.e., 25 arcsec = 100 pc) and orientation parameters of $PA$ = 22.5 degrees and $i$ = 56\\degr. We use the recent {\\tt HERM33ES} SPIRE observations at 250, 350, and 500 \\micron\\, combined with CO(2--1) observations from the M33CO@IRAM project, as a tracer of the molecular component, and a high-resolution mosaic of VLA HI data \\citep[both from][]{Gratier10}. The SPIRE data were first processed as described in K10 and then converted from Jy/beam to brightness units (MJy/sr). To estimate the dust temperature, the 250\\micron\\ data were convolved to the 350\\micron\\ beamsize ($\\sim 25\"$) and, assuming a single temperature grey body with an emissivity $\\propto \\nu^\\beta$ with $\\beta =2$, a temperature was derived from the flux ratio. At the temperatures of the cool component seen in M~33, \\citet{Dupac03} find $\\beta \\approx 2$. To minimize the effect of the uncertainty in $\\beta$, we chose to use adjacent bands to estimate temperature. The 250/350\\micron\\ ratio provides more accurate temperatures than the 350/500 micron ratio even for temperatures below 10~K. A 15\\% variation (or uncertainty) in the 250/350\\micron\\ ratios corresponds to a temperature change of 0.9, 2.1, and 4.1 K at temperatures of 10, 15, and 20 K but the same uncertainty in the 350/500\\micron\\ ratios yields 1.3, 3.2, and 6.3 K errors for the same dust temperatures. A further advantage is that we obtain the dust temperature at a resolution typical of giant molecular clouds (GMC), $\\sim 100$ pc. Ideally, a multi-component fit would be used but this requires high S/N data at many wavelengths over the whole disk. By assuming that the cool dust component dominates the emission beyond 250\\micron, we could calculate the dust temperature out to the optical radius of M~33. The resulting temperature map is shown in Fig. 1. Figure 2 shows a comparison with the temperature of the cool component of the preferred two-component (warm plus cool dust) model fit to data between 24\\micron\\ and 500\\micron\\ from K10. The two-component model uses $\\beta=1.5$ and the temperatures are higher out to 6~kpc. The temperatures in the radial bins have been estimated by averaging the temperatures rather than averaging the emission as in K10. The latter yields slightly higher temperatures because the dust is usually warmer where emission is strong (compare Fig. 1 with Fig 1. in K10). K10 calculate the dust column density using $\\kappa=0.4 (\\nu/250{\\rm GHz})^2 \\ {\\rm cm}^2$g$^{-1}$ of dust, which is equivalent to a dust cross-section per H-atom of $\\sigma = 1.1 \\times 10^{-25} (\\lambda/250\\mu {\\rm m})^{-2} {\\rm cm}^2$ for a hydrogen gas-to-dust mass ratio of 140. In the following we use the SPIRE 250/350 \\micron\\ color temperature because the data cover a greater area and agree well with the cool dust temperature of the two-component fit, showing the domination of the cool dust at these wavelengths. No correction for line contamination was subtracted from the SPIRE data. At these frequencies, the CO lines contribute very little to the continuum flux, unlike at 1.3mm or 850\\micron\\ \\citep[e.g.][]{Braine_n3079}. Figure 3 shows the SPIRE 250\\micron\\ emission with the CO(2--1) emission as contours at 25$''$ resolution to show how closely the CO emission follows the dust emission peaks. \\citet{Gratier10} show a detailed comparison of the CO and \\HI\\ emission with star formation tracers such as H$\\alpha$ and {\\it Spitzer} 8 and 24\\micron\\ maps. The 250\\micron -bright regions are detected in CO and the general morphology of the cool dust emission is seen in the \\HI\\ image. \\begin{figure}[h] \\centering \\includegraphics[width=6.2cm,angle=-90]{tdust_rad4b.eps} \\\\ \\caption{Radial distribution of the average dust temperature in the stellar disk of M\\,33. The solid line joining the triangles shows the 250/350 \\micron\\ color temperature and for comparison we show the cool dust temperature (large dots) obtained by K10 in their preferred two-component model with $\\beta=1.5$. The error bars on the triangles indicate a 15\\% calibration uncertainty in the 250/350 \\micron\\ ratio converted to temperature.} \\end{figure} \\begin{figure}[h] \\centering \\includegraphics[width=8.8cm,angle=0]{s250_co_figb.eps} \\\\ \\caption{SPIRE 250\\micron\\ emission with the IRAM CO(2--1) emission at 25$''$ resolution superposed as contours and with the polygons to the upper left used to calculate the dust cross-section per H-atom. The beams are shown as white dots in the lower left corner. The CO contours are at 0.5 (red), 2 (blue), and 4 K km s$^{-1}$ in the main beam scale. The large polygon indicates the area covered so far by the IRAM CO survey.} \\end{figure} ", "conclusions": "" }, "1005/1005.3278_arXiv.txt": { "abstract": "We report CCD photometry of the cataclysmic variable V1113 Cygni. During two campaigns, lasting from May to August 2003 and from March to June 2005, we recorded two superoutburst. In the obtained light curves we detected clear superhumps with a mean period $P_{\\rm sh}=0.07891(3)$ days ($113.63\\pm0.04$ min). That fact confirms that the star is a member of SU UMa class of dwarf novae. During the first observed superoutburst the superhump period was decreasing with an enormous rate of $\\dot P = -4.5(8)\\times 10^{-4}$ which is one of the highest values ever observed in SU UMa systems. \\noindent {\\bf Key words:} \\textit{Stars: individual: V1113 Cyg - binaries: close - novae, cataclysmic variables} ", "introduction": "Cataclysmic variable stars are close binary systems containing white dwarf (the primary) and red dwarf (the secondary). In those systems the secondary loses its mass through the inner Lagrangian point in the Roche-lobe and the primary star accretes it (Warner 1995). SU UMa class of dwarf novae is one of the subclasses of cataclysmic variable stars. SU UMa-type stars show two types of outbursts: short (outbursts) and long (superoutbursts). Superoutbursts are about one magnitude brighter than ordinary outbursts and have a duration a few days longer than the outbursts. In the light curves of superoutbursts one can see periodic light oscillations (superhumps) with a period a few percent longer than the orbital period of the binary system. The thermal-tidal instability model proposed by Osaki (1996) is now known as the best explanation of the mechanism of the superoutbursts and superhumps (for a review see Warner 1995, Osaki 1996, Hellier 2001). However, in recent years some serious criticism concerning both the nature of superoutburst and nature of superhumps has appeared (Schreiber and Lasota 2007, Smak 2004, 2008, 2009). V1113 Cyg was discovered as a dwarf nova by Hoffmeister (1966), but received little attention for almost 30 years. Results of Kato et al. (1995) have shown that the star is SU UMa-type with superhumps' period of $0.0792(\\pm0.0001)$ d. They also reported that V1113 Cyg has a large ($\\sim$ 6 mag) superoutburst amplitude. In the second paper by Kato et al. (2001) the detection of 30 outburst from July 1994 to May 2001 is presented. They noticed low number of outbursts (only about two normal outbursts per each supercycle) and high number of superoutbursts (12 from 30 detected). They suspected variation of the mass transfer rate. Because results are inconclusive further observations were needed. Kato et al. (2009) reanalyzed the observations of Kato et al. (1995) and obtained new observations during the 2008 superoutburst. They reported decreasing superhump periods in both superoutburts with the rates of $-19.2(6.8)\\times 10^{-5}$ and $-5.2(4.7)\\times 10^{-5}$, respectively. ", "conclusions": "Recently, Kato et al. (2009) published a comprehensive survey of period variations of superhumps in SU UMa-type dwarf novae. The conlcuded that in many systems evolution of the superhump period can be divided into three distinct parts: an early stage with a stable and longer period (stage A), middle stage for which many systems (especially with $P_{\\rm sh}<0.08$ days) show positive superhump period derivative (stage B) and final stage with a shorter but again stable superhump period (stage C). However they reported that V1113 Cyg does not follow this scenario with decreasing superhump periods with the rates of $-19.2(6.8)\\times 10^{-5}$ and $-5.2(4.7)\\times 10^{-5}$ observed for 1994 and 2008 superoutbursts, respectively. V1113 Cyg has its superhump period close to 0.08 day borderline and thus can show different superhump period evolution. Our high negative $\\dot P$ value may indicate that the behaviour of the star is similar to that observed in SU UMa or DM Dra (see Figs. 4 and 7 in Kato et al. 2009). On the other hand, it is also possible that in 2003 our observations caught V1113 Cyg during the transition from stage A to B. As can be estimated from Fig. 4 of Kato et al. (2009), for stars with superhump period of around 0.078 days, such a transition occurs around cycle number $E\\approx 20$ assuming $E=0$ for the begining of the superoutburst (see $O-C$ diagrams of TT Boo, QY Per, RZ Leo or EG Aqr). The peak of the parabola shown in our Fig. 5 occurs also at $E\\approx 20$ but in this case the cycle numeration starts with the first night of the observations which is not the first night of the superoutburst. Assuming that the superoutburst started 1-2 days before our first observation the transition moment shifts to $E\\approx 32-33$ or even $E\\approx 45$. It is quite late in comparison to other stars putting some doubt in A to B transition hypothesis. To summarize the results of the observations of the 2003 and 2005 brightenings of V1113 Cyg we can confirm: \\begin{itemize} \\item Detection of the clear superhumps during those events directly proves that V1113 Cyg belongs to the group of SU UMa variables. \\item The amplitude of the superoutburst is at least ($\\sim$ 4) mag, but we never caught the star at the beginning of the superoutburst. According to Kato (2001) this variable has a large ($\\sim$ 6) mag outburst amplitude and it is a possible estimation. \\item During two years of observation the star went into at least two superoutbursts so V1113 Cyg has about one superoutburst a year. Eruptions of the star lasted about two weeks and were not shorter than 12 days (July 2005 superoutburst). \\item We did not record any normal outbursts. Lack of them may be an observational effect. But in general we agree with the results of Kato (2001) which suggest low ratio of outbursts to superoutbursts. \\item During the first observed superoutburst the superhump period was decreasing with an enormous rate of $\\dot P = -4.5(8)\\times 10^{-4}$ which is one of the highest values ever observed in SU UMa systems. \\end{itemize} \\bigskip \\noindent {\\bf Acknowledgments.} ~We acknowledge generous allocation of the Warsaw Observatory 0.6-m telescope time. Data from AAVSO observers are also appreciated. We thank to W. Pych for providing some useful software which was used in the analysis and to K. Mularczyk and P. K\\c{e}dzierski for their assistance in observations. This work was supported by the Polish MNiSW grant no. N203~301~335. KZ was supported by Foundation for the Polish Science through grant MISTRZ. {\\small" }, "1005/1005.3570_arXiv.txt": { "abstract": "On 1998 November 14, Saturn and its rings occulted the star GSC 0622-00345. The occultation latitude was 55{\\decdegree}5 S. This paper analyzes the \\mbox{2.3 {\\micron}} light curve derived by \\citeauthor{HarringtonFrench2010apjsatoc98I}. A fixed-baseline isothermal fit to the light curve has a temperature of \\mbox{140 {\\pm} 3 K}, assuming a mean molecular mass of 2.35 AMU. The thermal profile obtained by numerical inversion is valid between 1 and 60 {\\micro}bar. The vertical temperature gradient is \\math{>}0.2 K km\\sp{-1} more stable than the adiabatic lapse rate, but it still shows the alternating-rounded-spiked features seen in many temperature gradient profiles from other atmospheric occultations and usually attributed to breaking gravity (buoyancy) waves. We conduct a wavelet analysis of the thermal profile, and show that, even with our low level of noise, scintillation due to turbulence in Earth's atmosphere can produce large temperature swings in light-curve inversions. Spurious periodic features in the ``reliable'' region of a wavelet amplitude spectrum can exceed 0.3 K in our data. We also show that gravity-wave model fits to noisy isothermal light curves can lead to convincing wave ``detections''. We provide new significance tests for localized wavelet amplitudes, wave model fits, and global power spectra of inverted occultation light curves by assessing the effects of pre- and post-occultation noise on these parameters. Based on these tests, we detect several significant ridges and isolated peaks in wavelet amplitude, to which we fit a gravity wave model. We also strongly detect the global power spectrum of thermal fluctuations in Saturn's atmosphere, which resembles the ``universal'' (modified Desaubies) curve associated with saturated spectra of propagating gravity waves on Earth and Jupiter. \\if\\submitms y \\else \\comment{\\hfill\\herenote{DRAFT of {\\today} \\now}.} \\fi ", "introduction": "\\label{intro} Earth-based occultations remain an attractive method for measuring the thermal profile in the 1--100-{\\micro}bar region of a planetary atmosphere. Many such profiles for Saturn were recorded during the 28 Sgr occultation of 1989 July 3, which sampled the equatorial region from 6{\\decdegree}6 N--15{\\decdegree}2 S latitude \\citep{HubbardEtal1997icsatmes}. There is a single profile for the north polar region \\citep[82{\\decdegree}5--85{\\degrees} N]{CoorayEtal1998icnpsatoc}, and a northern low-latitude profile from the same event \\citep[19{\\decdegree} 16 N]{FrenchEtal1999phemu97satoc}. Saturn's central flash probes much deeper, around 2.5 mbar; \\citet{NicholsonEtal1995icsatcfl} obtained IR images of the flash during the 28 Sgr event, from which they inferred the zonal wind profile of the sampled latitudes along Saturn's limb. Occultations observed by a spacecraft near a giant planet probe the troposphere from the cloud deck (\\sim1 bar) to the mbar level at radio and infrared wavelengths. They probe the upper stratosphere and thermosphere (\\math{<}1 {\\micro}bar) in the ultraviolet. Earth-based visual and infrared occultations measure the thermal structure of the intervening mesosphere and stratosphere regions, which are not well sampled by spacecraft experiments. For Saturn, the {\\em Pioneer} radio \\citep{KlioreEtal1980scipioneersatoc, LindalEtal1985ajvoysatoc} and {\\em Voyager 2} extreme ultraviolet solar and stellar observations \\citep{SmithEtal1983jgrvoysatoc} sensed the equatorial region only. The {\\em Voyager 1} radio occultation sensed 75{\\degrees} S \\citep{TylerEtal1981scisatvoy1radoc}, while the {\\em Voyager 2} radio occultations sensed 36{\\decdegree}5 N and 31{\\degrees} S \\citep{TylerEtal1982scivoy2radoc}. The {\\em Cassini} radio experiment has performed a number of radio occultations in the equatorial region as well as at middle and high latitudes of Saturn \\citep{NagyEtal2006jgrCassinisatradio, KlioreEtal2009jgrSatOc}. The {\\em Cassini} Ultraviolet Imaging Spectrograph stellar occultation probed the upper atmosphere at 40{\\degrees} S and 66{\\degrees} N \\citep{Shemansky2008cosparSatcassiniuvis}, and the {\\em Cassini} Composite Infrared Spectrometer mapped Saturn's thermal atmospheric emission, resulting in temperature maps for both hemispheres for pressures ranging from 0.1 mbar to about 700 mbar \\citep{FlasarEtal2005sciSatcassini}. The {\\em Cassini} Visual and Infrared Mapping Spectrometer also has the capability of observing spectrally resolved near-infrared stellar occultations by Saturn's atmosphere \\citep{BrownEtal2004ssrVIMS}. Temperature profiles for a variety of atmospheres from both occultations and in situ observations show quasi-periodic structures that are usually attributed to propagating waves. Waves have been reported on Venus \\citep{HinsonJenkins1995icvenusradoc}, Earth (\\citealp{FrittsAlexander2003revgeogwdyn}, and references therein), Mars \\citep{CreaseyEtal2006grlMarsgw, FrittsEtal2006jgrMarsgwaero}, Jupiter \\citep{FrenchGierasch1974jasocwav, YoungEtal1997scijupgravwav, RaynaudEtal2003icjupoc, RaynaudEtal2004icjupbsco, YoungEtal2005icjupasi}, Saturn \\citep{CoorayEtal1998icnpsatoc, FouchetEtal2008natSaturn}, Titan \\citep{SicardyEtal1999ictit28sgr}, Uranus \\citep{YoungEtal2001icuoc}, Neptune \\citep{RoquesEtal1999aanepstrat3}, and Pluto \\citep{PersonEtal2008ajWavPluto, HubbardEtal2009icPlutowav, ToigoEtal2010icPlutoTides}. Both the behavior of individual waves and the form of wave power spectra can reveal properties of the underlying atmosphere. For example, the forcing, propagation, and dissipation of the waves both contribute to and depend on the sources and sinks of energy in the atmosphere, the background thermal state, and eddy and molecular diffusion. On 1998 November 14, Saturn and its rings occulted GSC 0622-00345, as predicted by \\citet{BoshMcDonald1992ajsatocs}. We obtained a light curve for atmospheric immersion, based on infrared imaging observations at the NASA Infrared Telescope Facility (IRTF) on Mauna Kea, HI. The high signal-to-noise ratio (S/N) allowed us to determine the vertical temperature profile of Saturn's stratosphere at 55{\\decdegree}5 S latitude, a region not sampled by previous stellar occultation observations (see Figure 1 and Table 1 of \\citealp{HarringtonFrench2010apjsatoc98I}, hereafter Paper I). Paper I presents the light curve and describes the new methods used to acquire and derive it. This paper presents the scientific analysis of the light curve. Subsequent sections cover isothermal model fits, numerical inversions to derive the thermal profile, noise tests, local and global wavelet spectrum analysis, a gravity-wave model based on wavelet reconstruction, exploration of the ``universal'' power spectrum of gravity waves, discussion of the global power spectrum, and our conclusions. For each analysis, we present new significance tests that determine the effects of real (non-Gaussian) noise. ", "conclusions": "We have analyzed a light curve based on IRTF observations of the 1998 November 14 occultation of GSC 0622-00345 by Saturn (Paper I). We presented and analyzed isothermal light-curve fits, an atmospheric thermal profile, a wavelet analysis, gravity wave modeling, and power spectra. The derived thermal profile varies over 142 -- 151 K in the pressure range 1 -- 60 {\\micro}bar at a latitude of 55{\\decdegree}5 S. The vertical temperature gradient is removed by more than 0.2 K km\\sp{-1} from the adiabatic lapse rate, indicating that the stratospheric region sounded by the occultation is statically stable. Our thermal gradient profile shows the same alternating-rounded-spiked appearance of other occultation profiles, including one for Saturn \\citep{CoorayEtal1998icnpsatoc}. This shape has previously been interpreted as evidence of gravity wave breaking \\citep{RaynaudEtal2003icjupoc, RaynaudEtal2004icjupbsco, YoungEtal2005icjupasi}. Our new noise tests, based on real noise sampled from our light curve's upper baseline rather than synthetic, uncorrelated Gaussian noise, showed that atmospheric scintillation (and similar correlated noise sources such as spacecraft pointing drifts) can introduce relatively strong, spurious temperature fluctuations into the thermal profile derived by inverting an atmospheric occultation light curve. For our data, the effect was many K for the raw inversions, but it mainly appeared at longer wavelengths in power spectra. However, the amplitudes at the shorter wavelengths found outside the COI were still of order 0.1 -- 1 K, comparable to the amplitudes of the gravity waves often identified in occultation data sets. We thus developed several significance tests for our power-spectrum analyses, and note that without such tests, one must be skeptical of gravity-wave detection claims in ground-based (and possibly some space-based) occultation inversions. We used a wavelet analysis to search for localized gravity wave trains. Based on the wavelet power and the use of significance tests only, the strongest candidate had an amplitude of 0.7 K, \\math{\\lambda\\sb{z}} = 12.5 km at \\math{z\\sb{\\rm max} = -35 km}, and lasted four cycles. It and several shorter-wavelength features stand over the global power spectrum at the 95% Alternative explanations for the observed periodic structures include sound waves, planetary waves, and non-transient features. Without knowing the horizontal structure of the wave we cannot rule out sound waves or planetary waves as the cause for the temperature fluctuations. However, one can make the argument that, for a given \\math{\\lambda\\sb{z}}, planetary waves typically have lower (time) frequencies and therefore dissipate lower in the atmosphere than the valid region of our profile. To take into account both phase and amplitude information, which could improve sensitivity over our amplitude-based noise limit, we fit a gravity-wave model to the amplitude and phase of the strongest features in the valid region of the wavelet spectrum, following \\citet{RaynaudEtal2004icjupbsco}. The model calculates the temperature amplitude vs.\\ height of a single, damped wave mode propagating throughout our profile's valid region. We performed fits both with and without a new parameter for constant vertical wind shear. The added parameter did not improve the fits, so we report the shearless fits here for consistency with prior results from this model. Our best candidate for a gravity wave that propagates continuously through the valid region, as assessed by our model, had \\math{\\lambda\\sb{z} = 40} km, \\math{\\lambda\\sb{h} = 200} km, and a period of 42 minutes. At this altitude the wave is not strongly affected by dissipation and achieves a maximum amplitude above the observed atmospheric region exceeding 1 K. According to Equation (\\ref{eq:detectability}), this wave was detectable if the angle between the line of sight and the horizontal direction of wave propagation exceeded 74{\\degrees}. However, similar fits to quasi-periodic features in thermal profiles derived from isothermal (i.e., waveless) light curves with real noise gave some fits of similar appearance. This demonstrates the need for a study of real noise to establish an amplitude criterion that discriminates real waves from noise. The reason wavelike (i.e., sinusoidal) features arise out of the noise is simple: by reconstructing only a limited range of wavelengths, the resulting profile is certainly sinusoidal and has a favored period. Wave models are sinusoidal, and ours even has parameters that allow the phase and amplitude to vary, so we will get a good fit if the amplitude of the reconstructed data does not vary much. We find that this circumstance occurs in the noise data sets often enough to require at least criteria for significant wave amplitude. Criteria involving the number of cycles outside the COI or the phase of the reconstructed data could potentially provide even-more-stringent limits. We derived global power spectra from our wavelet transforms, using only data outside the COI. This method greatly reduced the noise level of the spectra, which stand everywhere 2 -- 10 times above the noise level calculated using real-noise-contaminated isothermal light curves. The power spectra follow the modified-Desaubies form of the universal spectrum of gravity waves, though with a slightly more negative high-wavenumber exponent. Superposed on this spectrum one sees the signature of the discrete wave structures discussed above. That we see both the universal spectrum and individual features fit well by gravity-wave models lends confidence that we are indeed looking at a signal dominated by gravity waves. The amplitudes of all five wave-like features that we analyzed are well above the noise level. The wave model used to fit these features is based on the assumption that the waves propagate independently of each other. Interactions between the wave modes might be able to explain the discrepancies between the observed wave fluctuations and the model's single gravity waves. Note that the discrepancies are more significant when derived wavenumbers are larger than the observationally derived characteristic wavenumber. This is the saturated part of the spectrum where wave-wave interactions determine wave behavior and vertical propagation. Direct comparison of our derived temperatures with previous Saturn occultation measurements (see, e.g., Table VII of \\citealp{HubbardEtal1997icsatmes}) requires care, both because of differences in the assumed mean molecular mass and because Saturn's mean stratospheric temperature is strongly affected by seasonally varying insolation \\citep{BezardGautier1985icsatclimate}. Heating by inertia-gravity waves might also be important, at least in some regions and/or seasons \\citep{CoorayEtal1998icnpsatoc, FrenchGierasch1974jasocwav, YoungEtal1997scijupgravwav}. Detailed modeling of stratospheric temperatures requires taking account of non-LTE effects as well \\citep{Appleby1990icch4gp}. A clearer picture of zonal and seasonal variations in Saturn's stratospheric structure should emerge when Earth-based stellar occultations can be viewed in the context of data from the {\\em Cassini} orbiter." }, "1005/1005.4485_arXiv.txt": { "abstract": "Resistive relativistic magnetohydrodynamic (RRMHD) simulations are applied to investigate the system evolution of relativistic magnetic reconnection. A time-split Harten--Lan--van Leer method is employed. Under a localized resistivity, the system exhibits a fast reconnection jet with an Alfv\\'{e}nic Lorentz factor inside a narrow Petschek-type exhaust. Various shock structures are resolved in and around the plasmoid such as the post-plasmoid vertical shocks and the ``diamond-chain'' structure due to multiple shock reflections. Under a uniform resistivity, Sweet--Parker-type reconnection slowly evolves. Under a current-dependent resistivity, plasmoids are repeatedly formed in an elongated current sheet. It is concluded that the resistivity model is of critical importance for RRMHD modeling of relativistic magnetic reconnection. ", "introduction": "Magnetic reconnection \\citep{sweet,parker,petschek} is the driver of explosive events in space, astrophysical, and laboratory plasmas. The reconnection process attracts growing attentions to explain flaring events \\citep{lyut06,gia09} and the magnetic annihilation \\citep{coro90,lyu01} in relativistic plasma environments. Basic properties of relativistic magnetic reconnection have been discussed by relativistic magnetohydrodynamic (RMHD) theories \\citep{bf94b,lyu05,ten10}. In particular, \\citet{lyu05}'s careful work brought significant insights. In the last decade, modern simulation works have revealed many features of relativistic magnetic reconnection. From the viewpoint of kinetic physics, it is widely recognized that reconnection is an efficient particle accelerator \\citep{zeni01,zeni07,claus04}. It was further found that the system is crucially influenced by the guide-field magnetic topology \\citep{zeni05} and the radiative cooling effects \\citep{claus09}. In a fluid scale, the reconnection system has been explored by a resistive RMHD (RRMHD) \\citep{naoyuki06} and relativistic two-fluid \\citep{zeni09a,zeni09b} models. \\citet{zeni09a} demonstrated and extensively analyzed a quasi-steady Petschek-type reconnection. There is a strong demand for further development of RRMHD reconnection work. Unlike the kinetic and two-fluid models, the RRMHD model is free from kinetic scales such as the skin depth and the gyro radius, and so it is highly desirable to study stellar-scale problems. In addition, plasmas are considered to be collisional in optically thick, radiation-dominated environments \\citep{uz06}. Since such plasmas can be approximated by a single RMHD fluid, the RRMHD code with non-FLD\\footnote{FLD is flux-limited diffusion} radiative transfer will be necessary for future modeling of radiative relativistic reconnection. However, no other RRMHD work has come out since \\citet{naoyuki06}, because the RRMHD equations turned out to be numerically unstable. Attempts to better solve the RRMHD system have been undertaken. \\citet{kom07} pointed out that the non-ideal electric field behaves as stiff relaxation to the ideal RMHD condition. He split the equations into two parts: the stiff terms were analytically solved as an exponential decay, and the rest part was solved by a two-step Harten--Lan--van Leer (HLL) method. \\citet{pal09} employed a hybrid scheme of an implicit Runge-Kutta method for the electric field and an explicit one for the other variables. \\citet{dumbser09} developed a higher-order scheme in unstructured volumes. In this Letter, we present RRMHD simulations of relativistic magnetic reconnection, based on recent advances in numerical schemes \\citep{kom07}. ", "conclusions": "Our results corroborate the first RRMHD work by \\citet{naoyuki06} on many aspects such as Alfv\\'{e}nic outflow, the compression ratio, and the narrower slow shock structure. Judging from their inflow speed ($v_{in}/c_{A,in}\\sim \\mathcal{R}$), they observed faster reconnection probably due to the stronger localized resistivity of $S_1 \\sim \\mathcal{O}(1)$. Our results also agree with the two-fluid work by \\citet{zeni09a} such as the typical outflow parameters of the reference runs and the Petschek angle relations. Some visible differences are attributed to the two-fluid effects. The current layers look much sharper in this work, because the fluid inertial effects tend to smooth structures in the two-fluid model. The plasmoid looks longer in the $x$-direction. The outflow jet immediately hits the downstream plasmas in the $x$-direction in the RRMHD model, while plasma out-of-plane motion softens the reflection in the two-fluid model. Concluding this and those works, (1) Alfv\\'{e}nic outflow, (2) increasing reconnection rate, and (3) narrower opening angle, can be regarded as common features of relativistic Petschek reconnection. Since a narrower outflow exhaust suppresses the energy throughput, (3) does not immediately fit (2). However, \\citet{zeni09b} pointed out that the relativistic enthalpy flux transports huge energy per a cross section even in a narrower exhaust. Thanks to the stable RRMHD code, we successfully resolved shock structures in and around the plasmoid. To our knowledge, the post-plasmoid vertical shocks and the diamond-chain structure are new discoveries. The diamond chain develops when the plasmoid edge speed (a sizable fraction of $c_{A,in}$) exceeds the sound speed in the neutral Harris sheet. We predict that it also develops in nonrelativistic MHD simulations once the above condition is met. Since multiple shocks are confined in a narrow region, the plasmoid edge can be a potential site of shock acceleration of particles. From the numerical viewpoint, relativistic reconnection in the high-$\\sigma_{\\varepsilon}$ regimes is challenging for an explicit-type RRMHD code. Since the analytic part of the scheme contains the time constant of $-S\\gamma$ (e.g., Equation \\ref{eq:anom2}a), employing Alfv\\'{e}nic outflow speed, a code requires $ \\Delta t \\lesssim S_0^{-1} (1+\\sigma_{\\varepsilon,in})^{-1/2}$. As shown in Section 3, the outflow is faster near the plasmoid neck, and so the restriction becomes even severer. Thus, high time (and appropriate spatial) resolutions are necessary in high-$\\sigma_{\\varepsilon}$ and moderate-$S$ regimes of our interest. In addition, a plasma solution is not always stable even when the total momentum and energy are well handled. Let us extract the plasma momentum from Equation \\ref{eq:mom}, $\\partial_{t} \\vec{m} + \\div ( p\\vec{I} + w \\vec{u}\\vec{u} ) = \\vec{j} \\times \\vec{B}$. Focusing on the source term which is usually problematic, we see that the plasma stability condition, ${|\\Delta \\vec{m}|}/{w} \\sim {| \\vec{j} \\times \\vec{B} | \\Delta t }/{w} \\lesssim \\mathcal{O}(0.1),$ will be critical in higher-$\\sigma_{\\varepsilon}$ regimes, unless Ohm's law strongly limits the current. These issues need to be improved by an implicit approach \\citep{pal09}. From the physics viewpoint, those restrictions obviously come from an immature form of Ohm's law. Note that the RRMHD model contains the displacement current $\\partial_t\\vec{E}$ in Amp\\`{e}re's law (Equation \\ref{eq:E}). Since it operates in a short timescale of the plasma frequency, the nonrelativistic MHD code usually drops it. However, the RRMHD system needs it for the energy and momentum conservation. On the other hand, the RRMHD uses a simple Ohm's law with the scalar resistivity (Equation \\ref{eq:ohm}). This is a time-stationary form \\citep{bf93} and there is no general consensus on extended forms. Combining the two equations with different timescales, the RRMHD system inevitably becomes problematic, especially in its impulsive phase. Thus, a further development of Ohm's law (e.g., \\citet{ged96}) or the reconstruction of the entire RRMHD system (e.g., \\citet{koide09}) is highly desirable. In order to study large-scale problems which contain reconnection in arbitrary locations, we can no longer use the localized resistivity at a fixed location. We do need good physical or phenomenological anomalous resistivity models. Our attempt is a first step toward this direction, and the current-dependent resistivity model did exhibit different system evolution. This clearly tell us that the modeling of the effective resistivity, a long-standing problem of the entire reconnection research, is very important for RRMHD modeling of relativistic magnetic reconnection as well. In summary, we explored an RRMHD model of relativistic magnetic reconnection. We corroborated earlier works and further studied fine plasmoid structures. Three different resistivity models are examined and they showed different system evolutions. It is crucially important to model an effective resistivity for RRMHD modeling of magnetic reconnection." }, "1005/1005.4399_arXiv.txt": { "abstract": "% We study analytically the development of gravitational instability in an expanding shell having finite thickness. We consider three models for the radial density profile of the shell: (i) an analytic uniform-density model, (ii) a semi-analytic model obtained by numerical solution of the hydrostatic equilibrium equation, and (iii) a 3D hydrodynamic simulation. We show that all three profiles are in close agreement, and this allows us to use the first model to describe fragments in the radial direction of the shell. We then use non-linear equations describing the time-evolution of a uniform oblate spheroid to derive the growth rates of shell fragments having different sizes. This yields a dispersion relation which depends on the shell thickness, and hence on the pressure confining the shell. We compare this dispersion relation with the dispersion relation obtained using the standard thin-shell analysis, and show that, if the confining pressure is low, only large fragments are unstable. On the other hand, if the confining pressure is high, fragments smaller than predicted by the thin-shell analysis become unstable. Finally, we compare the new dispersion relation with the results of 3D hydrodynamic simulations, and show that the two are in good agreement. ", "introduction": "% Shells and bubbles are common morphological features in the interstellar media of galaxies, and have been observed at many different wavelengths. For instance, \\citet{2002ApJ...578..176M} and \\citet{2005A&A...437..101E} have identified $\\sim 1000$ expanding shells in HI surveys of the Milky Way. HI shells have also been found in the LMC \\citep{1998ApJ...503..674K}, the SMC \\citep{2005MNRAS.360.1171H}, and other nearby galaxies. In the infrared, \\citet{2006ApJ...649..759C,2007ApJ...670..428C} have assembled a catalogue of $\\sim 600$ shells found in the GLIMPSE survey of our Galaxy, using the SPITZER space telescope. Shells have been detected in the Wisconsin H$\\alpha$ mapper data \\citep[WHAM;][]{2003ApJS..149..405H}. Shells have also been studied in millimetre molecular line emission, radio continuum and X-rays. It is generally believed that a substantial fraction of these shells is formed by feedback from massive stars. In addition, it has been suggested by \\citet{1977ApJ...214..725E} that expanding shells can trigger star formation by the {\\it collect-and-collapse} mechanism. In this mechanism, a massive star (or a group of stars) injects energy into the interstellar medium in the form of stellar winds and ionising radiation, and creates a bubble of gas with high temperature. The bubble expands due to its high internal pressure, and sweeps up the ambient medium into a dense shell. The shell cools down, becomes gravitationally unstable, fragments and forms a new generation of stars. If the new generation includes massive stars, the whole process may repeat itself, leading to sequential self-propagating star formation. Various modifications to the collect-and-collapse mechanism are reviewed by \\citet{1998ASPC..148..150E}. The collect-and-collapse mechanism has been tested observationally in a series of papers by \\citet{2003A&A...408L..25D, 2005A&A...433..565D, 2006A&A...458..191D, 2008A&A...482..585D, 2009A&A...496..177D} and \\citet{2006A&A...446..171Z}. Using data at various wavelengths, they identify and study several objects in which star formation appears to have been triggered at the borders of HII regions. \\citet{2008ApJ...681.1341W,2009ApJ...694..546W} analyse six shells from the GLIMPSE survey, and determine the properties of young stellar objects concentrated in these shells. The gravitational instability of an expanding infinitesimally thin shell has been studied theoretically by several authors. \\citet{1983ApJ...274..152V} derives a dispersion relation (perturbation growth rate as a function of wavenumber) for an infinitesimally thin shell, using decomposition into spherical harmonics. A very similar dispersion relation is obtained by \\citet{1994ApJ...427..384E} using perturbation analysis of linearised two-dimensional hydrodynamic equations. \\citet{1994MNRAS.268..291W} use the equation of motion for a two-dimensional fragment forming on the surface of a shell, and find expressions for the size of the first and most unstable fragment, and for the time at which the instability starts. All these analyses use a very similar physical model (an infinitesimally thin shell which expands into a uniform medium), and although the mathematical description in each of them is very different, the results regarding fragment sizes and growth rates are very similar. We have tested these analytic predictions for the growth rate of gravitational instability, based on the thin-shell approximation, in \\citet[][hereafter Paper~I]{2009MNRAS.398.1537D}. We use two very different three-dimensional hydrodynamic codes (an Eulerian AMR code and a Lagrangian SPH code) to simulate the evolution and subsequent fragmentation of expanding self-gravitating shells. Our setup differs slightly from the one studied by Vishniac, Elmegreen and Whitworth, in that our shell does not accrete mass from the ambient medium. This modification requires only minor changes to the linear theory of the thin shell gravitational instability, which we describe in \\S\\ref{sec:thin_shell}. We make this modification because an accreting shell is prone to the Vishniac dynamical instability \\citep{1983ApJ...274..152V}, which would grow quickly and make an analysis of gravitational instability impossible. On the other hand, we want to have the option to keep the shell thin, in order to test the thin-shell approximation. Therefore, we fill the shell interior and exterior with a hot rarefied gas and the pressure of this gas confines the shell. However, because of its low density, the ram pressure of the accreted gas, which is crucial for the development of the Vishniac instability, is negligible. We also keep the pressure of the rarefied gas constant throughout the evolution, so that the shell is effectively in free fall, and development of the Rayleigh-Taylor instability is suppressed. In Paper~I, we find excellent agreement between the two numerical codes. However, for simulations with low confining pressure, in which the shell becomes thick, the simulations do not agree well with the predictions of the thin-shell approximation. The agreement is better for simulations in which the confining pressure is chosen so that the shell thickness stays approximately constant during its evolution. In this paper (Paper~II) we study the gravitational instability of a thick shell analytically, in order to understand how a thick shell fragments, and why the fragmentation differs from the predictions of the thin-shell approximation. We show that the confining pressure, which defines the shell thickness, is an important factor in determining the range of unstable wavelengths. Therefore, we call the new model {\\em pressure assisted gravitational instability} (PAGI). The outline of the paper is as follows. In \\S\\ref{sec:radprof} we derive a semi-analytic description of the equilibrium radial profile of the shell, and compare it with our 3D numerical simulations, and with the simple uniform-density model of the shell that we adopt in the subsequent analysis. Section \\S\\ref{sec:gi} deals in detail with modelling gravitational instability in a shell. We briefly derive the dispersion relation for a non-acreting shell, using the thin-shell approximation; we give a description of a fragment forming in a shell with non-zero thickness, and compare it with the description used in the thin-shell approximation; finally we derive a new dispersion relation for a thick pressure-confined shell. In \\S\\ref{sec:num}, we compare the new dispersion relation with growth rates measured in our 3D numerical simulations, in \\S\\ref{discussion} we discuss the astrophysical consequences of our results, and in \\S\\ref{conclusions} we summarise our conclusions. ", "conclusions": "% We have studied an isothermal ballistic shell, confined from both sides by a hot highly rarefied gas having non-zero pressure, $P_{_{\\rm EXT}}$. We have solved the equation of hydrostatic equilibrium in a frame of reference moving with the shell and shown that the resulting shell density profiles are in very good agreement with the profiles obtained from three-dimensional hydrodynamic simulations. We have also shown that the shell thickness measured at half of its peak density agrees well with the thickness of a simple uniform-density shell model. This allows us to model a fragment forming in the shell, due to gravitational instability, as a uniform-density oblate spheroid. We use this model to derive a dispersion relation for pressure assisted gravitational instability (PAGI). This dispersion relation (perturbation growth rate as a function of initial fragment wavenumber) depends on the pressure of the external medium: the higher $P_{_{\\rm EXT}}$, the larger the maximum unstable wavenumber (i.e. the smaller the smallest unstable fragment), and the faster the growth rate for an unstable fragment. If $P_{_{\\rm EXT}}\\!=\\!0$, the highest unstable wavenumber is $0.6$ times smaller than predicted by the standard thin-shell analysis using the same parameters; if $P_{_{\\rm EXT}}\\!=\\!\\infty$, the highest unstable wavenumber is $2.2$ times larger. The PAGI dispersion relation gives approximately the same range of unstable wavelengths as the thin-shell dispersion relation if $P_{_{\\rm EXT}}\\simeq P_{_{\\rm CRIT}}$, where $P_{_{\\rm CRIT}}$ is the critical external pressure for which self-gravity and external pressure contribute equally to confinement of the shell. Finally, we have demonstrated that the predictions of the PAGI dispersion relation agree rather well with the results of three-dimensional hydrodynamic simulations. In particular, (i) the PAGI dispersion relation predicts a maximum unstable wavenumber very similar to the simulations (but systematically $\\sim 25\\%$ higher); (ii) modulo this offset, the increase in $l_{_{\\rm PAGI}}^{^{\\rm MAX}}$ with $P_{_{\\rm EXT}}$ predicted by the PAGI dispersion relation is exactly mirrored by the simulations." }, "1005/1005.1269_arXiv.txt": { "abstract": "We present two \\textit{Suzaku} observations of the Galactic center microquasar \\mq~separated by approximately 700 days. The source was observed on both occasions after a transition to the spectrally hard state. Significant emission from \\mq~ is detected out to an energy of 300 keV, with no spectral break or turnover evident in the data. We tentatively measure a lower limit to the cut-off energy of $\\sim$ 380 keV. The spectra are found to be consistent with a Comptonized corona on both occasions, where the high energy emission is consistent with a hard power-law ($\\rm \\Gamma \\sim 1.8$) with a significant contribution from an accretion disc with a temperature of $\\sim$ 0.4 keV at soft X-ray energies. The measured value for the inner radius of the accretion disc is found to be inconsistent with the picture whereby the disc is truncated at large radii in the low-hard state and instead favours a radius close to the ISCO ($\\rm R_{in} \\sim 10 - 20~R_g$). ", "introduction": "Since the discovery of apparently super-luminal jets from the X-ray binary GRS 1915+105 \\citep{b12}, the Galactic microquasars have assumed a position of critical importance in our efforts to understand accretion physics and relativistic jet production \\citep{b40}. \\mq~ was discovered by the \\textit{Einstein} satellite \\citep{b1}. Subsequent observations revealed \\mq~to be the dominant source of hard X-rays ($\\rm > 20~keV$) in the direction of the Galactic center \\citep{b2,b3}, where the source is located approximately 50 arcmin from Sgr A$\\rm ^*$. The microquasar nature of \\mq~was discovered upon the observation of a double sided radio jet consistent with the X-ray position \\citep{b4}. Further VLA observations showed this radio source to be highly variable \\citep{b5}. Since its discovery \\mq~has been observed on numerous occasion at X-ray wavelengths. The column density towards this source is high given the proximity to the Galactic center and has been measured by \\textit{Chandra} to be $\\rm \\sim 1 \\times 10^{23}~cm^{-2}$ \\citep{b10}. Here, the spectrum was found to be consistent with a power-law ($\\rm \\Gamma \\sim 1.4$). \\textit{INTEGRAL} low-hard state observations have detected \\mq~up to energies of $\\sim$ 600 keV, where the spectrum is found to be consistent with a power-law ($\\rm \\Gamma \\sim 1.6$) up to 200 keV, with an additional component required at higher energies \\citep{b8}, see also \\citet{b9} for earlier \\textit{INTEGRAL/RXTE} observations. In the high-soft state, the hard X-ray flux decreases significantly, dropping below the \\textit{INTEGRAL} detection limit at energies $\\rm >$ 50 keV \\citep{b8}. \\citet{b11} reported on 5 years of \\textit{RXTE} monitoring, where they discovered a modulation with a period 12.73$\\rm \\pm$0.05 days which is attributed to the orbital period of the binary, in addition to a possible super-orbital modulation with a period of $\\sim$ 600 days. As the extinction at optical wavelengths is prohibitive ($\\rm A_v \\sim 50$), counterpart searches are required to take place in the infrared. While a number of candidate counterparts have been identified \\citep{b6,b7}, no variability has been observed from these, rendering them unlikely to be the actual counterpart. The current upper limit for the K$\\rm_s$-band magnitude of the counterpart is $\\geq$ 19.9 at the 95\\% confidence level. This is equivalent to a secondary spectral type of O or B if on the main sequence or K if a giant, at an assumed distance of 8.5 kpc. In this paper, we describe observations undertaken with the \\textit{Suzaku} X-ray observatory, while \\mq~was in the low-hard state. In \\S2, we describe the observations and extraction of source spectra. We proceed to analyze the data in \\S3. In \\S4, these results are compared to observations of other microquasars in the hard state, and finally our conclusions are presented in \\S5. ", "conclusions": "We present \\textit{Suzaku} observations of the Galactic center microquasar \\mq~ in two separate epochs taken after the system had transitioned into the low-hard state. The system is observed to be in the low-hard state at the time of our observations with an X-ray luminosity of $\\sim$ 1\\% Eddington. The spectra in each epoch are similar, being described by a model consisting of a soft thermal accretion disc component (T $\\sim$ 0.4 keV) and the broadband emission ($\\rm > 10 keV$) is found to be characterized by an unbroken power-law to at least 300 keV. Consistent with growing evidence from observations of numerous systems in the low-hard state (e.g. GX 339-4, SWIFT J1753.5-0127, XTE J1817-330, XTE J1118+480), we also find evidence that the accretion disc in \\mq~is not truncated at large radii in the low-hard state and instead resides close to the ISCO ($\\rm R_{in} \\sim 20~R_g$).\\\\" }, "1005/1005.1433_arXiv.txt": { "abstract": "{ Herschel PACS and SPIRE images have been obtained over a 30\\arcmin$\\times$30\\arcmin\\ area around the well-known carbon star CW Leo (IRC +10 216). An extended structure is found in an incomplete arc of $\\sim$22\\arcmin\\ diameter, which is cospatial with the termination shock due to interaction with the interstellar medium (ISM) as defined by Sahai \\& Chronopoulos from ultraviolet GALEX images. Fluxes are derived in the 70, 160, 250, 350, and 550 $\\mu$m bands in the region where the interaction with the ISM takes place, and this can be fitted with a modified black body with a temperature of 25$\\pm$3~K. Using the published proper motion and radial velocity for the star, we derive a heliocentric space motion of 25.1 \\ks. Using the PACS and SPIRE data and the analytical formula of the bow shock structure, we infer a de-projected standoff distance of the bow shock of $R_{0} = (8.0 \\pm 0.3) \\times 10^{17}$ cm. We also derive a relative velocity of the star with respect to the ISM of $(106.6 \\pm 8.7)/\\sqrt{n_{\\rm ISM}}$ \\ks, where n$_{\\rm ISM}$ is the number density of the local ISM. } ", "introduction": "Ever since the discovery paper by Becklin et al. (1969), the object IRC +10 216 (= AFGL 1381 = CW Leo) has spurred much interest. We now know that it is a carbon star in an advanced stage of stellar evolution on the asymptotic giant branch (AGB), pulsating and surrounded by an optically thick dust shell and large molecular circumstellar envelope (CSE). One aspect of study has been to constrain the properties of the CSE by answering questions such as what is the mass-loss rate and how has it changed with time, what kind of chemistry takes place, and what is the geometry and structure of the CSE? The deep optical images taken by Mauron \\& Huggins (1999, 2000), Mauron et al. (2003), and Le\\~{a}o et al. (2006) show that the dusty envelope is not smooth but consists of a series of arcs or incomplete shells. The average angular separation between the dust arcs suggests a timescale for the change in mass-loss rate of the order of 200--800 yr. The lack of kinematic information on the dust arcs precludes any firm conclusion about the true three-dimensional structure of the arcs or shells. From large-scale mapping at a relatively low angular resolution of the CO J=1-0 emission, Fong et al. (2003) discovered a series of large molecular arcs or shells at radii of $\\sim$100\\arcsec\\ in the outer envelope. They attribute these multiple shells as ``being the reverberations of a single Thermal Pulse erupting over 6000 yr ago.'' The timescale inferred from the spacing between these arcs is about 200--1000 yr. In the present paper, we discuss our new results on the outer shell of CW Leo from observations with the {\\em Herschel Space Observatory} (Pilbratt et al. 2010) and their connection to the results from {\\em GALaxy Evolution Explorer Space Observatory} (GALEX) by Sahai \\& Chronopoulos (2010, hereafter SC). For the present analysis, we adopted a distance of $d=$ 135\\,pc and a mass-loss rate \\mdot= 2.2$\\cdot 10^{-5}$ \\msolyr\\ (Groenewegen et al. 1998), a gas expansion velocity of V$_{\\rm exp}$ = 15.4 \\ks, a radial velocity V$_{\\rm LSR}= -25.5$ \\ks\\ (Groenewegen et al. 2002), corresponding to V$_{\\rm helio}= -18.6$\\,\\ks, and a proper motion (pm) $\\mu_{\\rm \\alpha} \\cos \\delta = +26 \\pm$6, $\\mu_{\\rm \\delta} =+4 \\pm$6 mas/yr (Menten et al. 2006). ", "conclusions": "\\label{discussion} \\subsection{Bow shock, thermal pulse, or both ?} The most widely accepted explanation of large detached shells is mass--loss variation (e.g., Olofsson et al. 1990, Zijlstra et al. 1992). AGB stars experience thermal pulses (TPs) during which intense mass loss ejections occur. A star can undergo several TPs, which would lead to the formation of concentric spherical detached shells (see Kerschbaum et al. 2010). Another explanation of detached envelopes is the interaction between the AGB wind and the interstellar medium (ISM) (e.g. Young et al. 1993, Martin et al. 2007, Ueta et al. 2006). In this scenario, the AGB wind is slowed down as it sweeps up material from the ISM. The piled up material forms a density enhancement that continue to expand due to the internal pressure. Shocks can occur if the relative velocity of the AGB wind with respect to the ISM is large enough. The thermal emission of the dust in the density enhancement at the shock interface between the stellar wind and the the ISM can be detected in the far-IR (Ueta et al. 2006, 2009). While a detached shell produced by TPs would be spherical, in the case of wind-ISM interaction the shape of the shell will depend on the space motion of the star through the ISM. A wind--ISM shell can look spherical as seen for R Cas (Ueta et al. 2009) if most of the space motion of the star relative to the ISM is in the radial direction. For stars with a high space motion (relative to the ISM), the shape of the bow shock will be more parabolic with the apex of the parabola in the direction of the star's motion relative to the ISM. In the case of CW\\,Leo, the shape of the extended emission and the position of the star suggest that the stellar wind has driven a shock into the ISM. The far-IR emission is probably caused by the thermal emission of the piled-up dust at the shock interface. One may think that the observed UV emission is produced by dust scattering of the interstellar light. However, the brightness ratio of FUV to NUV is $\\sim$10 (see Table~\\ref{TableF}), which is much higher than expected in that case (of the order of $\\sim$2.4; SC). The only other AGB star with UV data probing a wind--ISM interaction is Mira (Martin et al. 2007). Martin et al. suggest that collisional excitation of cool H$_2$ by hot electrons from the post-shock gas is responsible for the UV emission. The faint NUV emission can be explained by the H$_2$ emitting in the FUV band. SC suggest that the same mechanism may also be the dominant contributor to the FUV ring emission in CW Leo. The dust and the FUV emission have a similar spatial scale. For the planetary nebula NGC 6720, we note that van Hoof et al. (2010) find on the basis of PACS and SPIRE data that dust and H$_2$ are co-spatial and argue that H$_2$ has been formed on grain surfaces. \\begin{figure} \\begin{center} \\vspace{0cm} \\hspace{0cm} \\resizebox{9cm}{!}{ \\includegraphics[angle=90]{figures/14658fig2.ps}} \\caption{Intensity profiles as a function of the offset along the minor axis of the ellipse increasing from west to east. The profile for each wavelength was normalised to the intensity at 558$\\arcsec$ and shifted up for clarity. From bottom to top with normalisation factor in mJy/arcsec$^2$ and the vertical shift: FUV (0.065,0.0), 160~$\\mu$m (0.063,0.8), 250~$\\mu$m (0.023,1.1), 350~$\\mu$m (0.008,1.4), 500~$\\mu$m (0.004,1.8). The vertical dotted line indicates the position of the dust emission peak from the center of the ellipse.} \\label{Ipro} \\vspace{-4mm} \\end{center} \\end{figure} The structure seen in our data may also be the result of both mechanisms i.e., mass--loss variation and wind--ISM interaction. By estimating the interpulse period for CW Leo, we can check wether the dust shell we see is from an earlier TP that is now interacting with the ISM. Guelin et al. (1995) constrained the initial mass of CW Leo to be 3 \\less\\ $M$ \\less\\ 5 \\msol\\, based on the isotopic ratios of $^{24,25,26}$Mg. From the initial--final mass relation from Salaris et al. (2009), this implies a likely final mass (and essentially the current core mass) of 0.7--0.9 \\msol. From the core mass interpulse period relation of Wagenhuber \\& Groenewegen (1998) for solar metallicity stars, this implies an interpulse period of 6~000--33~000 years. Scaling the mass loss and the wind velocity in SC with our assumed values, the flow timescale in the unshocked and shocked winds is 19~900 and 56~000 yr (very long because of the very low velocity of 1.2~\\ks\\ in this region, see SC for details), for a total lower limit to the duration of the mass-loss phase of about 75~000 years. If TPs were to modulate the mass loss, as hypothesised for the origin for the detached shells discussed earlier, one could expect at least one other TP to have occurred during the time it took the envelope to expand this far. No obvious density enhancement is evident in the unshocked wind in the PACS and SPIRE images, which suggests that the interpulse period is at least 19~000 years. However, the dynamical and interpulse timescales are compatible, so it is possuble that the bow shock and dust emission are not only the result of a steady outflow interacting with the ISM, but might include the effect of an enhanced wind of short duration due to a recent TP (or pulses). \\begin{table}[!h] \\caption{Derived fluxes for the extended emission} \\begin{tabular}{cccccc} \\hline \\hline $\\lambda$ & Flux & $\\lambda$ & Flux & $\\lambda$ & Flux \\\\ ($\\mu$m) & (Jy) & ($\\mu$m) & (Jy) & ($\\mu$m) & (Jy) \\\\ \\hline 0.15 & $(20\\pm0.2)\\times10^{-4}$ & 70 & 3.51$\\pm$0.54 & 250 & 4.31$\\pm$0.66 \\\\ 0.23 & $ (2\\pm0.2)\\times10^{-4}$ & 160 & 7.70$\\pm$1.17 & 350 & 1.70$\\pm$0.27 \\\\ & & & & 550 & 0.73$\\pm$0.14 \\\\ \\hline \\end{tabular} \\label{TableF} \\vspace{-7mm} \\end{table} \\subsection{Space motion of the star and the ISM flow velocity} From the adopted radial velocity, distance, and proper motion, the heliocentric space velocity of the star is derived to be about 25.1 \\ks\\ at a heliocentric inclination angle of the space motion vector of 47.8 degrees (measured from the plane of the sky, away from us) and a PA of 81.3\\degr\\ for the proper-motion vector in the plane of the sky. In the PACS/SPIRE wavebands, the observed far-IR surface brightness is expected to be proportional to the column density of the dusty material in the shell because the optical depth of the shell is much lower than unity. Because the bow shock interface is a parabolic surface arbitrarily oriented in space, the column density tends to reach its highest value where the bow shock cone intersects with the plane of the sky including the central star. Thus, the apparent shape of the bow shock is the conic section of the bow shock. Therefore, given the analytical formula of the bow shock structure (Wilkin 1996), we can fit the observed surface brightness of the bow shock to the conic section of the bow shock cone to derive the heliocentric orientation of the bow shock (e.g. Ueta et al. 2008; 2009). From this fitting, we determined that the apex of the bow shock cone is oriented at $61\\fdg9 \\pm 0\\fdg3$ (this is degenerate, in a sense that it could be pointing away from us or towards us) with respect to the plane of the sky into the PA of $88^{\\circ}$ with the deprojected stand-off distance of $(8.0 \\pm 0.3) \\times 10^{17}$ cm. Using these numbers in the ram pressure balance equation $V_{\\star} = \\sqrt{\\frac{\\dot{M} \\; V_{\\rm exp}}{4 \\pi \\; \\mu_{\\rm H} \\; m_{\\rm H} \\; n_{\\rm ISM} \\; R_0^2} }$ for a mean nucleus number per hydrogen nucleus of $\\mu_{\\rm H}= 1.4$, the relative velocity of the star with respect to the ISM is $V_{\\star} = (106.6 \\pm 8.7)/\\sqrt{n_{\\rm ISM}}$ \\ks, where $n_{\\rm ISM}$ is the number density of the ISM local to CW Leo in cm$^{-3}$. By following the scheme of Johnson \\& Soderblom (1987, also see Ueta et al. 2008), the heliocentric space velocity components of the star can be converted to the heliocentric Galactic space velocity components [U, V, W] of [21.6$\\pm$3.9, 12.6$\\pm$3.5, -1.8$\\pm$3.3] \\ks\\ , and also to the LSR Galactic space velocity components of [30.6$\\pm$3.9, 24.6$\\pm$3.5, 5.2$\\pm$3.3] \\ks. The heliocentric ISM flow velocity is 117.6 \\ks\\ if the bow cone is facing us (i.e.~the apex pointing toward) or 82.6 \\ks\\ if the bow cone is facing away from us. \\begin{figure}[t] \\begin{center} \\vspace{0cm} \\hspace{0cm} \\resizebox{9cm}{!}{ \\includegraphics[angle=90]{figures/14658fig3.ps}} \\caption{Modified black body ($T$= 25$\\pm$3~K and $\\beta$= 1.6$\\pm$0.4) fit (dashed curve) to the derived fluxes (symbols) for the extended emission. } \\label{dustT} \\vspace{-4mm} \\end{center} \\end{figure} Wareing et al. successfully modelled bow shocks for AGB stars using 3-D hydrodynamical models (R\\,Hya and Mira in Wareing et al. 2006, 2007a). Their models were able to reproduce in great detail all the components of a bow shock (astropause, astrotail, vortices) and constrain the space velocity of the star, the ISM density, and the mass--loss rate. The models were built by varying the velocity of the star relative to the ISM, the ISM density and the mass loss rate and the effect of those physical parameters on the shape of the bow shock can be seen in Wareing et al. (2007b). From the Wareing models, we can see that for ISM densities $\\le$2~cm$^{-3}$ (which implies $V_{\\star} \\ge$75 \\ks) and after an evolution of 50~000~yr into the AGB phase, the bow shock does not resemble our data. The models have a more flattened bow shock with the star closer to the apex of the shock. This is for a mass--loss rate between 5$\\times 10^{-7}$~\\msolyr\\ and 10$^{-6}$~\\msolyr, which is an order of magnitude lower than the value assumed for CW\\,Leo. For a higher mass--loss rate, we would expect an even more important departure from sphericity with a significant tail of ejecta. This suggests that the density of the local ISM to CW\\,Leo is probably higher than 2~cm$^{-3}$ implying an upper limit of 75 \\ks to $V_{\\star}$, the stellar velocity relative to the ISM." }, "1005/1005.2606_arXiv.txt": { "abstract": "Sky coverage is one of the most important pieces of information about astronomical observations. We discuss possible representations, and present algorithms to create and manipulate shapes consisting of generalized spherical polygons with arbitrary complexity and size on the celestial sphere. This shape specification integrates well with our Hierarchical Triangular Mesh indexing toolbox, whose performance and capabilities are enhanced by the advanced features presented here. Our portable implementation of the relevant spherical geometry routines comes with wrapper functions for database queries, which are currently being used within several scientific catalog archives including the Sloan Digital Sky Survey, the Galaxy Evolution Explorer and the Hubble Legacy Archive projects as well as the Footprint Service of the Virtual Observatory. ", "introduction": "\\label{sec:intro} Astronomers have to keep accurate records of where their observations are located on the sky. Beyond the direction and angular size of the field of view, we have detailed information available about the precise sky coverage derived from the exact shape of our detectors. This coverage is invaluable for most statistical studies, e.g., luminosity functions or especially analyses of spatial clustering. The de facto standard of the Flexible Image Transport System \\citep[FITS;][]{fits} has reserved keywords for the World Coordinate System \\citep[WCS;][]{wcs} specification, and parameters that specify the transformation from image pixels to sky coordinates (and reverse), are present in the header of most FITS files. While the WCS is perfectly adequate for individual exposures, multiple observations are difficult to describe in a single system. Every field has potentially a separate coordinate system, hence moving from field to field is convoluted, and makes it cumbersome to answer even simple questions, e.g., whether two separate fields overlap. The footprint of a large-area survey might be complicated but the small-scale irregularities are even more problematic. Not only the depth of a survey changes as a function of the position on the sky, but parts of the observations are often censored for various reasons, such as bright stars blocking the view, satellite trails, artifacts from reflections. If we would like to represent all these on the sky, we need a scalable solution that works for shapes of arbitrary complexity and size from the subpixel level to the entire sky. We need tools to create and manipulate these descriptions right there where the data are and utilize the information efficiently. Geographic Information Systems (GIS) were designed with a similar goal in mind. There are however subtle differences, that are large enough that GIS systems are not quite applicable to astronomy directly. The modern mapping systems do not extend much beyond the basic features of projected maps but provide very efficient tools for finding nearby places and shortest routes, etc. Even the most complicated GIS shapes are limited to spherical polygons whose sides are great circles (or straight lines in the projection.) In astronomy, the approximations with such polygons would be unacceptably inaccurate or prohibitively redundant, hence there is need for an extended set of features to represent the geometries of the observations and surveys. Some of the concepts discussed in the paper are not new and have been introduced and studied previously in different fields. Books by \\citet{samet89,samet90} detail the quad-trees for spatial searches that \\citet{barrett94} advocated using for astronomy data. \\cite{fekete90a,fekete90b,short95} developed an icosahedron-based methodology for Earth sciences, and \\citet{kunszt00, kunszt01} introduced the convex description and the Hierarchical Triangular Mesh, a triangulation that was also used by \\citet{lee98} with a different numbering scheme. A similar representation of shapes is also found in \\citet{mangle1} and \\citet{mangle2}. \\citet{goodchild91,goodchild92} and \\citet{song00} introduced the Discrete Global Grid for GIS systems, an equal area variant of the same triangulation idea. The integration to relational databases is discussed in \\citet{gray04}. The goal this paper is to provide the astronomy community with a complete and consistent view of the current and much improved methodology built on a new fully functional spherical geometry framework, and describe the implementations and interfaces used in several astronomy archives and tools today including the Sloan Digital Sky Survey \\citep{cas}, the Hubble Legacy Archive \\citep{greene07}, the Galaxy Evolution Explorer and the NVO Footprint Service \\citep{adass_footprint}. In Section~\\ref{sec:shapes} we describe how to specify spherical shapes and manipulate them. Section~\\ref{sec:geom} deals with the spherical geometry of the region representations and Section~\\ref{sec:htm} discusses an efficient search method based on the Hierarchical Triangular Mesh. In Section~\\ref{sec:lib} we provide details of our software solution including the implementation of the database routines and their usage. Section~\\ref{sec:sum} summarizes the main results, along with current and future applications in astronomy. ", "conclusions": "" }, "1005/1005.5026_arXiv.txt": { "abstract": "The mass of the Galactic dark matter halo is under vivid discussion. A recent study by Xue et al. (2008, ApJ, 684, 1143) revised the Galactic halo mass downward by a factor of $\\sim$2 relative to previous work, based on the line-of-sight velocity distribution of $\\sim$2400 blue horizontal-branch (BHB) halo stars. The observations were interpreted in a statistical approach using cosmological galaxy formation simulations, as only four of the 6D phase-space coordinates were determined. Here we concentrate on a close investigation of the stars with highest negative radial velocity from that sample. For one star, SDSSJ153935.67+023909.8 (J1539+0239 for short), we succeed in measuring a significant proper motion, i.e. full phase-space information is obtained. We confirm the star to be a Population~II BHB star from an independent quantitative analysis of the {\\em SDSS} spectrum -- providing the first NLTE study of any halo BHB star -- and reconstruct its 3D trajectory in the Galactic potential. J1539+0239 turns out as the fastest halo star known to date, with a Galactic rest-frame velocity of 694$^{+300}_{-221}$\\,\\kms\\ (full uncertainty range from Monte Carlo error propagation) at its current position. The extreme kinematics of the star allows a significant lower limit to be put on the halo mass in order to keep it bound, of $M_{\\rm halo}\\ge1.7_{-1.1}^{+2.3}\\times10^{12}$\\,\\msun. We conclude that the Xue et al. results tend to underestimate the true halo mass as their most likely mass value is consistent with our analysis only at a level of 4\\%. However, our result confirms other studies that make use of the full phase-space information.\\\\ ", "introduction": "\\label{sec:intro} Knowledge of the properties of dark matter halos is an important issue for our understanding of galaxy formation and evolution, and for unveiling the nature of dark matter. The halo of the Milky Way therefore is of highest interest, as it allows unique observational constraints to be obtained for testing theoretical models \\citep[e.g.][]{1996ApJ...462..563N}. Several observational campaigns -- e.g. the Sloan Digital Sky Survey \\citep[\\sdss,][]{2000AJ....120.1579Y}, the RAdial Velocity Experiment \\citep[\\emph{RAVE},][]{2006AJ....132.1645S}, the Sloan Extension for Galactic Understanding and Exploration \\citep[\\emph{SEGUE},][]{2009AJ....137.4377Y} -- provide the tracers for studying halo properties, like the total mass of the halo and its extent. Several studies in the past decade determined the halo mass from ever increasing samples of halo stars, globular clusters and/or satellite galaxies. However, it is in fact only a few objects at highest velocity that are affecting largely the mass estimates \\citep{2003A&A...397..899S,2007MNRAS.379..755S}. While larger halo masses of about $2\\times10^{12}$\\,\\msun\\ were favoured earlier \\citep{1999MNRAS.310..645W,2003A&A...397..899S}, lower masses of about half this value were derived more recently \\citep{2005MNRAS.364..433B,2007MNRAS.379..755S,2008ApJ...684.1143X}. The precise value determines, e.g. among others, whether satellite galaxies like the Magellanic Clouds \\citep[e.g.][]{2006ApJ...652.1213K,2009AJ....137.4339C} or hyper-velocity star\\footnote{Hyper-velocity stars move at such high velocity that they may be gravitationally unbound to the Galaxy. Their supposed place of origin is the Galactic center, where they may have been accelerated by gravitational interactions with the super-massive black hole \\citep{hills88}.} candidates \\citep{2009ApJ...691L..63A} are on bound orbits, or not. Most of the previous studies had to rely substantially on the {\\em distributions} of radial velocities in their samples to derive their conclusions as full space motions for halo objects are unavailable in many cases at present. In such cases only four coordinates (i.e. two position values, distance and radial velocity, RV) of the 6D phase space are determined and the missing data (the proper motion components) are handled in a statistical approach. E.g., radial velocities for more than 10,000 blue halo stars from the \\sdss were measured in the yet most extensive study by \\citet{2008ApJ...684.1143X}. The sample was composed of blue horizontal-branch, blue straggler, and main-sequence stars with effective temperatures roughly between 7,000 and 10,000\\,K according to their colours. Here, we focus on the fastest stars of the Xue et al. sample in terms of {\\em negative} line-of-sight velocity, indicating a {\\em bound} orbit. For one of them we were able to measure a significant proper motion, which allowed a detailed 3D kinematic investigation when combined with a quantitative spectroscopic analysis that facilitated the determination of the star's distance. \\object{SDSSJ153935.67+023909.8} (J1539+0239 for short) is an inbound\\footnote{An extragalactic origin of the star as an unbound low-mass hyper-velocity star from another galaxy is imaginable but unlikely. The local volume is devoid of galaxies in a wide range (several 10\\,$\\deg$) around the infall direction of J1539+0239 at present \\citep[for the Local Group see][and later discoveries]{1999A&ARv...9..273V}. If ejected early after its formation, the star could have travelled $\\sim$8\\,Mpc during its lifetime, such that a final answer on this may be only obtained when full space motions of the nearby galaxies are known.} Population~II horizontal branch star with a Galactic rest-frame (GRF) velocity of $\\sim$700\\,\\kms\\ at its current position, making it the fastest halo object known. This allows a significant lower limit for the total halo mass of the Galaxy to be set. The present work provides a glimpse to the detailed kinematic investigations feasible once the European Space Agency's Gaia satellite mission \\citep[e.g.][]{2005ESASP.576.....T} becomes operational, at much higher precision. ", "conclusions": "\\label{sec:conclusion} We reported the quantitative spectral analysis of a high-velocity star from the sample of faint blue halo stars of \\citet{2008ApJ...684.1143X}. J1539+0239 was confirmed to be a Population~II blue horizontal branch star with a low metallicity of $[$Fe/H$]=-2.0$ and the characteristic enhancement of $\\alpha$-elements. Hereby, we performed a NLTE analysis of a halo BHB star for the first time. While the majority of the weak lines were confirmed to be formed close to LTE conditions, many of the stronger metal lines -- which are important diagnostics at the spectral resolution achieved within the {\\em SDSS} -- show pronounced NLTE strengthening, with the differences between the derived LTE and NLTE abundances amounting to 0.1\\,dex to 0.5\\,dex typically. In addition to information on the chemical composition, the radial velocity, proper motion and spectroscopic distance were derived and a detailed kinematical analysis was performed. Carrying out kinematical numerical experiments using the Galactic potential of \\citet{1991RMxAA..22..255A} in order to obtain an orbit of J1539+0239 gravitationally bound to the Milky Way, we found that the mass of the dark halo has to be at least $M_{\\rm halo}\\sim1.7_{-1.1}^{+2.3}\\times10^{12}$\\,\\msun (absolute uncertainties from extrema in MC error propagation) . This mass limit is in good agreement with several previous studies \\citep{1999MNRAS.310..645W,2003A&A...397..899S,2009ApJ...691L..63A}. However, the significantly lower most likely mass value of \\cite{2008ApJ...684.1143X} is consistent with our analysis only at a level of 4\\%, i.e. it likely underestimates the Galactic dark halo mass. We conclude, that if the kinematics of a halo star is extraordinary enough, and the errors within the analysis are small, even {\\em one} star alone can provide a significant lower limit to the dark matter halo mass, and to the total mass of the Milky Way (halo\\,+\\,bulge\\,+\\,disk), here $M_{\\rm total}\\ge1.8_{-1.1}^{+2.3}\\times10^{12}$\\,\\msun. The determining factor is that full kinematic information is available, as it will become routine in the era of the Gaia space mission, at much higher precision." }, "1005/1005.5356_arXiv.txt": { "abstract": "{} {We determine the physical properties (spin state and shape) of asteroid (21) Lutetia, target of the ESA Rosetta mission, to help in preparing for observations during the flyby on 2010 July 10 by predicting the orientation of Lutetia as seen from Rosetta.} {We use our novel KOALA inversion algorithm to determine the physical properties of asteroids from a combination of optical lightcurves, disk-resolved images, and stellar occultations, although the latter are not available for (21) Lutetia.} {We find the spin axis of (21) Lutetia to lie within 5\\degr~of ($\\lambda = 52$\\degr, $\\beta = -6$\\degr) in Ecliptic J2000 reference frame (equatorial $\\alpha = 52$\\degr, $\\delta = +12$\\degr), and determine an improved sidereal period of 8.168\\,270 $\\pm$ 0.000\\,001 h. This pole solution implies the southern hemisphere of Lutetia will be in ``\\textsl{seasonal}'' shadow at the time of the flyby. The apparent cross-section of Lutetia is triangular as seen ``\\textsl{pole-on}'' and more rectangular as seen ``\\textsl{equator-on}''. The best-fit model suggests the presence of several concavities. The largest of these is close to the north pole and may be associated with large impacts. } {} ", "introduction": "\\indent The origin and evolution of the Solar System and its implications for early planetesimal formation are key questions in planetary science. Unlike terrestrial planets, which have experienced significant mineralogical evolution, through endogenic activity, since their accretion, small Solar System bodies have remained essentially unaltered. Thus, a considerable amount of information regarding the primordial planetary processes that occurred during and immediately after the accretion of the early planetesimals is still present among this population. Consequently, studying asteroids is of prime importance in understanding the planetary formation processes \\citep{2002-AsteroidsIII-1-Bottke} and, first and foremost, requires a reliable knowledge of their physical properties (size, shape, spin, mass, density, internal structure, etc.) in addition to their compositions and dynamics. Statistical analyses of these parameters for a wide range of asteroids can provide relevant information about inter-relationships and formation scenarios. \\\\ \\indent In this respect, our observing program with adaptive optics, allowing diffraction-limited observations from the ground with 10 m-class telescopes, has now broken the barrier which separated asteroids from real planetary worlds \\citep[\\textsl{e.g.,}][]{2007-Icarus-191-Conrad, 2008-AA-478-Carry, 2009-Icarus-202-Drummond, 2010-Icarus-205-Carry-a, 2010-AA--Drummond}. Their shapes, topography, sizes, spins, surface features, albedos, and color variations can now be directly observed from the ground. This opens these objects to geological, rather than astronomical-only, study. While such surface detail is only possible for the largest asteroids, our main focus is on determining accurately the size, shape, and pole. Among them, we have observed (21) Lutetia, an asteroid that will be observed in-situ by the ESA Rosetta mission.\\\\ \\indent The Rosetta Mission will encounter its principal target, the comet 67P/Churyumov-Gerasimenko, in 2014. However, its interplanetary journey was designed to allow close encounters with two main-belt asteroids: (2867) \\v{S}teins and (21) Lutetia. The small asteroid (2867) \\v{S}teins was visited on 2008 September 5 at a minimum distance of about 800 km \\citep{2009-DPS-40-Schulz} and (21) Lutetia will be encountered on 2010 July 10. Knowing the geometry of the flyby (\\textsl{e.g.}, visible hemisphere, sub-spacecraft coordinates as function of time, and distance) before the encounter is crucial to optimize the observation sequence and schedule the on-board operations. The size of Lutetia \\citep[estimated at $\\sim$100 km, see][]{ 2002-AJ-123-Tedesco, PDSSBN-Iras, 2006-AA-447-Mueller} allows its apparent disk to be spatially resolved from Earth. Our goal is therefore to improve knowledge of its physical properties to prepare for the spacecraft flyby. \\\\ \\indent Lutetia, the Latin name for the city of Paris, is a main-belt asteroid (semi-major axis 2.44 AU) that has been studied extensively from the ground \\citep[see][for a review, primarily of recent observations]{2007-SSRv-128-Barucci}. Numerous studies have estimated indirectly its spin \\citep[by lightcurve, \\textsl{e.g.,}][]{ 1987-KFNT-3-Lupishko-a, 1992-AA-95-Dotto, 2003-Icarus-164-Torppa}. Size and albedo were reasonably well determined in the 1970s by \\citet{1977-Icarus-31-Morrison} using thermal radiometry (108--109 km), and by \\citet{1976-AJ-81-Zellner} using polarimetry (110 km). Five somewhat scattered IRAS scans \\citep[\\textsl{e.g.,}][]{2002-AJ-123-Tedesco, PDSSBN-Iras} yielded a higher albedo and smaller size than the dedicated observations in the 1970s. \\citet{2006-AA-447-Mueller} derived results from new radiometry that are roughly compatible with the earlier results or with the IRAS results, depending on which thermal model is used. \\citet{2008-AA-479-Carvano} later derived a lower albedo from ground-based observations, seemingly incompatible with previous works. Radar data analyzed by \\citet{1999-Icarus-140-Magri, 2007-Icarus-186-Magri} yielded an effective diameter for Lutetia of 116 km; reinterpretation of those data and new radar observations \\citep{2008-Icarus-195-Shepard} suggest an effective diameter of 100 $\\pm$ 11 km and an associated visual albedo of 0.20. Recent HST observations of Lutetia \\citep{2009-arXiv-Weaver} indicate a visual albedo of about 16\\%, a result based partly on the size/shape/pole determinations from our work in the present paper and \\citet{2010-AA--Drummond}.\\\\ \\indent Lutetia has been extensively studied using spectroscopy in the visible, near- and mid-infrared and its albedo measured by polarimetry and thermal radiometry \\citep{1975-ApJ-195-McCord, 1975-Icarus-25-Chapman, 1976-AJ-81-Zellner, 1978-Icarus-35-Bowell, 1995-Icarus-117-Rivkin, 1999-Icarus-140-Magri, 2000-Icarus-145-Rivkin, 2004-AA-425-Lazzarin, 2005-AA-430-Barucci, 2006-AA-454-Birlan, 2007-AA-470-Nedelcu, 2008-AA-477-Barucci, 2008-Icarus-195-Shepard, 2009-Icarus-202-DeMeo, 2009-Icarus-202-Vernazza, 2009-AA-498-Lazzarin, lazzarin-2010, 2010-AA-513-Perna, 2010-AA-Belskaya}. We present a discussion on Lutetia's taxonomy and composition in a companion paper \\citep{2010-AA--Drummond}.\\\\ \\indent Thermal infrared observations used to determine the size and albedo of Lutetia were initially inconsistent, with discrepancies in diameters and visible albedos reported \\citep[\\textsl{e.g.,}][]{% 1976-AJ-81-Zellner, 1989-Icarus-78-Lupishko, 1996-PSS-44-Belskaya, 2002-AJ-123-Tedesco, 2006-AA-447-Mueller, 2008-AA-479-Carvano}. \\citet{2006-AA-447-Mueller} and \\citet{2008-AA-479-Carvano}, however, interpreted these variations as an indication of surface heterogeneity, inferring that the terrain roughness of Lutetia increased toward northern latitudes\\footnote{our use of ``northern hemisphere'' refers to the hemisphere in the direction of the positive pole as defined by the right-hand rule from IAU recommendations \\citep{2007-CeMDA-98-Seidelmann}}, that the crater distribution is different over the northern/southern hemispheres, and includes a possibility of one or several large craters in Lutetia's northern hemisphere. Indeed, the \\textsl{convex} shape model derived from the inversion of 32 optical lightcurves \\citep{2003-Icarus-164-Torppa} displays a flat top near the north pole of Lutetia. \\citet{2002-AA-383-Kaasalainen} have shown that large flat regions in these convex models could be a site of concavities. The southern hemisphere is not expected to be free from craters however, as \\citet{2010-AA-513-Perna} detected a slight variation of the visible spectral slope, possibly due to the presence of large craters or albedo spots in the southern hemisphere.\\\\ \\indent In this paper, we present simultaneous analysis of adaptive-optics images obtained at the W. M. Keck and the European Southern Observatory (ESO) Very Large Telescope (VLT) observatories, together with lightcurves, and we determine the shape and spin state of Lutetia. In section~\\ref{sec: obs}, we present the observations, in section~\\ref{sec: shape} the shape of Lutetia, and finally, we describe the geometry of the upcoming Rosetta flyby in section~\\ref{sec: flyby}. ", "conclusions": "} \\indent We have reported disk-resolved imaging observations of (21) Lutetia obtained with the W. M. Keck and Very Large Telescope observatories in 2007, 2008, and 2009. We have derived the shape and spin of (21) Lutetia using the Knitted Occultation, Adaptive-optics, and Lightcurve Analysis (KOALA) method, which is based on combining these AO images with optical lightcurves gathered from over four decades.\\\\ \\indent The shape of (21) Lutetia is well described by a Camembert wedge, and our shape model suggests the presence of several concavities near its north pole and around its equator. The spin axis of Lutetia is tilted with respect to its orbital plane, much like Uranus, implying strong seasonal effects on its surface. At the time of the Rosetta flyby, (21) Lutetia's northern hemisphere will be illuminated while the southern hemisphere will be in long-term darkness, hindering the size determination from Rosetta.\\\\ \\indent The next opportunity to observe Lutetia's shortest dimension, impacting its volume determination, will occur in 2011 July, one year after Rosetta flyby, when the sub-Earth point will be close to its equator (SEP$_\\beta$ of +31\\degr). During this time, observations using large telescopes equipped with adaptive-optics will allow refinement of Lutetia's short dimension and thus improve the volume determination. This ground-based support will be essential to take advantage of the high-precision mass determination provided by the spacecraft deflection observed during the flyby." }, "1005/1005.0868_arXiv.txt": { "abstract": "In Brans-Dicke theory a non-linear self interaction of a scalar field $\\phi$ allows a possibility of realizing the late-time cosmic acceleration, while recovering the General Relativistic behavior at early cosmological epochs. We extend this to more general modified gravitational theories in which a de Sitter solution for dark energy exists without using a field potential. We derive a condition for the stability of the de Sitter point and study the background cosmological dynamics of such theories. We also restrict the allowed region of model parameters from the demand for the avoidance of ghosts and instabilities. A peculiar evolution of the field propagation speed allows us to distinguish those theories from the $\\Lambda$CDM model. ", "introduction": "Scalar-tensor gravitational theories have been widely studied as an alternative to General Relativity (GR). In these theories a scalar-field degree of freedom $\\phi$ is coupled to the Ricci scalar $R$ through a coupling of the form $F(\\phi)R$ \\cite{FujiiMaeda}. For example, such a coupling arises in low energy effective string theory as a result of the dilaton coupling with gravitons \\cite{GasVene}. The discovery of dark energy in 1998 \\cite{SNIa} has also stimulated the study for the modifications of gravity on large distances (see Refs.~\\cite{darkreview} for recent reviews). The well-known example of scalar-tensor theories is Brans-Dicke (BD) theory \\cite{BDtheory} in which $F(\\phi)$ is proportional to $\\phi$ with a non-canonical kinetic term $(-\\omega_{\\rm BD}/\\phi)(\\nabla \\phi)^2$, where $\\omega_{\\rm BD}$ is the so-called BD parameter. In original BD theory without a field potential, the BD parameter is constrained to be $\\omega_{\\rm BD}>40000$ from local gravity tests in the solar system \\cite{Will}. This comes from the fact that the coupling between the massless field $\\phi$ and non-relativistic matter needs to be suppressed to avoid the propagation of the fifth force. Under this bound, the deviation from GR (which corresponds to the limit $\\omega_{\\rm BD} \\to \\infty$) is too small to be detected in current cosmological observations. There are two ways to recover the General Relativistic behavior in high density regimes relevant to solar system experiments. One is the so-called chameleon mechanism \\cite{chame} in which the presence of the field potential $V(\\phi)$ allows a possibility of having a density-dependent effective mass of the field. Provided that the effective mass is sufficiently large in the regions of high density, a spherically symmetric body can have a thin-shell around its surface so that the effective coupling between $\\phi$ and matter is suppressed outside the body. This mechanism works not only for BD theory \\cite{chame,TUMTY} but also for metric $f(R)$ gravity \\cite{fRchame}, because the latter can be regarded as a special case of BD theory with $\\omega_{\\rm BD}=0$ \\cite{ohanlon,Chiba}. For viable dark energy models based on BD theory and metric $f(R)$ gravity the early cosmological evolution mimics that of GR, but the deviation from GR becomes important at late times (i.e. low density regimes) \\cite{TUMTY,fRviable,fRviable2}. Another way for the recovery of GR in the regions of high density is to introduce non-linear self interactions of a scalar field, e.g., $\\xi (\\phi) \\square \\phi (\\partial^{\\mu} \\phi \\partial_{\\mu} \\phi)$, where $\\xi$ is a function in terms of $\\phi$. There have been attempts to restrict the form of the Lagrangian by imposing the ``Galilean'' symmetry $\\partial_{\\mu} \\phi \\to \\partial_{\\mu} \\phi +b_{\\mu}$ \\cite{Nicolis,Deffayet,Chow,Rham,Gannouji}. The self interaction of the form $\\square \\phi (\\partial^{\\mu} \\phi \\partial_{\\mu} \\phi)$ respects the Galilean symmetry in the Minkowski background. In the Dvali-Gabadadze-Porrati (DGP) braneworld model \\cite{DGP} the field self interaction $\\square \\phi (\\partial^{\\mu} \\phi \\partial_{\\mu} \\phi)$ has been employed to recover the General Relativistic behavior for the length scale smaller than the so-called Vainshtein radius \\cite{DGPse1,DGPse2} (see also Refs.~\\cite{Babichev}). While this ``self screening'' mechanism (called the Vainshtein mechanism \\cite{Vain}) can be at work for consistency with solar system experiments, the DGP model is unfortunately plagued by a ghost problem \\cite{DGPse2,DGPghost} as well as incompatibility with a number of observational constraints \\cite{DGPobser}. In BD theory with the field self interaction term $\\xi (\\phi) \\square \\phi (\\partial^{\\mu} \\phi \\partial_{\\mu} \\phi)$, there exist de Sitter (dS) solutions for $\\xi(\\phi) \\propto \\phi^{-2}$ and $\\omega_{\\rm BD}<-4/3$ \\cite{Silva}\\footnote{For constant $\\xi (\\phi)$ there exist no consistent dS solutions, but the presence of other field self interaction terms that respect Galilean symmetry in the Minkowski background allows a possibility for giving rise to dS solutions \\cite{Gannouji}.}. Moreover the General Relativistic behavior is recovered in early cosmological epochs during which the field is nearly frozen. The solutions finally approach the dS attractor at which $\\dot{\\phi}/\\phi$ is constant. As long as $\\dot{\\phi}/\\phi$ is positive, one can avoid the appearance of ghosts and instabilities associated with the field propagation speed. Moreover this model gives rise to several interesting observational signatures such as the modified growth of matter perturbations and anti-correlations in the cross-correlation of large scale structure and the integrated Sachs-Wolfe effect in cosmic microwave background anisotropies \\cite{Silva,Kobayashi1}. In this paper we consider the general action (\\ref{action}) below without the field potential and derive the functional forms of $F(\\phi)$, $B(\\phi)$, and $\\xi (\\phi)$ from the requirement of obtaining dS solutions with $\\dot{\\phi}/\\phi=$\\,constant. The three functions are restricted to be of the power-law forms in terms of $\\phi$, which include BD theory with $\\xi (\\phi) \\propto \\phi^{-2}$ as a special case. We discuss the cosmological viability of such theories by analyzing the stability of fixed points for the background field equations. We also study conditions for the avoidance of ghosts and instabilities to find the region of viable model parameters. ", "conclusions": "\\label{conclude} We have constructed modified gravitational models of dark energy starting from the general action (\\ref{action}) without a field potential. The presence of a non-linear self interaction $\\xi(\\phi) \\square \\phi (\\partial^{\\mu} \\phi \\partial_{\\mu} \\phi)$ allows us to recover the General Relativistic behavior in the regions of high density. The functions $F(\\phi)$, $B(\\phi)$, and $\\xi(\\phi)$ are restricted in the power-law forms given in Eq.~(\\ref{funchoice}) from the demand of obtaining dS solutions at late times. The theory with $n=2$ corresponds to Brans-Dicke theory with the self-interaction $\\xi(\\phi) \\propto \\phi^{-2}$, which was recently studied in literature \\cite{Silva}. If $\\xi(\\phi)=0$, the theories with the functions $F(\\phi)=M_{\\rm pl}^2 (\\phi/M_{\\rm pl})^{3-n}$ and $B(\\phi)=\\omega ( \\phi/M_{\\rm pl})^{1-n}$ are equivalent to Brans-Dicke theory by introducing a new field $\\chi \\equiv F(\\phi)$. However, except for $n=2$, the presence of the field self interaction term does not allow us to express theories with the functions (\\ref{funchoice}) as Brans Dicke theory plus the term $\\xi(\\chi) \\square \\chi (\\partial^{\\mu} \\chi \\partial_{\\mu} \\chi)$ by such a field redefinition. We have derived the dynamical equations (\\ref{auto1})-(\\ref{auto3}) to discuss the background cosmological evolution for the theories with the functions (\\ref{funchoice}). The evolution of the dimensionless variables $x$ and $y$ during radiation and matter eras can be analytically estimated, which is consistent with the results obtained by numerical integrations. The presence of these epochs demands the condition $n \\le 3$. The variables $x$ and $y$ at the dS point, denoted as $x_{\\rm dS}$ and $y_{\\rm dS}$ respectively, are determined by solving Eqs.~(\\ref{xdS}) and (\\ref{ysolu}) for given $n$ and $\\omega$. We also studied the stability of the dS point by considering the evolution of homogenous curvature perturbations and found that the stability condition is given by $(n-3)x_{\\rm dS}<3$. The viable model parameter space can be restricted further by studying conditions for the avoidance of ghosts and instabilities. The no-ghost condition corresponds to $Q_s>0$, where $Q_s$ is defined in Eq.~(\\ref{Qdef}). Provided $F(\\phi)>0$, we require that $x>0$ to avoid ghosts during the cosmological evolution from the radiation era to the dS epoch. We note that the case $n=2$ is special because of the divergent behavior of the quantity $Q_s$ at the dS point. Interestingly the stability of the dS point is automatically satisfied for $n \\le 3$ and $x_{\\rm dS}>0$. The instability of perturbations can be avoided for $c_s^2>0$, where the propagation speed squared $c_s$ is defined in Eq.~(\\ref{cses}). At the dS point $c_s^2$ reduces to Eq.~(\\ref{eq:csdS}), which is positive for $n \\ge 2$ under the conditions $n \\le 3$ and $x_{\\rm dS}>0$. Hence the viable parameter region of $n$ is constrained to be $2 \\le n \\le 3$. For the existence of the dS point the parameter $\\omega$ is restricted in the range $\\omega<-n(n-3)^2$ from Eq.~(\\ref{xdS}). When $n \\neq 3$ the evolution of $c_s^2$ during radiation and matter dominated epochs can estimated as Eqs.~(\\ref{csra}) and (\\ref{csma}) respectively, which remain subluminal. During the transition from the matter era to the dS epoch the propagation speed can be superluminal, depending on the values of $\\omega$ and $n$. The avoidance of the temporal superluminal propagation gives the lower bound on $\\omega$ for each value of $n$. In Fig.~\\ref{omegafig} we have plotted the viable model parameter space and the region of the subluminal propagation in the $(n, \\omega)$ plane. When $n=3$ it is difficult to avoid the appearance of the superluminal mode during the matter era because of the specific evolution of the variables $x$ and $y$. It will be of interest to study the evolution of matter density perturbations and gravitational potentials to confront our theories with the observations of large scale structure, cosmic microwave background, and weak lensing. We leave the detailed analysis of cosmological perturbations for future work." }, "1005/1005.1163_arXiv.txt": { "abstract": "We present the first multi-fluid analysis of a dense neutron star core with a deconfined colour-flavour-locked superconducting quark component. Accounting only for the condensate and (finite temperature) phonons, we make progress by taking over results for superfluid $^4$He. The resultant two-fluid model accounts for a number of additional viscosity coefficients (compared to the Navier-Stokes equations) and we show how they enter the dissipation analysis for an oscillating star. We provide simple estimates for the gravitational-wave driven r-mode instability, demonstrating that the various phonon processes that we consider are not effective damping agents. Even though the results are likely of little direct astrophysical importance (since we consider an overly simplistic stellar model) our analysis represents significant technical progress, laying the foundation for more detailed numerical studies and preparing the ground for the inclusion of additional aspects (in particular associated with kaons) of the problem. ", "introduction": "The state of matter at extreme densities continues to be an issue of vigorous investigation. The problem is complicated, not only from the theoretical point of view, but also by the fact that laboratory experiments are restricted. For example, while colliders like RHIC at Brookhaven, GSI in Darmstadt and the LHC at CERN probe hot quark-gluon plasmas they will never be able to explore the cold, extreme high density, region of the QCD phase diagram. In order to test our understanding of the relevant physics we need to turn to astrophysics, and the dynamics of compact stars. In fact, ``neutron stars'' represent unique laboratories of such extreme physics. With core densities reaching about one order of magnitude beyond nuclear saturation, they are likely to contain exotic states of matter like hyperon phases with net strangeness and/or deconfined quarks. It is well-established that these states of matter should exhibit superfluidity/superconductivity at the relevant temperatures (neutron stars are born with temperatures $\\sim 10^{12}$~K and rapidly cool below $\\sim 10^9$~K). Moreover, observed radio pulsar glitches provide strong evidence for the presence of a partially decoupled superfluid component in these systems. The modelling of the dynamics of these, potentially \\underline{very} complex, objects presents a serious challenge. A key aspect of the problem concerns the fact that a superfluid system has additional dynamical degrees of freedom. This is well-known from experiments on laboratory systems like $^4$He, which exhibit a second sound associated with thermal waves \\cite{khalatnikov,putterman}. Analogous ``superfluid'' modes have been studied in detail for superfluid neutron-proton-electron mixtures relevant for the outer core of a neutron star \\cite{epstein,mendell,ac01}. For simplicity, these studies have almost exclusively ignored thermal effects (the work in \\cite{gusakov} is a notable exception). While this is a useful first approximation it is clear that the zero temperature assumption must be relaxed in a realistic model. Basically, due to the density dependence of the various superfluid pairing gaps (see \\cite{NPA} for a guide to the literature), there will \\underline{always} be regions in a neutron star where thermal effects are important (in the vicinity of the critical density at which the phase transition occurs). Understanding the nature of these transition regions, and their effect on various aspects of neutron star dynamics, is one of the main challenges for research in this area. In this paper, we describe a first attempt at modelling thermal dynamics in a superfluid neutron star. We focus on stars with a colour-flavour-locked (CFL) superconducting quark core \\cite{CFL} at finite temperatures. This is an interesting problem for several reasons. First of all, the simplest possible model for this system considers a quark condensate coupled to a gas of phonons. This problem is analogous to $^4$He at low temperatures, and hence we can bring our recent dissipative two-fluid model \\cite{helium} to bear on it (more or less directly). The lessons we learn from this exercise should inform the development of finite temperature models for the superfluids in the outer core and the neutron star crust. Secondly, even though there have been discussions of the dynamics of the different CFL phases, in particular in the context of the gravitational-wave driven r-mode instability (see \\cite{jaikumar} for references), the superfluid aspects have (as far as we are aware) not previously been accounted for. Our discussion begins to address the relevance of the additional degrees of freedom in these systems, and provides some insight into the nature of the different ``fluids'' involved. As an application with immediate astrophysical relevance, we will work out the inertial r-modes and estimate the relevant viscous damping rates for the ``simplest'' model of CFL matter. We consider the ``cool'' regime where the temperature is significantly below all the quasiparticle energy gaps. In this regime, dissipation may mainly occur due to phonon interactions. One reason for considering this model is that there are results in the literature for both bulk- and shear viscosity \\cite{manuel1,manuel2} as well as the mutual friction associated with superfluid vortices \\cite{mana1}. Of course, the simple ``condensate plus phonon'' model that we consider is not the whole story. It should apply at asymptotically high densities, but may not be the true ground state at lower densities. The discussion in \\cite{alford1,alford2,alford3} adds extra dimensions to the problem by considering the bulk viscosity due to kaons (allowing for flavour-changing processes), which will be present at higher temperatures. At first sight, this mechanism will only be relevant for very hot stars ($T>1$~MeV~$\\sim 10^{10}$~K) since it assumes that there is a thermal population of kaons. Below the critical energy where the kaons appear the contribution to the bulk viscosity is exponentially suppressed and may not be that important. However, in more recent work on the so-called CFL-K$^0$ phase \\cite{alford2}, it is argued that the main low temperature mechanism involves condensed kaons. This is important since the kaon condensate will remain present as $T \\to 0$. The upshot of this is that the problem requires a ``multi-fluid'' analysis at all temperatures. The simplest model would have three components; the quark condensate, the kaon condensate and finite temperature excitations (phonons and thermal kaons). This problem is more involved than the case that we focus on here. Nevertheless, our results provide an essential starting point for investigations into the dynamical role of the kaons. Most importantly, by providing the ``hydrodynamics'' view of the problem we illustrate the input needed to study various dynamical scenarios. This should stimulate further discussion between experts on different aspects of this multi-faceted problem, as required to make progress in the future. ", "conclusions": "We have presented the first true multi-fluid analysis of a dense neutron star core with a deconfined, colour-flavour-locked superconducting, quark core. By focussing on a cool system, and accounting only for the condensate and (finite temperature) phonons, we made progress by taking over much of the formalism from the analogous problem for superfluid $^4$He, the archetypal two-fluid laboratory system. The additional fluid degree of freedom, in the present case represented by the phonon gas, leads to the system not being well represented by the Navier-Stokes equations. In particular, a complete model requires a number of additional viscosity coefficients. Without an actual calculation it is not easy to establish whether the multi-fluid aspects are relevant or not. For example, in the case of the gravitational-wave driven instability of the f-modes it is known that the superfluid degree of freedom is very important, since the vortex mutual friction may completely suppress that instability below a critical temperature (see \\cite{lm1,fmode}). It is known that, because of the different nature of the associated velocity field, the mutual friction does not affect the r-mode instability in the same drastic fashion \\cite{rmode,lm}. These examples provide clear evidence that different problems need to be considered on a case-by-case basis. We have provided a detailed dissipative formulation for a system comprising a quark (CFL) condensate and phonons. The model builds on recent improvements in our understanding of the analogous problems of superfluid $^4$He \\cite{helium} and causal heat conductivity \\cite{heat}. A key ingredient is the massless entropy component that represents the phonon excitations. We have discussed how the superfluid constraint of irrotationality reduces the number of required viscosity coefficients to four (one shear and three bulk), and provided a translation between results in the literature \\cite{manuel1,manuel2} and the coefficients in our formalism. We also emphasised that many more dissipative channels may come into play in a rotating system where superfluid vortices are present \\cite{monster,helium}. In particular, we showed how the vortex mediated mutual friction is accounted for in the model, and translated the available results for the associated coefficients \\cite{mana1}. In order to be able to make relevant estimates for the r-mode instability, we developed a simple two-component equation of state based on the MIT bag model at zero temperatures with an additional phonon gas representing the thermal component. This example highlights the additional information that is needed in a multi-fluid analysis, in particular, regarding the entrainment coupling. Future work needs to provide a consistent equation of state, including \\underline{all} the key aspects. The model equation of state completed the formulation of the problem and we could, in principle, have carried out a numerical study of the r-modes. We opted not to do this, instead introducing a sequence of simplifying assumptions, because we felt that it would be useful to start by working out some less precise estimates for the relevant dissipation timescales. A numerical analysis of the problem should, of course, be encouraged. The problem has a number of interesting aspects, and may shed light on how one should deal with finite temperature superfluid neutron stars in general. This work was motivated by a desire to understand the different phases of CFL matter from a hydrodynamics point of view. It is, obviously, an interesting problem and the notion that observations of gravitational waves from relativistic stars may help shed light on the extreme QCD phase diagram is exciting. The r-mode instability has been discussed in this context for some time, see for example \\cite{manuel1,manuel2,mana1,alford1,alford2,madsen,strange}, but there has not been any previous discussion of the multi-fluid aspects of the problem. We hope that this work will stimulate a more detailed discussion between experts in the relevant areas. The fact that the various phonon processes that we have accounted for can be shown to have little effect on the r-mode instability should not discourage future efforts. In fact, the result could probably have been anticipated. The real challenge will be to account for additional degrees of freedom, especially associated with the kaons (either thermal or in a condensate) \\cite{alford1,alford2}, and the modelling of ``hybrid'' stars with quark cores and various phase transitions. These problems have additional features that, while possibly understood in principle, have never been considered in practice. This makes the modelling more complex, but also intriguing since the richer dynamics may lead to surprises." }, "1005/1005.3020_arXiv.txt": { "abstract": "Recent observations of excited CO emission lines from \\zsim 2 disc galaxies have shed light on the \\schmidt \\ relation at high-\\z \\ via observed \\sigmacojtwo \\ and \\sigmacojthree \\ relations. Here, we describe a novel methodology for utilising these observations of high-excitation CO to derive the underlying Schmidt (\\schmidt) relationship. To do this requires an understanding of the potential effects of differential CO excitation with SFR. If the most heavily star-forming galaxies have a larger fraction of their gas in highly excited CO states than the lower SFR galaxies, then the observed molecular Kennicutt-Schmidt index, $\\alpha$, will be less than the underlying \\schmidt \\ index, $N$. Utilising a combination of SPH models of galaxy evolution and molecular line radiative transfer, we present the first calculations of CO excitation in \\zsim 2 disc galaxies with the aim of developing a mapping between various observed \\sigmaco \\ relationships and the underlying \\schmidt relation. We find that even in relatively luminous \\zsim 2 discs, differential excitation does indeed exist, resulting in $\\alpha < N$ for highly excited CO lines. This means that an observed (e.g.) \\sigmacojthree \\ relation does not map linearly to a \\sigmahtwo \\ relation. We utilise our model results to provide a mapping from $\\alpha$ to $N$ for the range of Schmidt indices $N=1-2$. By comparing to recent observational surveys, we find that the observed \\sigmacojtwo \\ and \\sigmacojthree \\ relations suggest that an underlying SFR $\\propto \\rho^{1.5}$ relation describes \\zsim 2 disc galaxies. ", "introduction": "\\label{section:introduction} Since the original works by \\citet{sch59} and \\citet{ken98b} parameterising star formation rates (SFRs) in galaxies in terms of the scaling relation: \\begin{equation} \\label{equation:schmidt} SFR \\propto \\rho_{\\rm gas}^{\\rm N} \\end{equation} there have been considerable efforts by both the Galactic and extragalactic star formation communities to characterise the exponent $N$ \\citep[e.g. ][and references therein]{ken98a}. Constraining the Schmidt SFR relation in galaxies is desireable, both for understanding the physics of star formation on local scales, as well as for giving simulators recipes for modeling processes below typical numerical resolution scales. Because the volume density is not observable, most observed forms of Equation~\\ref{equation:schmidt} have been in terms of the SFR and gas surface densities\\footnote{To remain consistent with typical literature nomenclature, we will refer to the volumetric form of Equation~\\ref{equation:schmidt} as the Schmidt relation, and the surface density form as the Kennicutt-Schmidt (KS) relation. We will reserve the index, $N$, for the volumetric exponent, and $\\alpha$ for the surface density exponent.}. As a result of recent pioneering high resolution millimetre-wave surveys, two trends have become apparent. First, the SFR surface density in galaxies is well correlated with the {\\it molecular} gas surface density, though little relation exists between the SFR and HI atomic gas \\citep[e.g. ][]{big08,ler08,kru09}. Second, the {\\it global} integrated relationship between SFR and CO molecular gas surface densities in observations of local galaxies carries an exponent $N \\sim 1.5$ \\citep[e.g. ][]{san91,ken98b,gao04a,gao04b}. A variety of theories have been proposed to explain the $N \\sim 1.5$ index inferred from observations \\citep[e.g. ][]{sil97,tan00,elm02a,kru05,kru09b}, most of which rely on stars forming on a time scale proportional to the dynamical time scale. The recent advent of sensitive (sub)mm-wave interferometers has allowed, for the first time, observational constraints on the \\schmidt \\ relation in ``normal'' star-forming (e.g. not starbursting) galaxies at high redshift via the detection of CO rotational emission lines \\citep[][]{bou07,ion09,dad10a,dad10b,bot09,gen10,tac10}. However, owing to fact that most mm-wave interferometers operate in the 1-3 mm wavelength regime, most detections of CO emission lines at high-redshift are of relatively excited lines (e.g. CO (J=3-2), as opposed to the ground state transition which is typically used as a tracer of \\htwo \\ molecular gas. In principle, if the excitation of CO is relatively invariant in a given sample of galaxies, one can simply make an assumption regarding the CO (J=3-2) to (J=1-0) scaling ratio in high-\\z \\ galaxies, and derive an \\htwo \\ gas mass (or surface density) with the inferred CO (J=1-0) intensity. However, if the excitation of CO in galaxies is not constant with increasing SFR, then the exponent in the \\sfrco \\ relation may not naturally translate to an exponent in the \\schmidt \\ relation. In other words, $\\alpha$ does not always equal $N$ when probing high excitation CO lines. To see this, consider a sample of galaxies which are forming stars at a rate according to SFR $\\propto \\rho^{1.5}$. If the CO (J=1-0) line faithfully traces the \\htwo \\ gas mass, then one could expect an observed relationship $\\Sigma_{\\rm SFR} \\propto \\Sigma_{\\rm CO J=1-0}^{1.5}$ \\citep{kru07,nar08b}. However, if the galaxies with the highest SFR have a higher fraction of their CO gas excited into the CO J=3 state than the lowest SFR galaxies, one will observe a {\\it flatter} relationship between SFR and L$_{\\rm CO 3-2}$ than index $\\alpha=1.5$. The \\sigmacojthree \\ relation will only map linearly to the \\sigmahtwo \\ relation if the excitation of CO is invariant with SFR. In practice, this can only happen if the CO gas is thermalised in the observed lines (if a given excited state is in LTE). This effect of differential molecular excitation on observed molecular SFR scaling relations has been observed in the local Universe. The SFR-CO (J=1-0) relationship has an index $\\alpha \\approx 1.5$, while the SFR-CO (J=3-2) relationship has a flatter index $\\alpha \\approx 0.9$ \\citep{san91,yao03,nar05,ion09,bay09}. Similarly, while the SFR-HCN (J=1-0) relationship is linear in local galaxies, the SFR-HCN (J=3-2) index is decidedly sublinear \\citep[with index $\\alpha \\sim 0.7$; ][]{gao04a,gao04b,bus08,gra08,jun09}. Observations of individual star-forming clumps (which are massive enough to host stellar clusters) have been inconclusive regarding whether these global trends extend to smaller scales \\citep[][]{wu05,wu10}. In the absence of a more direct tracer of \\htwo \\ gas than high-excitation CO, the potential effects of differential molecular excitation with SFR need to be quantified in order to derive the underlying SFR relation in high-redshift galaxies. The few multi-line constraints of excitation in high-\\z \\ galaxies hints that CO may be subthermally excited even in the most luminous \\zsim 2 systems \\citep{wei07,dan09,car10,har10}. This indicates that applying a uniform mapping from (e.g.) CO (3-2) to CO (J=1-0) line intensities will indeed be problematic. In this arena, numerical models can offer guidance. Our aim in this paper is to calculate the mapping of observed \\sigmaco \\ relations of excited lines (e.g. CO J=2-1 and CO J=3-2) to an underlying \\schmidt \\ relation controlling the star formation. In \\citet{nar09,nar10b,nar10a} and \\citet{hay10}, we have developed a merger-driven model for the formation of high-redshift ULIRGs which shows reasonable correspondence with observed SEDs, CO emission properties and number counts (C. Hayward et al. in prep.). Here, we utilise the (idealised) progenitor disc galaxies of these model mergers to represent the star-forming discs at high-\\z \\ typically observed in CO emission line surveys \\citep[e.g. ][]{tac10,dad10b}. We combine these hydrodynamic simulations of disc galaxies with 3D non-LTE molecular line radiative transfer calculations in order to calculate the full statistical equilibrium excitation properties of the molecules. These methods allow us to determine the differential excitation of CO of \\zsim 2 disc galaxies with respect to SFR, and derive a mapping of an observed molecular Kennicutt-Schmidt law (\\sigmaco) to a \\schmidt \\ relationship. ", "conclusions": "\\begin{table*} \\centering \\caption{This Table provides the mapping between observed Kennicutt-Schmidt molecular surface-density indices, $\\alpha$, and underlying volumetric Schmidt indices, $N$. We provide the mapping for CO transitions J=1-0 through J=4-3. The numbers contained here constitute the information necessary to re-create a plot like Figure~\\ref{figure:ks_molslope} for any of the four modeled CO transitions and will aid interpretation of future observations of varying molecular transitions. The ``errors'' denote a 25\\% uncertainty level which encompasses variations in the final solution upon changing initial conditions in our simulation (primarily the equation of state and normalisation of the SFR relation).} \\begin{tabular}{@{}c||cccc@{}} \\hline Schmidt Index $N$ & SFR-CO (J=1-0)$^\\alpha$ & SFR-CO (J=2-1)$^\\alpha$ & SFR-CO (J=3-2)$^\\alpha$ & SFR-CO (J=4-3)$^\\alpha$ \\\\ \\hline 1 & 0.96$\\pm 0.24$ & 0.89$\\pm0.22$ &0.76$\\pm 0.19$ &0.49$\\pm 0.12$\\\\ 1.5& 1.47 $\\pm 0.37$ & 1.36$\\pm0.34$ &1.08$\\pm 0.27$ &0.77$\\pm 0.19$\\\\ 2& 2.13 $\\pm 0.53$ & 1.95$\\pm 0.48$ &1.45$\\pm 0.36$ &0.95$\\pm 0.24$\\\\ \\hline \\end{tabular} \\label{table:indices} \\end{table*} The principle result of this study is that, due to subthermal excitation in high-lying CO lines in high-\\z \\ galaxies, observed \\sigmaco \\ relations do not necessarily map linearly to \\sigmahtwo \\ relations. This is part of a broader theoretical framework first developed by \\citet{kru07} and \\citet{nar08b} which posits that the underlying density distribution of galaxies is crucial in mapping the underlying Schmidt relation to observed molecular Kennicutt-Schmidt relations. In this paper, we have expanded upon these studies by exploring the relationship between Schmidt and Kennicutt-Schmidt indices in models which aim to serve as reasonable representations of the \\zsim 2 star-forming disc galaxies being uncovered in sensitive optical/NIR observations \\citep[e.g. ][]{dad05,for09}. Indeed, the models studied here have been shown in previous publications to satisfy the star-forming {\\it BzK} colour-selection criteria \\citep{nar10b}, as well as serve as reasonable progenitors for luminous, merger-driven \\zsim 2 submillimetre and 24 \\micron-selected galaxies \\citep{hay10,nar10a}. We have additionally explored the dependence of the observed molecular Kennicutt-Schmidt index on varying underlying Schmidt relations. Our models, combined with the general theoretical framework of \\citet{kru07} and \\citet{nar08b} suggest that observed \\zsim 2 discs are subject to differential CO excitation with respect to SFR, and that the observed \\sigmacojtwo and \\sigmacojthree \\ relations may map to an underlying Schmidt index of $N=1.5$ controlling the SFR. An important verifying aspect to these models is that they are able to explain the multitude of observed SFR-\\lmol \\ relations in the local Universe. For example, a linear relationship has been observed between SFR and HCN (J=1-0) in local galaxies, a trend which has often been interpreted as evidence that a linear {\\it dense gas} Schmidt relation controlled the SFR. Our models suggest the linear SFR-HCN (J=1-0) relation is in reality a combined effect of an underlying \\schmidt \\ index of $N=1.5$ and differential excitation in HCN \\citep{kru07,nar08b}. This view has been observationally confirmed by \\citet{bus08}, who showed that local galaxies exhibit an {\\it sublinear} SFR-HCN (J=3-2) relation, and thus follow the trend of decreasing SFR-HCN$^\\alpha$ index with increasing transition number characteristic of these models \\citep[e.g. Figure~\\ref{figure:sfr_lmol}; see also Figure 7 of ][]{nar08b}. Similarly, this model satisfies the multi-line constraints of local CO observations. The SFR-CO (J=1-0) relation in local galaxies appears to have an index ranging from $\\sim 1.3-1.5$. At higher-lying transitions (e.g. CO J=3-2), the index drops to $\\sim 0.9$, in accordance with theoretical predictions \\citep{ion09}. We note that at lower bolometric luminosities, even the SFR-CO (J=1-0) relation may be subject to differential excitation and serve as a relatively poor tracer of the underlying Schmidt relation. Finally, with an eye toward ALMA, we comment on the role of spatial resolution in observational determinations of the KS relation at high-\\z. The exact mapping between the observed \\sigmaco \\ index, $\\alpha$, and the \\schmidt \\ index, $N$, depends on the level of thermalisation of the gas within the beam. It is dependent (to first order) on the mean gas density. Higher resolution observations which probe just the nucleus of the galaxy will probe higher mean densities, and allow even higher-lying CO emission lines to directly trace the underlying Schmidt SFR relation. Indeed, some very tentative observational evidence for this trend in local galaxies has been shown by \\citet{nar08d}. The trends shown in Figures~\\ref{figure:sfr_lmol} and \\ref{figure:ks_molslope} are for the central 6 kpc of the galaxy, comparable to the typical resolution of current interferometric observations of \\zsim 2 galaxies \\citep[e.g. ][]{tac10}. Our models suggest that observations of the central $\\sim$2 kpc will probe sufficiently dense gas that the observed \\sigmacojthree \\ index, $\\alpha$, will trace the underlying \\schmidt \\ index, $N$." }, "1005/1005.3811_arXiv.txt": { "abstract": "{We present wide-field {\\em Herschel}/PACS observations of A\\,1689, a massive galaxy cluster at $z{=}0.1832$, from our Open Time Key Programme. We detect 39 spectroscopically confirmed 100$\\mu$m-selected cluster members down to $1.5{\\times}10^{10}L_{\\odot}$. These galaxies are forming stars at rates in the range 1--10\\,M$_{\\odot}$/yr, and appear to comprise two distinct populations: two-thirds are unremarkable blue, late-type spirals found throughout the cluster; the remainder are dusty red sequence galaxies whose star formation is heavily obscured with A$({\\rm H}\\alpha){\\sim}2$\\,mag and are found only in the cluster outskirts. The specific-SFRs of these dusty red galaxies are lower than the blue late-types, suggesting that the former are in the process of being quenched, perhaps via pre-processing, the unobscured star formation being terminated first. We also detect an excess of 100$\\mu$m-selected galaxies extending ${\\sim}6$Mpc in length along an axis that runs NE-SW through the cluster center at ${\\ga}9$5\\% confidence. Qualitatively this structure is consistent with previous reports of substructure in X-ray, lensing, and near-infrared maps of this cluster, further supporting the view that this cluster is a dynamically active, merging system. } ", "introduction": "Over the last 30 years a large body of evidence has formed that star formation in cluster galaxies is much lower than in the surrounding field (e.g. Lewis et al. \\cite{lewis}). A variety of mechanisms have been proposed to quench star formation in infalling spirals such as ram-pressure stripping or harassment (see Haines et al. \\cite{haines07} for a review), but the key distinguishing evidence remains elusive. However most of these studies rely on optical spectroscopy using H$\\alpha$ or the [O{\\sc ii}] emission lines (e.g. Poggianti et al. \\cite{pog99}). Recent mid-infrared based studies with {\\em ISO} and then {\\em Spitzer} (e.g. Metcalfe et al. \\cite{metcalfe}; Coia et al. \\cite{coia}; Geach et al. \\cite{geach06}; Saintonge et al. \\cite{saintonge}) have revealed a significant population of dusty cluster galaxies whose star formation is heavily obscured at optical wavelengths, leading to [O{\\sc ii}]-based SFRs underestimating the true level by ${\\sim}1$0--3$0{\\times}$ (Duc et al. \\cite{duc}), while the age-dependence of the dust obscuration can lead to dusty star-bursts being in fact classified spectroscopically as post-starburst galaxies (Poggianti \\& Wu \\cite{poggianti}). Clusters are not isolated systems, but lie at the intersections of filaments, constantly accreting galaxies and galaxy groups along these filaments or directly from the field, and various studies suggest that the key sites of galaxy transformation are within these infalling structures rather than the cluster core (e.g. Balogh et al. \\cite{balogh04}; Moran et al. \\cite{moran}; Fadda et al. \\cite{fadda08}). These issues have motivated the Local Cluster Substructure (LoCuSS\\footnote{http://www.sr.bham.ac.uk/locuss}) \\emph{Herschel} Key Programme to study a large statistical sample of 30 massive galaxy clusters at $z{\\sim}0.2$ with panoramic FUV--FIR data from {\\em GALEX}, {\\em Spitzer} and {\\em Herschel} (Haines et al. \\cite{haines09a}; Smith et al. \\cite{smith10}). These provide a complete census of star formation in cluster galaxies extending to the infall regions, which can then be related to the dynamical status of the cluster as determined from complementary lensing, X-ray and dynamical analyses (Haines et al. \\cite{haines09b}; Pereira et al. \\cite{pereira}). \\begin{figure*} \\centering \\includegraphics*[height=101mm]{Herschelmap.eps}\\includegraphics[height=101mm]{A1689_merged.eps} \\caption{ ({\\em left:}) The {\\em Herschel}/PACS 100$\\mu$m map, with the spectroscopically confirmed cluster members indicated by blue open squares (scaling as 100$\\mu$m flux). Red contours indicate the projected $K$-band luminosity density of confirmed cluster members (weighted to account for spectroscopic incompleteness) estimated using the adaptive kernel method. ({\\em top-right}) Optical $g{-}i$ colour versus clustercentric radius. ({\\em bottom-right}) Redshift versus projected clustercentric radius. Blue, green and red symbols respectively indicate $K{\\le}K^{*}{+}2$ galaxies detected by: both {\\em Herschel} and {\\em Spitzer}; just {\\em Spitzer}; and neither telescope. Small black and grey points indicate low-mass ($K{>}K^{*}{+}2$) from ours and Czoske's spectroscopic surveys. } \\label{cmradius} \\label{phasespace} \\label{map} \\end{figure*} In this letter we present an analysis of the {\\em Herschel} (Pilbratt et al. \\cite{pilbratt}) FIR imaging of Abell 1689 at $z{=}0.1832$, one of the first clusters studied in the infrared with {\\em ISO} (Fadda et al. \\cite{fadda00}; Balogh et al. \\cite{balogh}, hereafter B02; Duc et al. \\cite{duc}, hereafter D02). Abell 1689 has many characteristics of a dynamically relaxed cluster, with a regular and highly symmetric X-ray morphology (Xue \\& Wu \\cite{xue}), the centroid of which coincides both with the brightest cluster galaxy and the centre of mass derived from strong lensing (Limousin et al. \\cite{limousin}). However, the significant discrepancies obtained between mass estimates derived from strong lensing and X-ray analyses (Andersson \\& Madejski \\cite{andersson}), the very high concentration value observed for the cluster, and its large ($\\theta_{E}{\\sim}45^{\\prime\\prime}$) Einstein radius, all strongly favour a triaxial model in which the cluster is aligned along the line-of-sight (Oguri et al. \\cite{oguri}; Corless, King \\& Clowe \\cite{corless}). Moreover, a substructure comprising ${\\sim}1$5\\% of the total cluster mass (Riemer-S{\\o}rensen et al. \\cite{riemer}) 1.5--2$^{\\prime}$ to the NE of the cluster centre has been identified from dynamical, X-ray and strong lensing analyses (Czoske \\cite{czoske}; Andersson \\& Madejski \\cite{andersson}; Limousin et al. \\cite{limousin}), while Kawaharada et al. (\\cite{kawaharada}, hereafter K10) find evidence for an accreting filament on 10--18\\,arcmin scales. We present the data in \\S2, results in \\S3, and summary and discussion in \\S4. Throughout we assume a cosmology with $\\Omega_{M}{=}0.3$, $\\Omega_{\\Lambda}{=}0.7$ and H$_{0}{=}70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$. ", "conclusions": "The results of the {\\em Herschel}/PACS photometry can be interpreted to suggest a filament of star-forming galaxies feeding the cluster A1689 at $z{=}0.1832$ from both sides. The direction of this filamentary structure coincides with the NE/SW axis of elongation of the cluster $K$-band light, and the observed asymmetry/substructure seen on 1--2\\,arcmin scales in previous X-ray and strong lensing analyses. K10 found anisotropies in gas temperature and entropy distributions in the cluster outskirts (10--18$^{\\prime}$), with a high $T_{X}{\\sim}5.4$\\,keV seen in the NE direction consistent with accretion flow along a filament, but low gas temperature (${\\sim}1.7$\\,keV) in all other directions. They also found evidence for a connecting large-scale structure (${\\sim}2$\\,deg) in the same NE direction in the form of SDSS galaxies with photometric redshifts consistent with the cluster. The 100$\\mu$m selection has made the filaments stand out clearly against the background and dense cluster region, in the same way as seen by Fadda et al. (\\cite{fadda08}) for 24$\\mu$m sources in Abell 1763, due to the enhanced star-forming activity among the filament galaxies. The SFRs of the {\\em Herschel} sources in A1689 are in the range 1--10\\,M$_{\\odot}{\\rm yr}^{-1}$, much lower than the 20--40\\,M$_{\\odot}{\\rm yr}^{-1}$ seen by Geach et al. (\\cite{geach06}, \\cite{geach}) for LIRGs in the rich cluster Cl\\,0024+16 at $z{=}0.4$. If this star formation is to build the bulges needed to turn the spirals infalling into clusters at $z{\\sim}0.4$ into the S0s found in local clusters it would require ${\\ga}$1--3\\,Gyr at the SFRs observed here and in other $z{\\sim}0.2$ clusters (Smith et al. \\cite{smith10}), rather than the ${\\sim}10^{8}$yr implied by Geach et al. (\\cite{geach})'s results. We find two populations of {\\em Herschel} sources in A1689, analogous to the blue cloud and dusty-red galaxies identified by Wolf et al. (\\cite{wolf05},\\cite{wolf09}) in A901/2, and we find similar EW(O{\\sc ii}), EW(H$\\delta$) and specific-SFRs to them. In particular, for the dusty red galaxies we obtain $\\langle$\\,EW(O{\\sc ii})$\\rangle{\\sim}5${\\AA}, due to having A(O{\\sc ii}$){\\sim}4$\\,mag placing them on the k+a/e(a) boundary, yet their specific-SFRs are systematically ${\\sim}3{\\times}$ lower than those of the blue cloud population (Fig.~\\ref{ssfr}), suggesting that they are currently being quenched. Whereas the ``blue cloud'' population are found throughout the cluster, the ``dusty red'' galaxies are found only outside the cluster core ($r{\\ga}1$\\,Mpc), where they make up half of the star-forming galaxies detected by {\\em Herschel} (Fig.~\\ref{cmradius}; top-right). This is consistent with the findings of STAGES that the dusty red galaxies prefer the intermediate-density environment of cluster infall regions (Gallazzi et al. \\cite{gallazzi}; Wolf et al. \\cite{wolf09}), and morphologically they found ${\\sim}70$\\% to be Sa/Sb spirals with a bright nucleus or inner bar/disk, as well as a few cases of interactions or merger remnants, suggestive of galaxy harassment, interactions or mergers for their origin. Four of the brightest eight {\\em Herschel} detections appear associated with groups: two are within a compact system of 3 galaxies ($\\sigma_{\\nu}{\\sim}120$\\,km\\,s$^{-1}$; labelled A in Fig.~\\ref{map}); one within a $\\sigma_{\\nu}{\\sim}320$\\,km\\,s$^{-1}$ group (16 members; group B); and the last within the substructure identified by Czoske (\\cite{czoske}; group C). These, plus similar results obtained by Pereira et al. (\\cite{pereira}) for Abell 1835, suggest that merger-induced starbursts and transformations in low-mass groups (pre-processing) could represent a significant contribution to the high-SFR cluster galaxy population, and an important evolutionary pathway for the formation of S0s. The availability of FUV--FIR imaging plus extensive spectroscopy for a sample of 30 clusters spanning the full range of morphologies and dynamical states within LoCuSS will allow us in future to estimate the relative fractions of infalling galaxies accreted by clusters from filaments, groups and the field, and measure the importance of pre-processing inside groups. \\begin{figure} \\centering \\includegraphics[width=68mm]{SSFR_hist.eps} \\caption{Histograms of specific-SFRs for 100$\\mu$m-detected cluster members within the red sequence (orange stripes) and blue cloud (blue).} \\label{ssfr} \\end{figure}" }, "1005/1005.0509_arXiv.txt": { "abstract": "{Star formation rate (SFR), metallicity and stellar mass are within the important parameters of star--forming (SF) galaxies that characterize their formation and evolution. They are known to be related to each other at low and high redshift in the mass--metallicity, mass--SFR, and metallicity--SFR relations.} % {In this work we demonstrate the existence of a plane in the 3D space defined by the axes SFR [log(SFR)(M$_{\\odot}$yr$^{-1}$)], gas metallicity [12+log(O/H)], and stellar mass [log(M$_{\\rm star}$/M$_{\\odot}$)] of SF galaxies.} {We used SF galaxies from the $``$main galaxy sample$\"$ of the Sloan Digital Sky Survey--Data Release 7 (SDSS --DR7) in the redshift range $0.04 < z < 0.1$ and $r-$magnitudes between 14.5 and 17.77. Metallicities, SFRs, and stellar masses were taken from the Max-Planck-Institute for Astrophysics--John Hopkins University (MPA-JHU) emission line analysis database.} {From a final sample of 32575 galaxies, we find for the first time a fundamental plane for field galaxies relating the SFR, gas metallicity, and stellar mass for SF galaxies in the local universe. One of the applications of this plane would be estimating stellar masses from SFR and metallicity. High redshift data from the literature at redshift $\\sim$0.85, 2.2, and 3.5, do not show evidence for evolution in this fundamental plane.} {} ", "introduction": "Relations between important properties of astrophysical objects often lead to the discovery of the so--called fundamental planes when three parameters are involved. The fundamental plane (FP) for elliptical galaxies \\citep{Djor87,Dressler87}, relates their luminosity, velocity dispersion (dynamics), and scale length (morphology). This FP represents an important tool to investigate the properties of early type and dwarf galaxies, to perform cosmological tests, and compute cosmological parameters. It is also an important diagnostic tool for galaxy evolution and mass--to--light ($M/L$) variations with redshift. Fundamental planes have also been defined for globular clusters \\citep{Meylan97} and galaxy clusters \\citep{Schaeffer93, Adami98}. The parameter space of globular clusters, elliptical galaxies, and galaxy clusters is properly described by a geometrical plane $L \\propto R^{\\alpha}\\sigma^{\\beta} $, where $L$ is the optical luminosity of the system, $R$ is a measure of the size scale, $\\sigma$ is the velocity dispersion of the system, and $\\alpha$ and $\\beta$ are free parameters. The FP for globular clusters, elliptical galaxies, and galaxy clusters have very similar slopes, which means that, accounting for differences in zero points, a single FP with a range of about nine orders of magnitude in luminosity can be defined \\citep{Schaeffer93,Ibarra09}. % The FP that we introduce here relates three fundamental parameters: the SFR [log(SFR)(M$_{\\odot}$yr$^{-1}$)], gas metallicity [12+log(O/H)], and stellar mass [log(M$_{\\rm star}$/M$_{\\odot}$)] of field SF galaxies. All these variables have been related in the past by the mass--metallicity ($M-Z$) relation \\citep{Lequeux79}, the mass--SFR relation \\citep{Brinchmann04}, and the metallicity--SFR relation \\citep{Lara10,Lopez-Sanchez10}. Also, some authors have studied the inter-dependence of those variables \\citep[e.g.][]{Hoopes07, Ellison08, Mannucci10}. In the present work, we propose the generalization of those relations defining a plane formed by a linear combination of two of those variables with respect to the third one. The $M-Z$ relates the mass and metallicity of galaxies, with massive galaxies showing higher metallicities than less massive ones, and it has been well established for the local universe ($z \\sim$ 0.1) by the work of \\citet{Tremonti04} using SDSS data. The $M-Z$ relation has also been studied at low redshifts $z \\sim 0.35$ \\citep{Lara09a,Lara09b}, at intermediate redshifts $z \\sim 0.7$ \\citep[e.g.,][]{Rodrigues08}, and at high redshift $z \\sim 2.2$ and $z \\sim 3.5$ \\citep[][respectively]{Erb06,Maiolino08}. The stellar mass of SF galaxies is also related to the SFR, in the sense that more massive galaxies show higher SFRs \\citep{Brinchmann04,Salim05}. However, \\citet{Brinchmann04} emphasized that at log(M$_{\\rm star}$/M$_{\\odot}$) $\\gtrsim$ 10, the distribution of SFRs broadens significantly and the correlation between stellar mass and SFR breaks down. At higher redshifts, \\citet{Noeske07} showed the existence of a $``$main sequence$\"$ (MS) for this relation over the redshift range $0.2 < z < 1.1$, with the slope of the MS moving as a whole as $z$ increases. The metallicity and SFR of SF galaxies are weakly correlated, as will be observed in Fig. 1. However, and despite of the high scatter, SFR increases with metallicity \\citep{Lara10,Lopez-Sanchez10}. This paper is structured as follows: Sect. 2 describes the data selection as well as the SFRs, metallicities and stellar masses estimations given by the Max-Planck-Institute for Astrophysics--John Hopkins University (MPA-JHU) group\\footnote{http://www.mpa-garching.mpg.de/SDSS} and adopted in this work. In Sect. 3 we define the FP for field galaxies, and conclusions are given in Sect. 4. ", "conclusions": "We have demonstrated the existence of a FP for field SF galaxies in the 3D space formed by the orthogonal coordinate axes log(M$_{\\rm star}$/ M$_{\\odot}$), log(SFR)(M$_{\\odot}\\rm yr^{-1}$), and 12+log(O/H); three of the fundamental parameters of galaxies. All those variables have been related previously in pairs as with the $M-Z$, metallicity--SFR, and mass-SFR relations, but this is the first time that the correlation for all of them has been quantified. The FP presented here allows estimating the stellar mass [log(M$_{\\rm star}$/ M$_{\\odot}$)] of field galaxies as a linear combination of 12+log(O/H) and log(SFR)(M$_{\\odot}\\rm yr^{-1}$). The scatter in the mass estimates using the FP (1$\\sigma$ error of 0.16) is lower that that obtained through the $M-Z$ and mass-SFR relations. The FP introduced here would be useful for deriving masses in spectroscopic surveys where the SFR and metallicity are estimated for emission line galaxies, for example, using the H$\\alpha$ luminosity to estimate the SFR \\citep[e.g.][]{Kennicutt98}, and any of the metallicity methods in the literature, such as the $R_{23}$ \\citep{Pagel79} or N2 \\citep{Denicolo02}, see \\citet{Kewley08} for a review. However, since this study has been carried out using emission line galaxies, this FP will be useful only when both, SFR and metallicity of galaxies can be estimated. Within the errors, there is no evidence of an evolution of the local FP when applied to high redshift samples. Which means that it could be useful even at high redshifts, where measuring the continuum and absorption lines for fitting models would be more difficult and time consuming. Then, we propose the use of this FP as an alternative tool to the existing methods to determine the stellar mass of galaxies at low and high redshifts." }, "1005/1005.5699_arXiv.txt": { "abstract": "{The standard cooling models of neutron stars predict temperatures of $T<10^{4}$~K for ages $t>10^{7}$~yr. However, the likely thermal emission detected from the millisecond pulsar J0437-4715, of spin-down age $t_s \\sim 7\\times10^9$~yr, implies a temperature $T\\sim 10^5$~K. Thus, a heating mechanism needs to be added to the cooling models in order to obtain agreement between theory and observation.} {Several internal heating mechanisms could be operating in neutron stars, such as magnetic field decay, dark matter accretion, crust cracking, superfluid vortex creep, and non-equilibrium reactions (``rotochemical heating''). We study these mechanisms to establish which could be the dominant source of thermal emission from old pulsars.} {We show by simple estimates that magnetic field decay, dark matter accretion, and crust cracking are unlikely\tto have a significant heating effect on old neutron stars. The thermal evolution for the other mechanisms is computed with the code of Fern\\'andez and Reisenegger. Given the dependence of the heating mechanisms on the spin-down parameters, we study the thermal evolution for two types of pulsars: young, slowly rotating ``classical'' pulsars and old, fast rotating millisecond pulsars.} {We find that magnetic field decay, dark matter accretion, and crust cracking do not produce any detectable heating of old pulsars. Rotochemical heating and vortex creep can be important both for classical pulsars and millisecond pulsars. More restrictive upper limits on the surface temperatures of classical pulsars could rule out vortex creep as the main source of thermal emission. Rotochemical heating in classical pulsars is driven by the chemical imbalance built up during their early spin-down, and is therefore strongly sensitive to their initial rotation period.} {} ", "introduction": "\\label{sec:intro} Neutron stars (NSs) are compact objects composed of a liquid core enveloped by a solid crust. The core is expected to contain superfluid neutrons and superconducting protons, while the crust contains heavy atomic nuclei arranged in a crystal lattice, coexisting with superfluid neutrons in its inner part. The high density of the NS core, up to $\\sim(3-9)\\rho_0$, where $\\rho_0$ is the saturation nuclear matter density, cannot be reproduced in terrestrial laboratories. This turns the NSs into natural laboratories. The study of their thermal evolution, confronting theory and observation, provides a useful test for the understanding of the properties of matter at supernuclear density. For all standard cooling models \\citep{yak04}, neutron stars cool down to surface temperatures $T_s < 10^4$~K within less than $10^7$~yr. Nevertheless, the observation of ultraviolet thermal emission from millisecond pulsar J0437-4715 \\citep{kar04}, whose spin-down age, corrected to the latest distance of $157$~pc \\citep{del08}, is $\\tau_{sd}\\sim 7\\times 10^9$~yr \\citep{van01}, shows a surface temperature of about $\\sim 10^5$~K for this pulsar. Hence, a heating mechanism needs be to added to the standard cooling models to obtain agreement between theory and observation. There are several heating mechanisms that can be present during the late stages of the thermal evolution. These include the frictional motion of superfluid neutron vortices \\citep{alp84,shi89,lar99}, rotochemical heating \\citep{reis95,reis97,fer05,petro09}, magnetic field decay \\citep{gol92,thom96,pon07}, and crust cracking \\citep{baym71,chen92}. Other mechanisms, based on more speculative hypotheses, such as a time variation of the gravitational constant \\citep{jof06}, the decay of exotic particles \\citep{han02}, or the accretion of dark matter particles \\citep{dela10,kouv10}, could in principle also heat old neutron stars. In \\citet{sch99} and \\citet{lar99}, several of the internal heating mechanisms cited are studied and confronted with observational data of neutron stars (surface temperature upper limits). However, the oversimplified description of some heating mechanisms \\citep[rotochemical heating in][]{sch99} and the until then non-detection of thermal emission from neutron stars older than $\\sim10^6$ yr, made it impossible to obtain reliable conclusions. The goal of this work is to provide a comparative analysis of the thermal evolution including different heating mechanisms. In order to do this, we discard some of them as not strong enough (magnetic field decay, dark matter accretion, and crust cracking), and we present a more detailed study of the most promising ones: vortex creep and rotochemical heating. We confront these mechanisms with the thermal emission detected in the millisecond pulsar J0437-4715 and the best available upper limits on the temperature of six other old pulsars. Owing to the dependence of heating mechanisms on spin-down parameters, which leads to different temperatures for different pulsars, we separately study the thermal evolution for two types of pulsars: young, strongly magnetized, and slowly rotating ``classical'' pulsars, and old, weakly magnetic, and fast rotating millisecond pulsars (MSPs). The paper is organized as follows. In Sect.~\\ref{sec:mech} we describe the different heating processes and assess their importance. In Sect.~\\ref{sec:eff} we show the effects of vortex creep and rotochemical heating in classical and millisecond pulsars. In Sect.~\\ref{sec:obs} the predictions are confronted with observed data. A summary of our main conclusions is given in Sect.~\\ref{sec:conc}. ", "conclusions": "\\label{sec:conc} We studied five heating mechanisms that can be operating in old neutron stars: magnetic field decay, dark matter accretion, crust cracking, vortex creep, and rotochemical heating, and compared them with pulsar observations. We found that magnetic field decay, dark matter accretion, and crust cracking cannot produce detectable heating in old pulsars. Owing to the high yield strain angle \\citep{hor09}, the crust cracking mechanism does not operate in classical pulsars, and probably only operates in MSPs. The vortex creep and rotochemical heating can be important both for classical pulsars and MSPs. In the evolutionary curves with vortex creep, and with the excess angular momentum, $J$, set to the lowest temperature compatible with the thermal emission detected from MSP J0437-4715, the predicted temperature turns out to be very near the observational upper limits for the classical pulsars B1929+10 and B1133+16. Likewise, rotochemical heating with modified Urca reactions and no Cooper pairing is only 1.7 $\\sigma$ below the temperature measured in the MSP J0437-4715. However, the temperature prediction of rotochemical heating can be raised if a superfluid core is considered in the model. The prediction of pinning energies in the inner crust of NSs via the semi-classical model for the vortex-nuclei interaction \\citep{don04} overestimates the temperatures of B0950+08 and B1929+10. The recent estimations of pinning energies via a quantum approach \\citep{avo08} are consistent with the observations of classical pulsars and MSPs. Finally, more stringent constraints on the temperature of some classical pulsars such as B0950+08 could rule out the vortex creep mechanism as the main source of the thermal emission detected in the MSP J0437-4715." }, "1005/1005.0844_arXiv.txt": { "abstract": "The growth of the supermassive black holes (BHs) that reside at the centres of most galaxies is intertwined with the physical processes that drive the formation of the galaxies themselves. The evolution of the relations between the mass of the BH, $\\mbh$, and the properties of its host therefore represent crucial aspects of the galaxy formation process. We use a cosmological simulation, as well as an analytical model, to investigate how and why the scaling relations for BHs evolve with cosmic time. We find that a simulation that reproduces the observed redshift zero relations between $\\mbh$ and the properties of its host galaxy, as well as the thermodynamic profiles of the intragroup medium, also reproduces the observed evolution in the ratio $\\mbh/\\ms$ for massive galaxies, although the evolution of the $\\mbh/\\sigma$ relation is in apparent conflict with observations. The simulation predicts that the relations between $\\mbh$ and the binding energies of both the galaxy and its dark matter halo do not evolve, while the ratio $\\mbh/\\mhalo$ increases with redshift. The simple, analytic model of \\citet{boot10}, in which the mass of the BH is controlled by the gravitational binding energy of its host halo, quantitatively reproduces the latter two results. Finally, we can explain the evolution in the relations between $\\mbh$ and the mass and binding energy of the stellar component of its host galaxy for massive galaxies ($\\ms\\sim 10^{11}\\,\\msun$) at low redshift ($z<1$) if these galaxies grow primarily through dry mergers. ", "introduction": "\\label{sec:intro} Over the past decade it has become clear that the supermassive black holes (BHs) found at the centres of virtually all galaxies with spheroidal components, have masses that are coupled to the properties of their host galaxies \\citep{magg98,ferr00,trem02,hari04,hopk07b}. Additionally, there exists evidence that BH masses are coupled to the properties of the dark matter haloes in which they reside \\citep{ferr02,boot10}. Further correlations between quasar activity \\citep[e.g.][]{boyl98} and the evolution of the cosmic star formation rate \\citep[e.g.][]{mada96} provide evidence that there exists a link between galactic star formation and accretion onto a central AGN. It has long been recognised that the growth of BHs is likely self-regulated \\citep{silk98} and that these tight correlations indicate that the growth of BHs is tightly intertwined with the physical processes that drive galaxy formation. However, despite a wide variety of theoretical and observational studies, the origin of these relations is still debated. The study of the evolution of the BH scaling relations therefore represents a crucial aspect of the galaxy formation process that may provide us with additional clues regarding the physical processes that give rise to the BH scaling relations. Addressing these questions observationally is challenging. Due to their extremely high luminosities, bright quasars provide a promising route to measuring BH masses at high redshift through the widths of low-ionization lines that are associated with the broad-line region close to the BH and using the assumption of virial equilibrium \\citep[e.g.][]{vest02}. It has, however, been claimed that this procedure systematically underestimates BH masses \\citep{jarv02}. Measuring galaxy masses for these objects is very difficult as the BH outshines the galaxy by a large factor \\citep[see e.g.\\ the discussion in ][]{merl09}. Since AGN surveys are biased towards more massive black holes, selection effects also need to be taken into account \\citep[e.g.][]{shen09,benn10}, which can make it difficult to distinguish between evolution in the normalization and in the scatter in the scaling relations \\citep{laue07}. In spite of these difficulties, measurements of the BH scaling relations have been made as far out as redshift three. \\citet{mclu06} found that the BHs associated with radio loud AGN residing in galaxies of a given stellar mass are a factor of four more massive at redshift two than in the local Universe. \\citet{deca09} studied the \\civ\\, line associated with the quasar broad line region in R-band selected hosts at both redshifts zero and three and found that BHs are typically a factor of seven more massive at high redshift for a given galaxy mass. These results are consistent with other observational studies \\citep{walt04,peng06,peng06a,merl09,gree10,benn10}. Taken together, these papers suggest an emerging consensus that at higher redshift BHs in hosts of a given mass are systematically more massive than in the local Universe, although see \\citet{jahn09} for one study that finds no significant evolution. The evolution of the relation between BH mass, $\\mbh$, and stellar velocity dispersion, $\\sigmas$, has been studied utilising the width of the \\oiii\\, line as a proxy for stellar velocity dispersion \\citep{nels96}. These studies suggest that the $\\mbh-\\sigmas$ relation either does not evolve \\citep{shie03,gask09}, or does so weakly, with BHs $\\sim 0.1-0.3$~dex more massive at $z=1$ \\citep{salv06,gu09,woo08,treu07}. The evolution of the BH scaling relations has also been studied using numerical simulations \\citep[e.g.][]{robe06,joha09} and semi-analytic models \\citep[e.g.][]{malb07,lama10,kisa10}. \\citet{robe06} employed simulations of idealised galaxy mergers, initialised to have properties typical of merger progenitors at various redshifts, to construct the relation between galaxy stellar mass, $\\ms$, and $\\sigmas$ as a function of redshift and found that, at a given value of $\\sigmas$, the corresponding $\\mbh$ decreases mildly with increasing redshift. At $z=1$ the simulations of \\citet{dima08} have BHs that lie slightly above the $z=0$ normalization of the $\\mbh-\\sigma$ relation. However, these simulations were stopped at $z=1$ and so cannot inform us about the evolution of the $\\mbh-\\sigmas$ toward lower redshift. However, for $z>1$ they predict a weak evolution in the $\\mbh-\\sigmas$ relation such that at higher redshift galaxies of a given velocity dispersion contain slightly less massive BHs. \\citet{joha09} employed similar numerical techniques to argue that it is unlikely that BHs are able to form significantly before their host bulges. Semi-analytic models that reproduce many redshift zero properties of galaxies also predict that, at a fixed $\\sigmas$, BH masses decrease with increasing redshift \\citep{malb07}. These theoretical models thus predict evolutionary trends that go in the opposite direction to those inferred from observations. Finally, the models of \\citet{hopk09} predict that, at a fixed stellar velocity dispersion, BH masses at higher redshift are either the same (for $\\mbh\\sim10^8\\,\\msun$) or slightly more massive (for $\\mbh>10^8\\,\\msun$) at fixed $\\sigmas$ than their redshift zero counterparts, in agreement with observation. On the other hand, the relation between BH mass and galaxy bulge mass shows a positive evolution in both semi-analytic models \\citep{malb07, hopk09} and numerical simulations \\citep{dima08}, the magnitude of which is comparable to that observed. The larger spread in the predictions for the evolution of the $\\mbh-\\sigmas$ relation may reflect that it is more difficult to predict velocity dispersions, which depend on both mass and size, than it is to predict masses. In \\citet[][hereafter BS09]{boot09} we presented self-consistent, hydrodynamical simulations of the co-evolution of the BH and galaxy populations that reproduce the redshift zero BH scaling relations. These same simulations also match group temperature, entropy and metallicity profiles, as well as the stellar masses and age distributions of brightest group galaxies \\citep{mcca10}. In \\citet[]{boot10} (hereafter BS10) we used the same simulations, as well as an analytic model, to demonstrate that $\\mbh$ is determined by the mass of the dark matter (DM) halo with a secondary dependence on the halo concentration, of the form that would be expected if the halo binding energy were the fundamental property that controls the mass of the BH. In the present work we use the same models to investigate why and how the BH scaling relations evolve for massive galaxies. This paper is organised as follows. In Sec.~\\ref{sec:method} we summarise the numerical methods employed in this study and the simulation analysed. In Sec.~\\ref{sec:ev} we present predictions for the evolution of the BH scaling relations and compare them to observations. We find that the evolution in the $\\mbh-\\ms$ relation predicted by the simulations is in excellent agreement with the observations, while the measured weak evolution in the $\\mbh-\\sigma$ relation is in apparent disagreement, and predict that while BH mass increases with redshift for fixed halo mass, the relations between $\\mbh$ and the binding energies of both the host galaxies and DM haloes do not evolve. We demonstrate in \\ref{sec:explanation} that the a analytic description in which $\\mbh$ is coupled to the DM halo binding energy can reproduce the evolution of the relation between BH and halo mass. Furthermore, we show that the evolution in the relations between the BH and the stellar mass and binding energy can be understood in terms of the more fundamental relation with the binding energy of the dark halo and the growth of massive galaxies through dry mergers. Finally, we summarise our main conclusions in Sec.~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} We have used a self-consistent cosmological simulation that reproduces the observed redshift zero relations between $\\mbh$ and both galaxy and halo properties (as well as the thermodynamic profiles of the intragroup medium) to investigate how, and why, these relations evolve through time. The relation between BH mass and host galaxy mass predicted by the simulation is consistent with available observations at $z<2$, which are currently confined to $\\ms\\sim 10^{11}\\,\\msun$ galaxies. For such galaxies we predict that the ratio $\\mbh/\\ms \\propto (1+z)^{\\alphas}$, with $\\alphas \\approx 0.5$, and $\\mbh/\\sigmas^4 \\propto (1+z)^{\\alphas}$ with $\\alphas\\approx-0.3$, in apparent conflict with recent observations. The ratio between the BH mass and the binding energy of the dark halo is independent of redshift, in agreement with BS10 who argued that the BH mass is controlled by the halo binding energy. The simple analytic model of BS10, in which the BH mass is assumed to scale in proportion to the binding energy of the dark halo, not only reproduces the simulated redshift zero $\\mbh-\\mhalo$ relation, but also its evolution. For a fixed halo mass BHs are more massive at higher redshift because the haloes are more compact and thus more tightly bound. Assuming an NFW halo density profile and the evolution of the halo concentration-mass relation predicted by simulations, the model can quantitatively account for the predicted evolution. The simulation predicts that the ratio between the BH mass and the binding energy of the stellar component of the galaxy is also independent of redshift (at least for $\\ms\\sim10^{11}\\,\\msun$ and $z<1$), even though BS10 demonstrated explicitly that the correlations between BH mass and stellar properties are not fundamental. This result is, however, consistent with a picture in which massive galaxies grow primarily through dry mergers at low redshift, which we showed to be the case in the simulation. Combined with the observed evolution in the $\\ms-\\sigmas$ relation, this idea can quantitatively account for the evolution in the $\\mbh-\\ms$ relation. One interesting implication of this scenario is that the evolution of the relations between BHs and the properties of their host galaxies may differ for galaxies that do not grow predominantly through dry mergers, as would be expected for lower masses and at higher redshifts. We will investigate this further in the a future work, employing higher resolution simulations." }, "1005/1005.2888_arXiv.txt": { "abstract": "Evidence is presented indicating that in the hard state of Cygnus X-1, the coronal magnetic field might be below equipartition with radiation (suggesting that the corona is not powered by magnetic field dissipation) and that the ion temperature in the corona is significantly lower than what predicted by ADAF like models. It is also shown that the current estimates of the jet power set interesting contraints on the jet velocity (which is at least mildly relativistic), the accretion efficiency (which is large in both spectral states), and the nature of the X-ray emitting region (which is unlikely to be the jet). ", "introduction": "\\label{sec:intro} Black hole binaries are observed in two main X-ray spectral states, namely the Hard State (HS) and the Soft State (SS), see [1]. The hard X-ray emission in both spectral states is well represented by Comptonisation by an hybrid thermal/non-thermal electron distribution. In the HS the temperature and optical depth of the thermal electrons are higher, and the slope of the non-thermal tail seem steeper than in the SS. Consequently, the hard X-ray emission is dominated by thermal Comptonisation in the HS and by non-thermal Comptonisation in the SS. The HS is known to be associated with the presence of a compact radio jet which observed in the SS. Here we focus on the prototypical black hole source Cygnus X-1. In section~\\ref{sec:model}, we present a relatively simple coupled kinetic-radiation model that allows us to understand the origin of the very different spectral shapes observed in the two spectral states as well as the spectral evolution during state transitions (see e.g. [2]). A thorough investigation of the model and its discussion in the context of the observations can be found in [3]; the present paper summarises our main results. Then in section~\\ref{sec:jet}, we summarise the arguments developed in [4] showing that the present estimates of the jet power of Cygnus X-1 imply that the jet has a relativistic velocity, that the accretion proceeds efficiently in the HS and that the X-ray emission is unlikely to be produced in the jet. ", "conclusions": "In both spectral states of black hole binaries the coronal emission can be powered by a similar non-thermal acceleration mechanism. In the HS the synchrotron and $e$-$e$ Coulomb boilers redistribute the energy of the non-thermal particles to form and keep a quasi-thermal electron distribution at a relatively high temperature, so that most of the luminosity is released through quasi-thermal Comptonisation. In the SS, the soft photon flux from the accretion disc becomes very strong and cools down the electrons, reducing the thermal Compton emissivity. This change in the soft photon flux could be caused either because the inner radius of the truncated disc moves inward into the central hot accretion flow, or, in the framework of accretion disc corona models, because the disc temperature increases dramatically. Then most of the emission is produced by disc photons up-scattered by the non-thermal cooling electrons. Our comparison of simulations with the high energy spectra of Cygnus X-1 in the HS allowed us to set upper limits on the magnetic field and the proton temperature. Our results indicate that, { as long as the non-thermal MeV tail is produced in the same region as the bulk of the Comptonised emission}, the magnetic field in the HS is below equipatition with radiation (unlike what is assumed in most accretion disc corona models). The proton temperature is found to be significantly lower than predicted by standard 2-temperature accretion flow models ($kT_{\\rm i}<$2 MeV). We also note that such accretion flows are usually radiatively inneficient while the jet energetics suggests efficient accretion in the HS. The present estimates of the jet power also suggest that the jet is Thomson thin with a velocity which is {at least} mildly relativistic, which does not make it a favoured location for the production of the observed X-ray emission." }, "1005/1005.3076_arXiv.txt": { "abstract": "{ In \\textit{Herschel} images of the Galactic plane and many star forming regions, a major factor limiting our ability to extract faint compact sources is cirrus confusion noise, operationally defined as the ``statistical error to be expected in photometric measurements due to confusion in a background of fluctuating surface brightness.'' The histogram of the flux densities of extracted sources shows a distinctive faint-end cutoff below which the catalog suffers from incompleteness and the flux densities become unreliable. This empirical cutoff should be closely related to the estimated cirrus noise and we show that this is the case. We compute the cirrus noise directly, both on \\textit{Herschel} images from which the bright sources have been removed and on simulated images of cirrus with statistically similar fluctuations. We connect these direct estimates with those from power spectrum analysis, which has been used extensively to predict the cirrus noise and provides insight into how it depends on various statistical properties and photometric operational parameters. We report multi-wavelength power spectra of diffuse Galactic dust emission from Hi-GAL observations at 70 to 500\\,$\\mu m$ within Galactic plane fields at $l= 30 \\degr$ and $l= 59 \\degr$. We find that the exponent of the power spectrum is about $-3$. At $250\\ \\mu{\\rm m}$, the amplitude of the power spectrum increases roughly as the square of the median brightness of the map and so the expected cirrus noise scales linearly with the median brightness. For a given region, the wavelength dependence of the amplitude can be described by the square of the spectral energy distribution (SED) of the dust emission. Generally, the confusion noise will be a worse problem at longer wavelengths, because of the combination of lower angular resolution and the rising power spectrum of cirrus toward lower spatial frequencies, but the photometric signal to noise will also depend on the relative SED of the source compared to the cirrus.} ", "introduction": "Cirrus noise, which is operationally defined as the ``statistical error to be expected in photometric measurements due to confusion in a background of fluctuating surface brightness'' \\citep{gautier}, is a major issue limiting the cataloging of compact sources that underpins the study of the early stages of star formation in the interstellar medium. Examination of wide range of mass of the stellar precursors, requires measurement of sources with a wide range of luminosity, or at each wavelength, flux density. Stars form where there is abundant material, and so the cirrus brightness in the field tends to be high. Furthermore, many studies, both targeted and unbiased, are in the Galactic plane, which is also bright. Cirrus noise varies with cirrus brightness (Sect.~\\ref{sect:cnps}), introducing further complexity to the problem. Cirrus fluctuations characteristically decrease with decreasing spatial scale (Sect.~\\ref{sect:cnps}), but even with the improved angular resolution of \\textit{Herschel}, cirrus noise remains a dominant factor. The \\textit{Herschel} observation planning tool HSpot (www.ipac.caltech.edu/Herschel/hspot.shtml) has a built-in confusion noise estimator to provide ``on-line guidance on where to expect fundamental detection limits for point sources that cannot be improved by increasing the integration time.'' With such guidance the serious impact of cirrus noise was anticipated, and as shown below it can now be quantified directly using submillimeter data. To appreciate the problem at its fundamental level, consider actual catalogs of extracted sources \\citep{MolinariThisVolume,mol2010detect} in two degree-sized Hi-GAL Galactic plane fields \\citep{molinari2010} (see Sect.~\\ref{sect:obs}). \\begin{figure}% \\centering \\includegraphics[width=0.4\\textwidth, angle=0]{14684fg1.ps} \\caption{ Solid histogram: 136 sources cataloged at 250\\,$\\mu m$ in a representative large degree-sized sub-region of Hi-GAL $l= 30\\degr$ field (f30: see Sect.~\\ref{sect:obs}). Dashed histogram: 70 sources cataloged in a sub-region of Hi-GAL $l= 59\\degr$ field (f59) that has a factor 5.8 lower median brightness.} \\label{fig:hist2} \\end{figure} The solid histogram in Figure~\\ref{fig:hist2} shows the flux densities of sources cataloged at 250\\,$\\mu m$ in the brighter field. Note the falloff in source counts at flux densities less than 10~Jy. This falloff is not an intrinsic property of the underlying population of sources. We show below that this falloff is the expected consequence of cirrus noise, which can be quantified independently of the making of the catalogs. In the fainter field (dashed histogram), the cutoff is lower because of decreased cirrus noise. Clearly, it will be important to account for the varying dramatic effects of cirrus noise in order to recover the statistics of the intrinsic faint-source population.% Cirrus noise is operationally defined and so for a particular measuring strategy, such as fitting compact sources with Gaussians as used in Hi-GAL, this can actually be estimated directly from the source-removed maps (Sect.~\\ref{sect:cnps}). There is also an extensive literature on ``estimating confusion noise due to extended structures given some statistical properties of the sky'' \\citep{gautier}; see \\citep{kiss-galcir,jeong2005,ma-dust,roy2010}. The cirrus brightness statistics are well described by a power spectrum, appropriate for Gaussian random fields. \\citet{gautier} found some non-Gaussianity and \\citet{ma-dust} found non-vanishing skewness and excess kurtosis in the underlying brightness fluctuation fields. Nevertheless, these are not large effects and for estimating the variance the power spectrum is demonstrably still a powerful tool, particularly because of the insight it provides into how the cirrus noise depends on various statistical and operational parameters. Therefore, we connect the new direct estimates with what can be obtained using power spectra (Sect.~\\ref{sect:cnps}). In Sect.~\\ref{sect:prop} we show that reliable power spectra can be obtained even from first-generation images processed for Hi-GAL. Finally, we return to how estimates of cirrus noise should ultimately explain the faint-end cutoff in forthcoming source catalogs (Sect.~\\ref{sect:noise}). ", "conclusions": "" }, "1005/1005.1073_arXiv.txt": { "abstract": "{ The constituents of the cosmic IR background (CIB) are studied at its peak wavelengths (100 and 160 $\\mu$m) by exploiting Herschel/PACS observations of the GOODS-N, Lockman Hole, and COSMOS fields in the {\\em PACS Evolutionary Probe} (PEP) guaranteed-time survey. The GOODS-N data reach 3$\\sigma$ depths of $\\sim$3.0 mJy at 100 $\\mu$m and $\\sim$5.7 mJy at 160 $\\mu$m. At these levels, source densities are 40 and 18 beams/source, respectively, thus hitting the confusion limit at 160$\\mu$m. Differential number counts extend from a few mJy up to 100-200 mJy, and are approximated as a double power law, with the break lying between 5 and 10 mJy. The available ancillary information allows us to split number counts into redshift bins. At $z\\le0.5$ we isolate a class of luminous sources ($L_{IR}\\sim10^{11}$ L$_\\odot$), whose SEDs resemble late-spiral galaxies, peaking at $\\sim$130 $\\mu$m restframe and significantly colder than what is expected on the basis of pre-Herschel models. By integrating number counts over the whole covered flux range, we obtain a surface brightness of $6.36\\pm1.67$ and $6.58\\pm1.62$ $[$nW m$^{-2}$ sr$^{-1}]$ at 100 and 160 $\\mu$m, resolving $\\sim$45\\% and $\\sim$52\\% of the CIB, respectively. When stacking 24 $\\mu$m sources, the inferred CIB lies within 1.1$\\sigma$ and 0.5$\\sigma$ from direct measurements in the two bands, and fractions increase to 50\\% and 75\\%. Most of this resolved CIB fraction was radiated at $z\\le1.0$, with 160 $\\mu$m sources found at higher redshift than 100 $\\mu$m ones. } ", "introduction": "The cosmic IR background \\citep[CIB,][]{puget1996,hauser1998} accounts for roughly half of the total extragalactic background light \\citep[EBL,][]{hauser2001,lagache2005}, i.e., half of the energy radiated by all galaxies, at all cosmic epochs, at any wavelength \\citep{dole2006}. It is therefore a crucial constraint on modes and times of galaxy formation. Deep cosmological surveys carried out with the {\\em Infrared Space Observatory} (ISO, see Genzel \\& Cesarsky \\citeyear{genzel2000} for a summary) and {\\em Spitzer Space Telescope} \\citep[][for a review]{soifer2008} produced large samples of mid-IR sources and deep number counts \\citep{elbaz2002,papovich2004}. These surveys led to mid-IR CIB lower limits within a factor of two from the upper constraints set by TeV cosmic opacity measurements \\citep[e.g.][]{franceschini2009}. However, at CIB peak wavelengths ($100-200$ $\\mu$m), the nature of individual galaxies building up the EBL is poorly known. Past surveys produced limited samples of distant far-IR objects \\citep[e.g.][]{frayer2009}, mainly due to the small apertures of the available instruments and the low sensitivity in the far-IR. In the 160 $\\mu$m Spitzer/MIPS band, $\\sim$7\\% of the CIB was resolved into individually detected objects \\citep{dole2004}, and it was only through stacking the 24 $\\mu$m sources that most ($60-70$\\%) of the far-IR CIB could be recovered \\citep{dole2006,bethermin2010}. Similarly, at longer wavelengths, only stacking of 24 $\\mu$m sources on BLAST maps could account for the majority of the EBL at 250, 350, and 500 $\\mu$m \\citep{marsden2009}. With the favorable diffraction limit of the large {\\em Herschel} 3.5 m mirror (Pilbratt et al., \\citeyear{pilbratt2010}), and the high sensitivity of its {\\em Photodetector Array Camera \\& Spectrometer} (PACS, 70, 100, 160 $\\mu$m; Poglitsch et al., \\citeyear{poglitsch2010}), confusion and blending of sources are much less of a limitation. We are now able to resolve a large fraction of the CIB at its peak into individual galaxies. The {\\em PACS Evolutionary Probe} (PEP) extragalactic survey samples 4 different tiers: from the wide and shallow COSMOS field, through medium size areas like the Lockman Hole, all the way down to the 160 and 100 $\\mu$m confusion limit in the pencil-beam, very deep observations in GOODS-N and GOODS-S, and even beyond by exploiting gravitational lensing in low-redshift galaxy clusters (e.g. Abell 2218, Altieri et al. \\citeyear{altieri2010}). \\begin{figure*}[!ht] \\centering \\rotatebox{-90}{ \\includegraphics[height=0.425\\textwidth]{14610f1a.eps} } \\rotatebox{-90}{ \\includegraphics[height=0.425\\textwidth]{14610f1b.eps} } \\caption{Number counts at 100 ({\\em left}) and 160 $\\mu$m ({\\em right}), normalized to the Euclidean slope. Filled/open symbols belong to flux bins above/below the 80\\% completeness limit. Models belong to \\citet{lagache2004}, \\citet{franceschini2009}, \\citet{rowanrobinson2009}, \\citet{leborgne2009}, \\citet{valiante2009}, \\citet{lacey2009}, Gruppioni \\& Pozzi (in prep.). Shaded areas represent ISO and Spitzer data \\citep{rodighiero2004,heraudeau2004,bethermin2010}. {\\em Inset}: Collection of PACS number counts, including this work (red, limited to relative errors $\\le$15\\% for clarity sake), PEP Abell 2218 (grey, Altieri et al. \\citeyear{altieri2010}) and HerMES-PACS (light-blue, Aussel et al. \\citeyear{aussel2010}). The solid lines in the insets mark the trends expected for a non-evolving population of galaxies.} \\label{fig:pep_counts} \\end{figure*} Here we exploit the {\\em science demonstration phase} (SDP) observations of the GOODS-N field, complemented with the COSMOS and Lockman Hole wider and shallower layers of the survey, to build far-IR galaxy number counts and derive the fraction of CIB resolved by PEP. We then take advantage of the extensive multi-wavelength coverage of the GOODS-N field, with the aim of identifying the CIB contributors and the epoch when the bulk of the Universe IR energy budget was emitted. ", "conclusions": "\\label{sect:discussion} \\begin{figure*}[!ht] \\centering \\includegraphics[height=0.4\\textwidth]{14610f3a.eps} \\includegraphics[height=0.4\\textwidth]{14610f3b.eps} \\caption{Slicing the PEP GOODS-N population into redshift bins. Each PACS source has been associated to an optical/mid-IR counterpart as described in Appendix \\ref{sect:data}. See Fig. \\ref{fig:pep_counts} and text for details on the models shown.} \\label{fig:pep_counts_redshift} \\end{figure*} In the attempt to reproduce the observed ISO and Spitzer number counts, several authors built ``backward'' evolutionary models, including luminosity and/or density evolution, as well as different galaxy populations. In Fig. \\ref{fig:pep_counts}, we overlay recent models onto the observed PACS counts. We include in this collection also the \\citet{lacey2009} $\\Lambda$-CDM semi-analytical model (SAM), complemented with radiative transfer dust reprocessing. The most successful models are the \\citet{valiante2009}, including luminosity-dependent distribution functions for the galaxy IR SEDs and their AGN contribution, and \\citet{rowanrobinson2009}, employing analytic evolutionary functions without discontinuities and 4 galaxy populations. Gruppioni and \\citet{lagache2004} overestimate the amplitude of the number counts peak in both bands, while \\citet{franceschini2009} and \\citet{leborgne2009} reproduce the counts fairly well only in one channel (100 and 160 $\\mu$m, respectively). It is worth recalling that to date most of these models have been fine-tuned to reproduce mid-IR and sub-mm statistics, while a big gap in wavelength was affecting far-IR predictions. Most include a luminosity evolution $\\propto(1+z)^{3.0-3.5}$, but the redshift limit for this slope, the details of density evolution, or the adopted galaxy zoo vary significantly from author to author. Besides spanning a much wider range of observational data (UV, optical, near-IR luminosity functions, galaxy sizes, metallicity, etc.), the SAM approach suffers here for a limited flexibility in the choice of parameters, and significantly overestimates the bright end of PACS counts. Moreover, it cannot reproduce the peak, especially in the green band. Thanks to the rich ancillary dataset in GOODS-N, we split the far-IR number counts into redshift bins (Fig. \\ref{fig:pep_counts_redshift}). This elaboration offers a remarkable chance to set detailed constraints on the evolution of the galaxy populations adopted in current recipes. This view highlights some new features and the main problems of the models under discussion. First, in the lowest redshift bin, $0.02$. Most of the resolved 160 $\\mu$m CIB is produced at higher redshift, with the $0.51$ increases by more than 10\\%. On the other hand, at 160 $\\mu$m only the relative fraction of the highest redshift bin varies by more than 5\\%, while the others remain unchanged within $1-2$\\%. Also in the case of stacking, the 160 $\\mu$m CIB relative redshift distribution is more populated at high redshift than the 100 $\\mu$m one. As expected, this indicates that PACS galaxies with redder observed colors lie on average at higher redshift than bluer ones. \\onltab{1}{ \\begin{table*}[!ht] \\centering \\begin{tabular}{r | c c | c c | c c | c c | c c | c c} \\hline \\hline \\multicolumn{1}{c|}{$S_{center}$} & \\multicolumn{2}{c|}{GOODS-N 100$\\mu$m} & \\multicolumn{2}{c|}{LH 100$\\mu$m} & \\multicolumn{2}{c|}{COSMOS 100$\\mu$m} & \\multicolumn{2}{c|}{GOODS-N 160$\\mu$m} & \\multicolumn{2}{c|}{LH 160$\\mu$m} & \\multicolumn{2}{c}{COSMOS 160$\\mu$m}\\\\ \\hline 2.84 & 3.19e+04 & 0.208 & -- \t & -- & \t --\t & --\t& --\t & --\t& \t--\t&\t-- &\t --\t &\t-- \\\\ 3.57 & 4.67e+04 & 0.146 & -- \t & -- & \t --\t & --\t& --\t & --\t& \t--\t&\t-- &\t --\t &\t-- \\\\ 4.50 & 5.57e+04 & 0.149 & -- \t & -- & \t --\t & --\t& --\t & --\t& \t--\t&\t-- &\t --\t &\t-- \\\\ 5.66 & 6.37e+04 & 0.162 & 7.00e+04\t & 0.086 & \t --\t & --\t& 1.17e+05 & 0.129 & \t--\t&\t-- &\t --\t &\t-- \\\\ 7.13 & 6.90e+04 & 0.185 & 7.22e+04\t & 0.100 & \t 7.23e+04 & 0.028\t& 1.42e+05 &\t 0.173 & \t--\t&\t-- &\t --\t &\t-- \\\\ 8.97 & 6.43e+04 & 0.234 & 6.58e+04\t & 0.128 & \t 8.10e+04 & 0.031\t& 1.63e+05 &\t 0.142 & 1.43e+05\t& 0.086 &\t --\t &\t-- \\\\ 11.30 & 6.29e+04 & 0.287 & 7.24e+04\t & 0.146 & \t 7.47e+04 & 0.039\t& 1.74e+05 &\t 0.163 & 1.67e+05\t& 0.093 &\t --\t &\t-- \\\\ 14.22 & 6.81e+04 & 0.333 & 8.13e+04\t & 0.167 & \t 6.81e+04 & 0.050\t& 1.66e+05 &\t 0.200 & 1.82e+05\t& 0.106 &\t --\t &\t-- \\\\ 17.91 & 1.11e+05 & 0.302 & 7.00e+04\t & 0.217 & \t 7.33e+04 & 0.057\t& 1.77e+05 &\t 0.237 & 1.70e+05\t& 0.132 &\t 2.03e+05 &\t0.033\\\\ 22.54 & 8.13e+04 & 0.443 & 6.66e+04\t & 0.269 & \t 7.15e+04 & 0.071\t& 1.97e+05 &\t 0.266 & 1.53e+05\t& 0.171 &\t 2.07e+05 &\t0.039\\\\ 28.38 & 1.11e+05 & 0.438 & 6.70e+04\t & 0.317 & \t 6.70e+04 & 0.087\t& 1.83e+05 &\t 0.334 & 1.85e+05\t& 0.183 &\t 1.95e+05 &\t0.049\\\\ 35.73 & 6.63e+04 & 0.701 & 4.01e+04\t & 0.529 & \t 5.74e+04 & 0.116\t& 1.88e+05 &\t 0.394 & 1.66e+05\t& 0.238 &\t 1.80e+05 &\t0.062\\\\ 44.98 & 6.68e+04 & 0.814 & 5.97e+04\t & 0.484 & \t 5.91e+04 & 0.135\t& 1.49e+05 &\t 0.527 & 1.42e+05\t& 0.312 &\t 1.48e+05 &\t0.082\\\\ 56.62 & --\t & -- & $\\le$1.81e+04\t & 1.000 & \t 5.06e+04 & 0.175\t& 1.44e+05 &\t 0.641 & 1.14e+05\t& 0.415 &\t 1.41e+05 &\t0.100\\\\ 71.29 & --\t & -- &\t $\\le$2.40e+04\t & 1.000 & \t 5.64e+04 & 0.198\t& 1.01e+05 &\t 0.883 & 7.01e+04\t& 0.658 &\t 1.16e+05 &\t0.135\\\\ 89.74 & --\t & -- &\t 8.46e+04\t & 0.655 & \t 3.89e+04 & 0.289\t& -- &\t -- & 6.37e+04\t& 0.828 &\t 9.19e+04 &\t0.185\\\\ 112.98 & --\t & -- &\t\t--\t & --\t & \t 2.51e+04 & 0.453\t& --\t & --\t& 8.19e+04\t& 0.833 &\t 7.70e+04 &\t0.239\\\\ 142.23 & --\t & -- &\t\t--\t & --\t & \t 3.81e+04 & 0.412\t& --\t & --\t& --\t& -- &\t 8.38e+04 &\t0.271\\\\ 179.06 & --\t & -- &\t\t--\t & --\t & \t --\t & --\t& --\t & --\t& --\t& -- &\t 6.57e+04 &\t0.366\\\\ \\hline \\end{tabular} \\tablefoot{Fluxes are provided in $[$mJy$]$. For each field/wavelength, we list counts in units of $[$deg$^{-2}$ mJy$^{1.5}]$. Errors are given as relative fractions, and include both Poisson statistics and propagation of photometric uncertainties.} \\caption{PEP number counts, normalized to the Euclidean slope.} \\label{tab:pep_counts} \\end{table*} }" }, "1005/1005.0149_arXiv.txt": { "abstract": "{} { An optical search is carried out for supernova remnants (SNRs) in the Sc type nearby spiral galaxy M74, using ground-based observations at TUBITAK National Observatory (TUG, Antalya/Turkey) and Special Astrophysics Observatory (SAO, Russia). Observations are supplemented by the spectral analysis of archived X-ray data from $XMM$-$Newton$\\ and $Chandra$.} { A survey of M74 covering $\\sim$ 9 arcmin$^{2}$ with [S II], H$\\alpha$, and their continuum filters. Interference filter images of M74 are obtained with 1.5 m Russian Turkish Telescope (RTT150) at TUG and spectral data are taken with the 6 m Bolsoi Azimuthal Telescope (BTA) at SAO. The emission nebulae with continuum-subtracted line ratio values of [S II]$\\lambda$$\\lambda$6716,6731 /H$\\alpha$ $\\geq$ 0.4 are identified as SNRs. A follow-up spectroscopy is obtained to confirm optical SNR identifications.} { We have identified nine new SNR candidates in M74 with [S II]/H$\\alpha$ $\\geq$ 0.4 as the basic criterion. We obtain [S II]/H$\\alpha$ ratio in the range from 0.40 to 0.91 and H$\\alpha$ intensities from 2.8 $\\times$ $10^{-15}$ erg cm$^{-2}$ s$^{-1}$ to 1.7 $\\times$ $10^{-14}$ erg cm$^{-2}$ s$^{-1}$. We also present spectral follow-up observations of these SNR candidates. However, we are able to spectrally confirm only three of them (SNR2, SNR3, and SNR5). The lack of confirmation for the rest might be due to the contamination by the nearby H II emission regions as well as due to the inaccurate positioning of the long slit on these objects. In addition, we search the $XMM$-$Newton$\\ and $Chandra$ Observatory archival data for the X-ray counterparts to the optically identified candidates. We find positional coincidence with only three SNR candidates, SNR1, SNR2, and SNR8. The spectrum of SNR2 yields a shock temperature of 10.8 keV with an ionization timescale of 1.6 $\\times$ 10$^{10}$ s cm$^{-3}$ indicating a relatively young remnant in an early Sedov phase which is not supported by our optical wavelength analysis. Given the high luminosity of 10$^{39}$ erg s$^{-1}$ and the characteristics of the X-ray spectrum, we favor an Ultra Luminous X-ray Source interpretation for this source associated with an SNR. We calculate an X-ray flux upper limit of 9.0 $\\times$ $10^{-15}$ erg cm$^{-2}$ s$^{-1}$ for the rest of the SNRs including spectroscopically identified SNR3 and SNR5.} {} ", "introduction": "M74 (NGC 628) is an Sc type spiral galaxy with a $\\sim$ 6$^{\\circ}$ inclination angle at an assumed distance of 7.3 Mpc (Sharina et al.1996, Soria et al. 2004). It is the brightest member of the small M74 group in the Pisces constellation. M74 has spiral arms well formed with bright blue star clusters and dark cosmic dust lanes. In the early 2000, two supernovae (SN 2002ap and SN 2003gd) were identified in M74 and studied extensively. The supernova event SN 2002ap is one of the four Type Ic SNe observed in X-rays (Van Dyk et al. 2003). SN 2003gd, which is found to be a nearby Type II-P (plateau) event has a confirmed red giant progenitor (Hendry et al. 2005). Leievre \\& Roy (2000) have recovered well over 100 small and isolated H II regions and measured their fluxes in the extreme outer disk of M74 with R$>$R$_{25}$ making use of deep H$\\alpha$ narrowband imaging. They show that massive star formation rate measured by the azimuthally averaged H$\\alpha$ surface brightness decreases monotonically from the center out to R$\\sim$20 kpc beyond which it drops rapidly. Elmegreen et al. (2006) have examined the size and luminosity distributions of the star forming regions in the galaxy with the Hubble Space Telescope. Their results suggest clumping of stars having a scale free nature in the galaxy disk. Fathi et al. (2007) have studied the internal kinematics of M74 over its entire face using Fabry-Perot interferometry with a good angular resolution, and confirmed the presence of an inner rapidly rotating disc-like component caused by the slow secular evolution of the large-scale spiral arms together with the oval structure. They also detect over 300 H II regions in the galaxy with calibrated luminosities and diameters up to about 300 pc using continuum subtracted narrow band images in H$\\alpha$. M74 is observed in the X-rays using the $Chandra$ and $XMM$-$Newton$\\ data by Soria et al. (2004). 74 discrete X-ray sources are detected in the combined $Chandra$ observations to a detection limit of $\\approx$ 6 $\\times$ 10$^{36}$ erg s$^{-1}$. They estimate 15$\\%$ of the M74 discrete sources to have soft colors. They also calculate thermal SNRs with luminosities $\\approx$ 2 $\\times$ 10$^{37}$ erg s$^{-1}$. Two bright X-ray sources (XMMU J013636.5+155036, CXOU J13651.1+154547) are detected by $XMM$-$Newton$\\ in M74 (Soria et al. 2004), while only the latter is detected by Chandra (Krauss et al. 2003). It is established that both of the sources are variable. It is also argued that they can be Ultra Luminous X-ray sources (ULX), however their true nature remains unclear. Radio observations of M74 have been carried out during the Arecibo Galaxy Environment Survey (AGES). This survey is designed to investigate M74 group environment in 21 cm with a better sensitivity and higher spatial and velocity resolution than previous observations (Auld et al. 2006). They obtain the spatial distribution of HI for M74 and other selected galaxies. In the UV regime, the surface brightness and color profiles for M74 are calculated by Cornett et al. (1994). They show that M74 disk has sustained significant star formation over the last five hundred million years. In this paper, we present our results on the search for new supernova remnant (SNR) candidates in M74 using the observations by RTT150 and 6 m-BTA telescopes. We also make use of archived $Chandra$ data to look for their X-ray counterparts. SNR studies are important in the theories of interstellar medium (ISM), and star formation since supernovae inject large amounts of matter, and energy into the ISM. However, in spite of the quite large number of Galactic SNRs their observations are impeded by several limitations such as the uncertainty in distances to individual objects and high extinctions along the line of sight in several regions of the Galactic plane. The limitations and uncertainties are inherently much less in extragalactic samples. Assuming that all SNRs are of the same distance to us for a given galaxy, we can easily compare their observed properties. Relative positions of such SNRs are determined with more precision. Once we know the positions of SNRs, their distances with respect to H II regions and spiral arms can easily be calculated (Matonick \\& Fesen 1997; Blair \\& Long 1997, 2004). SNRs are identified in a number of nearby spiral galaxies using optical observations (e.g. D$^\\prime$Odorico, Dopita \\& Benveuti 1980; Braun \\& Walterbos 1993; Magnier et al. 1995; Matonick \\& Fesen 1997; Matonick et al. 1997; Gordon et al. 1998,1999; Blair \\& Long 1997, 2004; Sonbas et al. 2009), X-ray observations (Pence et al. 2001; Ghavamian et al. 2005), and radio observations (Lacey et al. 1997; Lacey \\& Duric 2001; Hyman et al. 2001). Multiwavelength surveys of SNRs are also carried out by Pannuti et al. (2000, 2002, 2007). In our work, we use a well known and accepted criterion (S II/H$\\alpha$ $\\geq$ 0.4) proposed by Mathewson \\& Clarke (1973) to differentiate SNRs from typical H II regions. Rationale for this lies in the fact that in a typical H II region, sulfur is usually expected to be in the form of $S^{++}$ due to strong photoionizing fluxes from central hot stars, making the ratio [S II]/H$\\alpha$, typically, in the range $\\sim$ 0.1 - 0.3. Shock waves produced by SN explosion propagate through the surrounding medium. The matter cools sufficiently behind these waves and variety of ionization states occur including S$^{+}$. This might be the reason for the increased [S II]/H$\\alpha$ ratio observed in SNRs. It follows that almost all discrete emission nebulae satisfying the above criterion are expected to be shock-heated. The organization of the paper is as follows; In section 2, we discuss our imaging and spectroscopic observations and related data reduction. Search results in the optical band, the identification of SNRs, and search for their X-ray counterparts are presented in section 3. Finally, conclusions and discussions of our results are provided in section 4. ", "conclusions": "We conducted a survey of SNRs using optical imaging and spectroscopic measurements in the nearby spiral galaxy M74. In this survey, we used blinking between continuum-subtracted H$\\alpha$ and continuum-subtracted [S II] images to deduce SNR candidates for a given ratio between the lines. The SNRs were confirmed by additional spectroscopic observations and finally, comparing the optically detected SNRs with archived Chandra observations. Nine SNRs were identified in M74 by our method. Three of these have shown positional coincidences with X-ray sources. SNR2 is our best candidate to be an SNR using optical wavelength analysis and the X-ray source is consistent with a ULX interpretation, making it one of the few associations between a SNR and a ULX. We calculated an X-ray flux upper limit of 9.0 $\\times$ 10$^{-15}$ erg cm$^{-2}$ s$^{-1}$ and a luminosity upper limit of 5.0 $\\times$ 10$^{37}$ erg s$^{-1}$ for our new SNR candidates excluding SNR1, SNR2 and SNR8. Our spectroscopically detected SNRs have shock velocities $<$ 85 kms$^{-1}$ indicating that they fall into a low shock velocity range implying very old remnants from which little X-ray emission is expected. According to the SN rate calculation of Matonick \\& Fesen (1997) in such galaxies, about half of all SNe are of Type Ib/c or Type II, produced by massive stars. In turn, only half of these SNe are located in regions with enough ambient density to produce a detectable SNR. This means that, only about a quarter of all SN events may leave easily detectable optical remnants. In this respect four times more SN events should have exploded in M74 and we detected nine SNRs. With this reasoning, we would expect to see $\\sim$ 40 SNRs. If the optically observable lifetime of an SNR is assumed to be $\\sim$ 20,000 years (Braun, Goss, \\& Lyne 1989), a crude SN occurrence rate for M74 can be estimated giving a value of about 1 per $\\sim$ 500 years. In the light of observation of two recent SNe from M74, our result can be reconciled if we accept the fact that more SNe occur at places that are difficult to reach optically, (i.e. in gas-rich regions of the galaxy). \\begin {acknowledgements} We thank the TUBITAK National Observatory (TUG) and the Special Astrophysical observatory (SAO) for their support for observing times and equipments. \\end {acknowledgements}" }, "1005/1005.2466_arXiv.txt": { "abstract": " ", "introduction": "\\label{sec:1} In 1998, two independent supernovae (SN) observation groups found that our universe is undergoing an accelerated expansion at the present stage, through the observations of distant type Ia supernovae~\\cite{SN}. This implies that there exists a mysterious component, dark energy, which has large enough negative pressure, responsible for the cosmic acceleration. Many other astronomical observations, such as surveys of the large scale structure (LSS)~\\cite{LSS} and measurements of the cosmic microwave background (CMB) anisotropy~\\cite{CMB}, also firmly indicate that dark energy is the dominant component in the present-day universe. It is commonly believed that exploring the nature of dark energy is one of the focuses in the realm of both cosmology and theoretical physics today. The most obvious candidate for dark energy is the famous Einstein's cosmological constant $\\lambda$ which has the equation of state $w=-1$. However, as is well known, the cosmological constant is plagued with the ``fine-tuning'' and ``cosmic coincidence'' problems \\cite{problems}. Another promising candidate for dark energy is the dynamical scalar field, a slowly varying, spatially homogeneous component. An example of scalar-field dark energy is the so-called quintessence\\cite{quintessence}, a scalar field $\\phi$ slowly evolving down its potential $V(\\phi)$. Provided that the evolution of the field is slow enough, the kinetic energy density is less than the potential energy density, giving rise to the negative pressure responsible to the cosmic acceleration. So far, in order to alleviate the cosmological-constant problems and explain the accelerated expansion, a wide variety of scalar-field dark energy models have been proposed. Besides quintessence, these also include phantom, $k$-essence, tachyon, ghost condensate and quintom amongst many. However, we should note that the mainstream viewpoint regards the scalar-field dark energy models as a low-energy effective description of the underlying theory of dark energy. It is generally believed by theorists that we cannot entirely understand the nature of dark energy before a complete theory of quantum gravity is established. However, although we are lacking a quantum gravity theory today, we still can make some efforts to explore the nature of dark energy according to some principles of quantum gravity. The holographic dark energy model \\cite{Li:2004rb} is just an appropriate example, which is constructed in light of the holographic principle of quantum gravity theory. That is to say, the holographic dark energy model possesses some significant features of an underlying theory of dark energy. More recently, a new model consistent with the holographic principle, the agegraphic dark energy model, has been proposed in ~\\cite{Cai:2007us}, which takes into account the uncertainty relation of quantum mechanics together with the gravitational effect in general relativity. While, by far, a complete theory of dark energy has not been established presently, we can, however, speculate on the underlying theory of dark energy by taking some principles of quantum gravity into account. The agegraphic dark energy model is no doubt a tentative in this way. Now, we are interested in that if we assume the holographic/agegraphic vacuum energy scenario as the underlying theory of dark energy, how the low-energy effective scalar-field model can be used to describe it. In this direction, some work has been done, see, e.g., \\cite{holoscalar,Zhang:2009un,Zhang:2008mb,Cui:2009ns}. The agegraphic versions of scalar-field models, such as quintessence and tachyon, have been constructed \\cite{Zhang:2008mb,Cui:2009ns}. In this paper, we focus on the canonical scalar-field description of the agegraphic dark energy, namely, the ``agegraphic quintessence'' \\cite{Zhang:2008mb}. In recent years, cosmological-constant/dark-energy problem has been studied by string theorists within the string framework. It is generally considered that string theory is the most promising consistent theory of quantum gravity. Based on the KKLT mechanism \\cite{Kachru:2003aw}, a vast number of metastable de Sitter vacua have been constructed through the flux compactification on a Calabi-Yau manifold. These string vacua can be described by the low-energy effective theories. Furthermore, it is realized that the vast series of semiclassically consistent field theories are actually inconsistent. These inconsistent effective field theories are believed to locate in the so-called ``swampland'' \\cite{Vafa:2005ui}. The self-consistent landscape is surrounded by the swampland. Vafa has proposed some criterion to the consistent effective field theories \\cite{Vafa:2005ui}. Moreover, it was conjectured by Arkani-Hamed et al. \\cite{weakgrav} that the gravity is the weakest force, which helps rule out those effective field theories in the swampland. Arkani-Hamed et al. pointed out \\cite{weakgrav} that when considering the quantum gravity, the gravity and other gauge forces should not be treated separately. For example, in four dimensions a new intrinsic UV cutoff for the U(1) gauge theory with single scalar field, $\\Lambda=g M_{\\rm{p}}$, is suggested, where $g$ is the gauge coupling \\cite{weakgrav}. In \\cite{hls}, the weak gravity conjecture together with the requirement that the IR cutoff should be smaller than the UV cutoff leads to an upper bound for the cosmological constant. In addition, for the inflationary cosmology, the application of the weak gravity conjecture shows that the chaotic inflation model is in the swampland \\cite{Huangchaotic}. This conjecture even implies that the eternal inflation may not be achieved \\cite{HLW}. Furthermore, Huang conjectured \\cite{Huang:2007qz} that the variation of the inflaton should be smaller than the Planck scale $M_{\\rm{p}}$, and this can make stringent constraint on the spectral index. More recently, the weak gravity conjecture has been applied to the dark-energy problem. It is suggested that the variation of the quintessence field value $\\phi $ should be less than $M_{\\rm{p}}$. This criterion may give important theoretical constraints on the equation-of-state parameter of quintessence models, and some of these constraints are even stringent than those of the present experiments \\cite{Huang:2007mv}. The criterion $|\\Delta\\phi(z)|/M_{\\rm{p}}\\leq 1$ has also been used to put theoretical constraints on other canonical scalar-field dark energy models; see, e.g., \\cite{Ma:2007av,Wu:2007pq,Chen:2008gi}. In this paper we shall investigate the possible theoretical limits on the parameter $n$ of the agegraphic quintessence from the weak gravity conjecture. In the next section, we will briefly review the new agegraphic dark energy model proposed in~\\cite{Wei:2007ty}. In Sec.~\\ref{sec3}, we will give the possible theoretical limits on the parameter $n$ of the agegraphic quintessence model from the weak gravity conjecture. Conclusion will be given in Sec.~\\ref{sec4}. \\setcounter{equation}{0} ", "conclusions": "\\label{sec4} To summarize, in this paper we have investigated the theoretical limits on the parameter $n$ of the agegraphic quintessence model by considering that the variation of the quintessence scalar field $\\phi$ should be less than the Planck mass $M_{\\rm{p}}$. The agegraphic dark energy can mimic the behavior of a quintessence scalar-field dark energy, so the quintessence model can be used to effectively describe the agegraphic dark energy. In this paper, we have tested the single-field and multi-field agegraphic quintessence models by using the weak gravity conjecture. We believe that the low-energy effective field theory is not applicable in the trans-Planckian field space. We have shown that for both single-field and multi-field agegraphic quintessence models the weak gravity conjecture leads to the same theoretical limit, $n\\lesssim 2.5$, which is inconsistent with the current observational constraint $2.637 0.7$). Interestingly, there are also relatively few radial velocity (RV) planets discovered in the period interval of $10$--$100$~days (top panel). The depletion of planets in this regime can be explained by the stellar-mass-dependent lifetime of the protoplanetary disk and its effect on migration \\citetext{\\citealp{burkert:2007} and \\citealp{currie:2009}}. The 10.86~day period planet \\hatcurb{} was discovered by the Hungarian-made Automated Telescope Network \\citep[HATNet;][]{bakos:2004}. Operational since 2003, it has covered approximately 11\\% of the Northern sky, searching for TEPs around bright stars ($8\\lesssim I \\lesssim 12.5$\\,mag). HATNet employs six wide field instruments: four at the Fred Lawrence Whipple Observatory (FLWO) in Arizona, and two on the roof of the Submillimeter Array hangar (SMA) of SAO in Hawaii. Since 2006, HATNet has announced and published 14 TEPs. This work describes the fifteenth such discovery, the first TEP with a period above $10$~days discovered by a ground-based photometric survey. ", "conclusions": "\\label{sec:discussion} The transiting extrasolar planet (TEP) \\hatcurb{} reported in this paper is among the few with orbital periods greater than $8$~days.\\footnote{Currently, there are no TEPs between $5.6$ and $8$~days, and, together with \\hatcurb{}, there are only four TEPs with periods greater than $10$~days. The periods are more densely distributed below $5$~days.} We summarize the properties of the currently known long-period TEPs in \\reftabl{ltep}. \\ifthenelse{\\boolean{emulateapj}}{ \\begin{deluxetable*}{lccrrcccrcccccr} }{ \\begin{deluxetable}{lccrrcccrcccccr} \\rotate } \\tabletypesize{\\scriptsize} \\tablewidth{0pc} \\tablecaption{Long-period TEPs\\tablenotemark{a} \\label{tab:ltep}} \\tablehead{ \\colhead{Name} & \\colhead{$\\mpl$} & \\colhead{$\\rpl$} & \\colhead{$M_{\\rm core}$\\tablenotemark{b}} & \\colhead{$T_{\\rm eq}$\\tablenotemark{c}} & \\colhead{Period} & \\colhead{a} & \\colhead{e} & \\colhead{$T_{14}$} & \\colhead{Star(\\teff)} & \\colhead{[Fe/H]} & \\colhead{Age\\tablenotemark{e}} & \\colhead{$\\mstar$} & \\colhead{$\\rstar$} & \\colhead{V} \t\\\\ \\colhead{} & \\colhead{[$\\mjup$]} & \\colhead{[$\\rjup$]} & \\colhead{[\\mearth]} & \\colhead{[K]} & \\colhead{[d]} & \\colhead{[AU]} & \\colhead{} & \\colhead{[h]} & \\colhead{[K]} & \\colhead{[dex]} & \\colhead{[Gyr]} & \\colhead{[$\\msun$]} & \\colhead{[$\\rsun$]} & \\colhead{[mag]} } \\startdata WASP-8b & 2.25 & 1.05 & 52 & 940 & 8.15872 & 0.080 & 0.31 & 3.5 & G6 (5600) & $+0.17$ & 4.0 & 1.03 & 0.95 & 9.9 \\\\ CoRoT-6b & 2.96 & 1.17 & 0 & 1020 & 8.88659 & 0.086 & 0.10\\tablenotemark{d} & 3.8 & F5 (6090) & $-0.20$ & 3.0 & 1.06 & 1.03 & 13.9 \\\\ CoRoT-4b & 0.72 & 1.19 & 0 & 1070 & 9.20205 & 0.090 & 0.00 & 4.4 & F0 (6190) & $+0.00$ & 1.0 & 1.10 & 1.15 & 13.7 \\\\ HAT-P-15b & \\hatcurPPmshort & \\hatcurPPrshort & $10$ & 910 & \\hatcurLCPshort & $0.096$ & $0.19$ & $5.5$ & G5 (5568) & \\hatcurSMEiizfehshort & \\hatcurISOageshort & \\hatcurISOmshort & \\hatcurISOrshort & 12.2 \\\\ HD~17156b &\t 3.21 & 1.02 & 136 & 880 &\t21.2169 & 0.162 & 0.68 & 3.2 & G0 (6079) & $+0.24$ & 3.1 & 1.24 & 1.45 & 8.2 \\\\ CoRoT-9b &\t 0.84 & 1.05 & 18 & 410 &\t95.2738 & 0.407 & 0.11 & 8.1 & G3 (5625) & $-0.01$ & 3.0 & 0.99 & 0.94 & 13.7 \\\\ HD~80606b &\t 4.08 & 0.98 & 195 & 400 &\t111.436 & 0.455 & 0.93 & 11.9 & G5 (5574) & $+0.43$ & 7.6 & 1.01 & 1.01 & 8.9 \\\\ [-1.5ex] \\enddata \\tablenotetext{a}{All parameters are from {\\em http://exoplanet.eu/}, except for those of WASP-8b \\cite[see][]{queloz:2010} and HD~80606b \\citetext{see \\citealp{moutou:2009} (\\teff) and \\citealp{hebrard:2010}}.} \\tablenotetext{b}{Based on the models of \\cite{fortney:2007}; zero core mass indicates that the planet has larger radius than predicted by the models at zero core mass.} \\tablenotetext{c}{Computed at a distance given by the semi-major axis.} \\tablenotetext{d}{Upper limit.} \\tablenotetext{e}{Ages have been taken either from `exoplanet.eu' or from currently published papers. For CoRoT-6b we adopted an intermediate value from the range of $2.5$--$4.0$~Gyr listed by \\cite{fridlund:2010}. For CoRoT-9b \\cite{deeg:2010} give a wide range of $0.2$--$8$~Gyr, and we adopted a plausible intermediate-to-low age.} \\ifthenelse{\\boolean{emulateapj}}{ \\end{deluxetable*} }{ \\end{deluxetable} } \\hatcurb{} is the longest period TEP discovered by ground-based photometric surveys. With a duration of $5.5$~hours, the transit of \\hatcurb{} is at the limit in length that can be pursued from a single site ground-based observation (the only planet with a longer period than \\hatcurb{} that can still be followed up from a single site is HD~17156b). The possibility of performing single-site full followup observations is a valuable property, since systematics are more difficult to handle when low signal-to-noise ratio data from different sites are combined. The orbit is eccentric with a high significance. Except for CoRoT-4b (and perhaps for CoRoT-6b), all of the 7 long-period planets have this property. Most of the planet masses are greater than $2\\mjup$. The host star masses are lower than $\\sim1.2$\\msun\\,, in agreement with the upper mass limit of $\\sim1.4$\\msun\\, of the extrasolar planets situated in the `period valley' between $\\sim0.1$~AU and $\\sim1$~AU \\citep{bowler:2010}. The relative depletion of planets in the $0.1$--$1$~AU region was recognized a few years ago when the number of extrasolar planets grew to the level that made more reliable statistical investigations possible \\citep{johnson:2007}. All host stars with semi-major axes smaller than $\\sim1$~AU have masses lower than $\\sim1.4$\\msun\\,. Above $\\sim1$~AU there are several stars with $M\\sim0.5$--$2.5$\\msun\\, \\citep{bowler:2010}. The distribution of the periods (or semi-major axes) of all known extrasolar planets is bimodal with a high significance (see upper panel of \\reffigl{pdistr}). Although the sample of multiple systems is not very extensive, it is interesting to note that the distribution becomes more uniform when considering only these systems \\cite{wright:2009}. Although some aspects of these properties can be explained by engulfment of planets in evolved giants \\citep{villaver:2009}, early stage of planet formation should play more important role. According to the models considering the formation phase \\citetext{\\citealp{burkert:2007} and \\citealp{currie:2009}, see also \\citealp{kretke:2009} for the role of magnetorotational instability in the protoplanetary disk} the bimodal distribution is attributed to the stellar-mass-dependent lifetime of the protoplanetary gas disk. For higher mass stars, the gas-depletion time scale is shorter due to accretion and increased EUV radiation, and therefore planets may stop inward migration too early and get stranded on orbits at high semi-major axes. For stellar masses $\\la1.2$\\msun\\, planets do migrate. The rate of migration is determined by the ratio of the viscous and depletion time scales.\\footnote{With a careful filtering of the currently available data we found that the valley exist at all host star masses, albeit for stars with $\\ga1.2$\\msun\\, the effect is indeed much stronger.} We investigated the agreement between the derived planetary mass and radius and predictions from current theoretical models. We interpolated the models of \\cite{fortney:2007} to the observed planet mass of $\\hatcurPPmlong$~$\\mjup$ and semi-major axis of equivalent solar irradiance (which is the same as the observed semi-major axis in this case). The iso-core lines in the $\\log(age)$--$\\log(R_{\\rm p})$ plane are shown in \\reffigl{fortney-model}. Although the errors are large, and the derived core mass (especially at the low end) depends very sensitively on the stellar/planet parameters, it seems likely that \\hatcurb{} is a H/He-dominated gas giant planet with a core mass most likely about $10$\\mearth\\,, but which could range from zero to perhaps $50$\\mearth\\,. \\hatcurb{} does not exhibit the ``radius anomaly'', where -- as is frequently the case for hot Jupiters -- the planetary radius is significantly larger than predicted by models such as that of \\cite{fortney:2007}. Furthermore, we find that for a hypothetical planet with the same mass and radius as those of \\hatcurb{}, the derived core mass is only a weak function of the period (semi-major axis) between $\\sim10$ and $1000$~days (it varies between $0$ and $10$\\mearth\\,). However, below $10$~days the core mass increases up to $60$\\mearth\\, (at $1$~day), and starts especially being sensitive below $4$~days. These all show that \\hatcurb{} resides in the period regime where standard planet structure models (i.e., the ones already tested on Solar System planets) are satisfactory, without resorting to additional effects (e.g., extra heat source, tidal heating, enhanced atmospheric opacities, etc., see \\citealp{miller:2009}) required by many of the hot Jupiters. In \\reftabl{ltep} we also added the derived core masses for the other long-period TEPs. We see that two of the shorter period members still exhibit some radius anomalies. For CoRoT-4b and CoRoT-6b, tidal heating \\citep{miller:2009} is unlikely, since they both have close to circular orbits. On the other hand, considering the possible ambiguities in the observed absolute planetary radii, it is possible that future followup observation will result in downward radius corrections of $3$--$5$\\% necessary to solve the apparent anomaly for these planets. This level of radius correction is not uncommon in the followup works of increasing accuracy \\citetext{e.g., the case of HAT-P-1b -- \\citealp{winn:2007} or that of HD~80606b -- \\citealp{fossey:2009} versus \\citealp{hebrard:2010}, and especially WASP-8b, with $10$\\% radius decrease in the very recent analysis of \\citealp{queloz:2010}}. It is also useful to compare the overall dependence of the planetary radius on the incident stellar flux. \\reffigl{flux-radius} shows this dependence for TEPs in the mass range of $(0.7$--$5.0)$\\mjup. Although the stellar fluxes for the long-period TEPs are still $2$--$3$ orders of magnitude larger than that of Jupiter, it seems that this is already enough to cause a substantial decrease in the planetary radius, and bring down the otherwise inflated values to `more normal' ones, corresponding to that of the Jupiter. We note that this happens in a quite wide mass range. \\begin{figure}[!ht] \\plotone{\\hatcurhtr_Fortney.eps} \\caption{ Position of \\hatcurb{} on the age --- planet radius diagram. The models of \\cite{fortney:2007} have been interpolated to the derived mass and solar equivalent semi-major axis of \\hatcur{}. Open circles are the tabulated/interpolated values from \\cite{fortney:2007}, continuous lines have been obtained by linear interpolation. Lines are labeled by the core mass (relative to the total planet mass). } \\label{fig:fortney-model} \\end{figure} \\begin{figure}[!ht] \\plotone{\\hatcurhtr_radius-fluxaver.eps} \\caption{ Incident stellar flux versus planet radius for TEPs of $0.7\\la M_{\\rm p}/M_{\\rm J}\\la 5.0$. The long-period planets of \\reftabl{ltep} are shown by whiter shade (green), the triangle at the left corner shows the position of Jupiter. } \\label{fig:flux-radius} \\end{figure} With V=$12.2$~mag, \\hatcur{} is still bright enough to be the subject of ground- and space-based observations. At a distance of $a=\\hatcurPParel$~AU from its host star, \\hatcurb{} allows us to investigate the properties of a colder planet, not influenced by the complicated hydrodynamical and radiative effects which may apply to canonical hot Jupiters (see, e.g., the discussion of the core mass insensitivity above). At the same time, the period of $\\hatcurLCPvshort$~days is still on a relatively short time scale, allowing frequent followup observations. The feasibility of possible follow-up observations is briefly summarized below. \\begin{itemize} \\item {\\em Search for other planets.} Long-term radial velocity (RV) and transit timing monitoring is justified due to the signature of excess RV scatter that is unlikely to be due to stellar activity (see \\refsecl{hispec}). Although we have made a very thorough search both in the RV and in the photometric data for a trace of a second planet, we have not found anything convincing. Due to the relatively small value of \\hatcurRVjitter\\ms\\, of the ``jitter'', the mass of the possible hidden companion is estimated in the sub-Neptune regime (except if there is a planet in, e.g., a 2:1 resonance, hiding in the eccentric solution -- see \\citealp{anglada:2010}). We note that here, the eccentricity of the orbit of \\hatcurb{} is a less forcing argument for searching for a putative second planet, because at the semi-major axis of \\hatcurb{} the circularization time scale due to tidal dissipation may be on the order of several Gyrs \\citep{matsumura:2008}. \\item {\\em Inclination of the orbital and stellar spin axes.} The measurement of the Rossiter-McLaughlin (R-M) effect for \\hatcur{} is somewhat challenging, due to its relatively low projected rotation rate of \\vsini$=$\\hatcurSMEiivsin\\,. The expected size of the effect is $\\sim20$\\ms\\,, similar to the one currently measured for WASP-8b with the same \\vsini\\, \\citep{queloz:2010}. Three of the seven long-period planets listed in \\reftabl{ltep} have yielded R-M data sufficient to compute the projected inclination of the stellar rotational and orbital axes. HD~17156 is aligned \\citep{narita:2009}, while HD~80606b is tilted \\citep{hebrard:2010}, and WASP-8b \\citep{queloz:2010} is apparently strongly tilted, suggesting a retrograde orbit. It is clear that measuring the R-M effect of long-period planets in more evolved systems such as \\hatcur{} is very important, due to the expected relaxation of the dynamics to a final equilibrium state. \\item {\\em Thermal emission.} With a proper choice of the infrared waveband and repeated observations, the detection of the occultation of \\hatcurb{} by ground-based instruments may be feasible. Assuming blackbody radiation, at a fixed wavelength the occultation depth is equal to $\\delta_{\\rm occ}=(R_{\\rm p}/R_{\\rm star})^2(T_{\\rm pl}/T_{\\rm star})$ \\citep{winn:2010}. Assuming $T_{\\rm eq}=T_{\\rm pl}$, at the moment of occultation (which is the periastron in the case of \\hatcurb{}) we have $T_{\\rm pl}\\sim1000$~K. It follows then that $\\delta_{\\rm occ}\\sim0.002$, which is measurable, especially if we consider that the occultation lasts for $\\sim4$~hours; short-enough to cover the event in one night and long-enough to gather sufficient number of data points (especially in the case of repeated observations) to reach the level of detection of several $\\sigma$. \\item {\\em Atmospheric absorption}. Measuring the small change in the transit depth due to the varying absorption level in different wavelengths (i.e., performing transmission spectroscopy) is even more challenging than detecting the occultation event. This is because the change in the transit depth is proportional to the scale-height $H$, which is proportional to the ratio of the local temperature to the gravity $T/g$ \\citep{winn:2010}. Since \\hatcurb{} is relatively cold and has a higher gravity than `standard' hot Jupiters, it is expected that the signal will be about $25$\\% of what is usually measured (assuming `standard' hot Jupiter $T$ and $g$ of $2000$~K and $20$\\mss\\,). Considering that the relative radius variation even for short-period targets like HD~189733 is of the order of $0.1$\\% \\citep{sing:2009}, measuring absorption features in colder planets such as \\hatcurb{} would indeed be very difficult. \\end{itemize} With the continuing work of ground-based wide-field photometric surveys, we expect to discover further TEPs in the $10$--$30$~day period regime. Since these will be mostly bright ($V\\la13$~mag) targets with orbital periods still considered to be relatively short, a wide range of followup works will be feasible. Therefore, these planets will fill the gap between the classical hot Jupiters and the long-period, but likely much fainter planetary systems (more akin to the Solar System) to be discovered by ongoing and future space projects." }, "1005/1005.3526_arXiv.txt": { "abstract": "Processes such as the solar wind sputtering and micrometeorite impacts can modify optical properties of surfaces of airless bodies. This explains why spectra of the main belt asteroids, exposed to these `space weathering' processes over eons, do not match the laboratory spectra of ordinary chondrite (OC) meteorites. In contrast, an important fraction of Near Earth Asteroids (NEAs), defined as Q-types in the asteroid taxonomy, display spectral attributes that are a good match to OCs. Here we study the possibility that the Q-type NEAs underwent recent encounters with the terrestrial planets and that the tidal gravity (or other effects) during these encounters exposed fresh OC material on the surface (thus giving it the Q-type spectral properties). We used numerical integrations to determine the statistics of encounters of NEAs to planets. The results were used to calculate the fraction and orbital distribution of Q-type asteroids expected in the model as a function of the space weathering timescale, $t_{\\rm sw}$ (see main text for definition), and maximum distance, $r^*$, at which planetary encounters can reset the surface. We found that $t_{\\rm sw}\\sim10^6$ yr (at 1 AU) and $r^* \\sim 5$ $R_{\\rm pl}$, where $R_{\\rm pl}$ is the planetary radius, best fit the data. Values $t_{\\rm sw}<10^5$~yr would require that $r^*>20$ $R_{\\rm pl}$, which is probably implausible because these very distant encounters should be irrelevant. Also, the fraction of Q-type NEAs would be probably much larger than the one observed if $t_{\\rm sw} > 10^7$ yr. We found that $t_{\\rm sw} \\propto q^2$, where $q$ is the perihelion distance, expected if the solar wind sputtering controls $t_{\\rm sw}$, provides a better match to the orbital distribution of Q-type NEAs than models with fixed $t_{\\rm sw}$. We also discuss how the Earth magnetosphere and radiation effects such as YORP can influence the spectral properties of NEAs. ", "introduction": "Measurements of the spectral properties of Near Earth Asteroids (NEAs) provide important evidence concerning the relationship between asteroids and the most common class of meteorites known as the ordinary chondrites (OCs). The tendency toward seeing OC-like spectral attributes among NEAs has been noted in multi-filter color observations (Rabinowitz 1998, Whiteley 2001), and in visible and near-infrared specrophotometric surveys (Binzel et al. 1996, 2004, 2010). In contrast, no spectral analogs of OCs have been found to date among the $\\sim$2000 surveyed main belt asteroids (MBAs), except for a case related to identified recent asteroid collisions (Moth\\'e-Diniz and Nesvorn\\'y 2008). The lack of spectrophotometric analogs for OC meteorites in the main belt is a long-debated and fundamental problem. It is now generally accepted that processes similar to those acting on the Moon, such as solar wind sputtering and micrometeorite impacts (Gold 1955, Pieters et al. 2000, see Hapke 2001 and Chapman 2004 for reviews), can darken and redden the initially OC-like spectrum of a fresh asteroid surface, giving it the `weathered' appearance (see Chapman 1996 and Clark et al. 2001, 2002a,b for direct evidence for asteroid space weathering processes from the NEAR-Shoemaker and Galileo spacecrafts). In the following text we will refer to processes that alter optical properties of surfaces of airless bodies as the `space weathering' (SW) effects. Since the SW processes should affect the MBAs and NEAs in roughly the same way (see, e.g., Marchi et al. 2006 for a study of the SW dependence on heliocentric distance), it may seem puzzling why a significant fraction of NEAs has an unweathered appearance (Binzel et al. 2004) while practically all spectroscopically surveyed MBAs are weathered. Several explanations have been proposed. To simplify the discussion of different models described below, we will use the following terminology taken from the standard asteroid taxonomy (Bus and Binzel 2002, DeMeo et al. 2009). We will define three categories of asteroid spectra: (1) Q-type spectra with deep absorption bands and shallow spectral slope similar to that of most OC meteorites in the RELAB database\\footnote{http://www.planetary.brown.edu/relab/}; (2) S-type spectra with shallow absorption bands and relatively steep spectral slope similar to that of weathered OCs; and (3) Sq-type spectra as the intermediate case between S and Q. See Bus and Binzel (2002) and DeMeo et al. (2009) for a formal definition of these categories. We will assume that asteroids with the Q-type spectra have essentially unweathered surfaces with OC-like mineralogy; the Sq- and S-type asteroids will be assumed to have moderately and strongly weathered surfaces with OC-like mineralogy. About 20\\% of chondritic NEAs surveyed at visible and/or near-infrared wavelengths have Q-type spectra while this fraction is essentially zero among MBAs. The standard interpretation of these results has been based on the presumption that Q-type asteroids are likely to be small. Indeed, the current spectrophotometric surveys of MBAs are largely incomplete in the size range of typical NEAs with diameters less then a few km. Special emphasis has therefore been given to asteroid-size-dependent processes, such as immaturity of regoliths on small asteroids and/or the shorter collisional lifetime of smaller asteroids (e.g., Johnson and Fanale 1973, Binzel et al. 1996, 2004, Rabinowitz 1998, Whiteley 2001). One possibility is that the observed spectral variations may be related to particle-size effects (Johnson and Fanale 1973), where the decreasing gravity results in a different size distribution of surface particles on typically smaller NEAs than on larger MBAs. However, the photometric parameters indicative of particle-size effects show little evidence of an asteroid diameter dependence, thereby giving doubt to this explanation (e.g., Clark et al. 2001, Masiero et al. 2009). Binzel et al. (2004) hypothesized that the SW size-dependency was because the survival lifetime against catastrophic disruption decreases with decreasing size. Thus, on average, as we examine smaller and smaller objects, we should see younger and younger surfaces. Surfaces showing Q-type spectral properties should thus exist, on average, only among the smallest asteroids, which become easy spectroscopic targets only when they enter into NEA space. Large, OC-like asteroids in the main belt should have, on average, space-weathered spectral properties, explaining why they are taxonomically classified as S types. ", "conclusions": "Several tidal effects may disturb the surface of a NEA during a distant planetary encounter.\\footnote{We only discuss distant encounters here. It is clear that the SL9-like tidal disruption, binary formation events, or events with significant mass shedding will erase any pre-existing surface features.} For example: (1) The interior structure of a rubble-pile asteroid may find a new equilibrium by re-arranging its components. This motion can produce landslides, degrade craters, ballistically displace surface material, or even remove the original layers from the asteroid. (2) Tidal stresses applied to a fractured interior may produce seismic shakes similar to, or perhaps more effective than, those generated by impacts. Consequently, surface morphology may be modified. (3) The tidal torque may spin up an asteroid. In surface segments where the centrifugal force exceeds gravity, regolith layers will be removed by carrying away the excess angular momentum. More subtle changes can occur in other surface parts of a spun-up asteroid. (4) If the tidal force becomes comparable to the object's gravity during encounter, an asteroid with large enough internal strength and a strengthless regolith may lose its regolith layer. These effects and their dependence on the encounter distance and speed are poorly understood. Some insights into this problem can be obtained from Richardson et al. (1998), where the authors performed numerical simulations of the effects of tidal gravity on a small asteroid with strengthless (rubble-pile) interior. In the most favorable case (slow encounter speed, fast prograde rotation), they found that significant mass shedding can occur up to $\\approx$5~$R_{\\rm pl}$. This sets a soft constraint on $r^*$. On one hand, $r^*$ can be larger than 5 $R_{\\rm pl}$ because the optically-active thin surface layer may be vulnerable to even tiniest perturbations that were not considered in the Richardson et al. model. On the other hand, when averaging over all encounter geometries and plausible asteroid spin states, the mean $r^*$ can become lower than 5 $R_{\\rm pl}$. Thus, for the lack of additional constraints on $r^*$, we will tentatively assume below, as a guideline for discussion, that $r^* \\sim 5$ $R_{\\rm pl}$. If we set $r^*=5$ $R_{\\rm pl}$ our results described in \\S3 and \\S4 imply that $t_{\\rm sw}\\sim1$~My. At first sight, this SW timescale seems to be comparable to that obtained from comparative studies of asteroid families in the main belt and OC meteorites in the RELAB database (NJWI05, Vernazza et al. 2009). For example, Vernazza et al. (2009) proposed that the SW timescale is $\\lesssim1$ My. Their result hinges on observations of two largest members of the Datura family that formed by a catastrophic breakup $\\approx$0.5 My ago (Nesvorn\\'y et al. 2006, Vokrouhlick\\'y et al. 2009). These two objects, 1270 Datura and 90265 2003CL5, appear to be significantly (but not completely) space weathered (Moth\\'e-Diniz and Nesvorn\\'y 2008), which implies that the SW timescale should be comparable to or shorter than the Datura family's age. This poses a problem because $t_{\\rm sw} \\lesssim 0.5$~My does not fit the NEA constraint (unless $r^* > 5$ $R_{\\rm pl}$). Below we discuss possible solutions to this problem. Observations of 2001 WY35, one of the smallest known members of the Datura family (absolute magnitude $H=17$), indicate that this object is not space weathered at all (Moth\\'e-Diniz and Nesvorn\\'y 2008). If these observations were correct, they would indicate that (at least some) km-sized asteroids may weather on timescales significantly longer than $\\approx$0.5~My. For example, small km-sized fragments ejected from asteroid breakup events may not retain/accumulate sufficient regolith layer on their surface in the immediate aftermath of the collision. The SW effects may be delayed for such objects until a particulate (SW-sensitive) surface layer develops on their surface, for example, by subsequent impact shattering of the exposed rock. Thus, the regolith formation and `gardening' can be an important part of the problem (Jedicke et al. 2004, Willman et al. 2008). We should not forget that the two constraints on the SW timescale discussed here come from studies of two distinct population of objects that are affected by different physical processes. The asteroids in the main-belt families are born by violent collisions and spend most of their lifetime beyond 2 AU. The NEAs, on the other hand, are exposed to more extreme solar-wind and temperature environment. They are olivine-rich and may therefore be more susceptible to SW effects than an average MBA (Sasaki et al. 2001, Marchi et al. 2005). While large impacts on NEAs should be rare, bombardment of their surface by $D\\sim100$ $\\mu$m particles should be more intense than on MBAs due to the larger number density of micrometeoroids at $<$2 AU (Gr\\\"un et al. 1985). Also, distant planetary encounters of NEAs should produce more gentle effects than catastrophic collisions of MBAs, thus giving the initial surface different attributes. In summary, the {\\it proper} SW timescale that measures the progression of SW under ideal conditions (e.g., in absence of regolith gardening) may be substantially shorter than the {\\it apparent} SW timescale that arises from combination of different effects, and these effects most likely operate on different timescales in the NEA and MBA environments. Another interesting possibility is related to the effects of the Earth magnetosphere on loose particulate material on a small asteroid's surface. The Earth magnetosphere extends to $\\approx$12 $R_{\\rm Earth}$ in the direction toward the Sun, $\\approx$15 $R_{\\rm Earth}$ in apex and antapex directions, and $\\approx$25 $R_{\\rm Earth}$ the anti-helion direction. The tail region stretches well past 200 $R_{\\rm Earth}$, and the way it ends is not well-known. The magnetic field ranges from 30-60 $\\mu$T at the Earth's surface and falls roughly as $1/r^3$ with distance $r$ toward the edge of the magnetosphere. Thus, if a 100-m-sized NEA passes at distance 10 $R_{\\rm Earth}$, a 10-$\\mu$m surface dust grain subject to the Lorentz force would levitate if previously charged to $>10^8$ e. Such charge is plausible for an asteroid surface of sufficiently high electrical resistivity. It is not clear, however, whether the Lorentz force effect can be more significant than the electrostatic levitation (Lee 1996) and/or van der Waals forces (Scheeres and Hartzell 2010). While speculative, the effects of Earth magnetosphere could possibly allow for larger $r^*$ values than those expected for tidal gravity. This could perhaps help to resolve some of the discrepancy between different measurements of the SW timescale discussed above. For example, with $r^*=20$ $R_{\\rm Earth}$, Fig. \\ref{prob2} would imply that $t_{\\rm sw} \\approx$ 0.25 My. The orbits of Q-type NEAs in Fig. \\ref{ics} hint on bimodal distribution with a group of 7 objects with $a\\lesssim1$ AU and largely spread inclination values, and 12 objects with $a\\gtrsim1.5$ AU and $i\\lesssim10^\\circ$. Using planetary encounters as the main agent that resets the SW clock, we were not able to fit both groups simultaneously. We found that the model with fixed $t_{\\rm sw}$ can match the low-$a$ group but it fails to fit the observed fraction of Qs with $a>1.5$ AU (e.g., Fig. \\ref{m2}). On the other hand, the model with $t_{\\rm sw} \\propto q^2$ matches the high-$a$ group (Fig. \\ref{m4}) but it fails to produce the observed large fraction of Qs with $a\\lesssim1$ AU and $q<0.5$ AU. This is puzzling. The problem may be related to biases in the current sparse spectrophotometric data. Alternatively, we may be missing some important physical effect in the model. For example, a small irregular object can be spun up by a radiation effect known as YORP and shed mass (e.g., Walsh et al. 2009; see Bottke et al. 2006 for a recent review of YORP). This could lead to a partial or global removal of the space weathered material and exposure of fresh material on the surface. This effect can therefore be important. Unfortunately, the timescale on which the surface of a typical NEA can be reset by YORP is poorly understood. Kaasalainen et al. (2007) determined that 1862 Apollo ($a=1.47$ AU, $D=1.4$ km) is spun up by YORP on a characteristic timescale $t_{\\rm YORP} = \\omega/(d\\omega/dt) \\sim 2.6$ My, where $\\omega$ is the spin rate. Starting from its current 3-hour spin period, 1862 Apollo is thus expected to be spun up to a $\\sim$2-hour period (and start shedding mass) in $\\sim$2 My. This timescale is probably at least slightly longer than the one on which the surface of 1862 Apollo should be reset by planetary encounters indicating that the YORP effect can be ignored for 1862 Apollo. Since $t_{\\rm YORP} \\propto D^2 a^2 \\sqrt{1-e^2}$ (e.g., Nesvorn\\'y et al. 2007), however, the YORP effect can become more important than planetary encounters for small NEAs that orbit closer to the Sun than 1862 Apollo. For example, a sub-km NEA with $a<1$ AU can have $t_{\\rm YORP}$ several times shorter than 1862 Apollo. We therefore speculate that the YORP effect can contribute to the observed excess of Q-type NEAs in these low-$a$ orbits. A detailed analysis of this problem goes beyond the scope of this paper." }, "1005/1005.1653_arXiv.txt": { "abstract": "We examine the clustering properties of a population of quasars drawn from fully hydrodynamic cosmological simulations that directly follow black hole growth. We find that the black hole correlation function is best described by two distinct components: contributions from BH pairs occupying the same dark matter halo ('1-halo term', $\\xi_{\\rm{BH,1h}}$) which dominate at scales below $\\sim 300 \\: \\rm{kpc \\: h^{-1}}$, and contributions from BHs occupying separate halos ('2-halo term', $\\xi_{\\rm{BH,2h}}$ ) which dominate at larger scales. From the 2-halo BH term we find a typical host halo mass for faint-end quasars (those probed in our simulation volumes) ranging from $M\\sim 10^{11}$ to a few $10^{12} M_{\\odot}$ from $z=5$ to $z=1$ respectively (consistent with the mean halo host mass). The BH correlation function shows a luminosity dependence as a function of redshift, though weak enough to be consistent with observational constraints. At small scales, the high resolution of our simulations allows us to probe the 1-halo clustering in detail, finding that $\\xi_{\\rm{BH,1h}}$ follows an approximate power law, lacking the characteristic decrease in slope at small scales found in 1-halo terms for galaxies and dark matter. We show that this difference is a direct result of a boost in the small-scale quasar bias caused by galaxies hosting multiple quasars (1-subhalo term) following a merger event, typically between a large central subgroup and a smaller, satellite subgroup hosting a relatively small black hole. We show that our predicted small-scale excess caused by such mergers is in good agreement with both the slope and amplitude indicated by recent small-scale measurements. Finally, we note the excess to be a strong function of halo mass, such that the observed excess is well matched by the multiple black holes of intermediate mass ($10^7-10^8$ M$_{\\odot}$) found in hosts of $M \\sim 4-8 \\times 10^{11} M_\\odot$, a range well probed by our simulations. ", "introduction": "With supermassive black holes being found at the centre of most galaxies \\citep{1995ARA&A..33..581K}, interest in quasars has increased significantly, with substantial investigation into fundamental relations between black hole masses and their host galaxies' properties \\citep{1998AJ....115.2285M, 2000ApJ...539L...9F, 2000ApJ...539L..13G, 2002ApJ...574..740T, 2007ApJ...655...77G}. In addition to these relations, statistical studies of the spatial clustering of quasars provide the potential to better understand the relation between quasars, their hosts and the underlying dark matter distribution, as well as estimate quasar lifetimes \\citep[see, e.g.,][]{HaimanHui2001, MartiniWeinberg2001} across a relatively large range of redshift. For example, strong clustering would suggest quasars should reside in massive groups. If so, they should be rare and in order to reproduce the quasar luminosity density, they must have long lifetimes. Conversely, low correlation would suggest more common quasars, and thus shorter quasar lifetimes. Early studies of quasar clustering produced varying results for the clustering amplitude, with no clear agreement on overall evolution with redshift, some suggesting minimal or decreasing clustering evolution \\citep{MoFang1993, CroomShanks1996}, while others found an increase in clustering with redshift \\citep{Kundic1997, LaFranca1998}. These findings were generally poorly constrained due to the small sizes of available quasar samples. With the emergence of large scale surveys such as Sloan Digital Sky Survey \\citep{2000AJ....120.1579Y} and the Two-degree Field QSO Redshift Survey \\citep{2002MNRAS.333..279L}, substantially larger catalogs have been compiled, permitting more detailed investigation into the clustering properties of quasars, and many recent studies have been made into this area \\citep[e.g.][]{LaFranca1998, Porciani2004, Croom2005, Shen2007, Myers2007, daAngela2008, Shen2009, Ross2009}. These recent studies have found evidence for an increase in clustering amplitude with redshift \\citep{LaFranca1998, Porciani2004}, primarily for $z>2$, in agreement with predictions from simulations \\citep[see, e.g.][]{Bonoli2009, Croton2009}. In addition to overall evolution, the luminosity dependence (if any) of large-scale clustering can provide significant insight into what quasar populations dominate different luminosity ranges. For example, the model of \\citet{2005ApJ...630..705H, 2005ApJ...630..716H, 2005ApJ...632...81H, 2005ApJ...625L..71H, 2006ApJS..163....1H} suggests that bright and faint quasars are similar objects which are observed at different phases of their lifetimes, rather than being fundamentally different populations of quasars (as simpler, 'on-off' models assume). This model would suggest that both bright and faint quasars should populate similar halos. Thus, while there may be some correlation between peak luminosity and host halo mass, clustering dependence on instantaneous luminosity should be relatively weak, particularly when compared to more traditional 'on-off' models of quasar luminosity \\citep{Lidz2006}. Recent observational studies have generally found a lack of luminosity dependence in the correlation function \\citep[see, e.g.,][]{Croom2005, Myers2007, daAngela2008}, though \\citet{Shen2009} found evidence for some, though weak, luminosity dependence. Several semi-analytic models have also been used, finding differing luminosity dependences, such as a significant dependence for sufficient luminosity ranges, but limited when considering only luminosities probed by observation \\citep{Bonoli2009}, or weak dependence at low redshift ($z<1$), but stronger at higher redshift \\citep{Croton2009}. In addition to large scale behavior, the possibility of excess quasar clustering on very small scales has arisen in several recent studies. While some observed quasar pairs are believed to be the result of gravitationally lensed quasars, it has been proposed that others may be physically distinct quasar binaries, which would suggest quasars cluster much more strongly on small scales than extrapolation of large scale clustering would imply \\citep{Djorgovski1991, Hewett1998, Kochanek1999, Mortlock1999}, suggesting a connection between galaxy mergers and quasar activity \\citep[see, e.g.][]{Kochanek1999}. However, investigating the smallest scale clustering has typically been problematic due to observational limitations (such as fiber collisions preventing small-separation pairs from being distinguished as distinct objects) and sample sizes insufficient for probing the smallest scales, where quasar pairs are rare. There have been several studies probing clustering at sub-Mpc scales, generally finding no excess clustering relative to an extrapolation of the large-scale clustering behavior \\citep[see, e.g.][]{Shen2009b, Padmanabhan2009}. However, these studies have been limited to scales above 100 $\\rm{kpc \\: h^{-1}}$, while several recent studies have managed to probe even smaller scales, where they do indeed find a significant excess \\citep{Hennawi2006, Myers2007II, Myers2008}. In particular, \\citet{Hennawi2006} studied binary quasars from SDSS and 2dF Quasar Survey to compute quasar clustering for scales as small as 20 $\\rm{kpc} \\: h^{-1}$ (comoving), and found significant excess clustering relative to the large scale extrapolation (by an order of magnitude at comoving scales below 100 $\\rm{kpc} \\: h^{-1}$, and growing stronger with decreasing scale). This excess implies that the quasars are more strongly clustered than galaxies at these small scales, supporting the theory that quasar activity is triggered by galaxy interactions. Using the quasar sample from \\citet{Myers2007}, \\citet{Myers2007II} found only a slight excess in small-scale clustering, and put an upper limit for the excess at a factor of 4.3\\underline{+}1.3 for physical scales of $\\sim 28 \\: \\rm{kpc} \\: h^{-1}$. They suggest that the significantly larger excess of \\citet{Hennawi2006} is a result of a selection effect, possibly due to studies tending to target tracers of the Ly$\\alpha$ forest, causing a bias toward $z > 2$, which may be more highly clustered. \\citet{Myers2008} used a complete spectroscopic sample of quasars over physical scales of 23.7-29.9 $\\rm{kpc} \\: h^{-1}$ from SDSS to find an excess clustering factor of $\\sim$4, consistent with the upper limit of \\citet{Myers2007II}, which, while 2$\\sigma$ below the excess found by \\citet{Hennawi2006}, nonetheless supports the general finding of a clustering excess which may be a result of galaxy interactions. In this paper, we use cosmological hydrodynamic simulations which directly model the growth, accretion, and feedback processes of black holes to investigate the properties and underlying causes of black hole clustering. Although the simulation volume limits our analysis to black hole luminosities and host group masses below those typically studied, the self-consistent modeling of black holes allows us to study the clustering behavior without post-processing models. Additionally, the high resolution allows us to investigate clustering behavior at extremely small scales, well below those studied with semi-analytic models, thereby providing a means of using simulations to investigate the observed small-scale excess for the first time, and provide a physical explanation for the underlying cause. In Section 2 we describe the numerical modeling for the black holes formation and accretion (Section 2.1) the simulation parameters used (Section 2.2), the details of the subgroup finder (Section 2.3) and our method of calculating correlation functions (Section 2.4). In Section 3 we investigate the quasar clustering properties at both large and small scales, and we summarize our results in Section 4. \\begin{figure*} \\centering \\includegraphics[width=14cm]{plots/temp4.eps} \\caption{An example of the distribution of black holes in the simulations: The same slice ($2 \\: \\rm{Mpc} \\: h^{-1}$ thick) through the D6 simulation at z=1,2,3,4. The positions of black holes in different luminosities bins ($L < 10^8 L_\\odot$ - Orange; $10^8 L_\\odot < L < 10^9 L_\\odot$ - Pink; $10^9 L_\\odot < L < 10^{10} L_\\odot$ - Blue; $L > 10^{10} L_\\odot$ - Green.) are plotted on top of the gas density distribution (shown in the the gray scale).} \\label{simulationslice} \\end{figure*} ", "conclusions": "In this paper we have investigated the clustering of black holes within hydrodynamic cosmological simulations, its redshift evolution, luminosity dependence, and particularly the small-scale behavior. We have shown that the large scale clustering of black holes traces that of the galaxies within their host groups, and provides a predictor of the typical host mass, which for our simulations is found to be on the order of a few $10^{11} M_\\odot$. Although well below the typically found masses of $\\sim 2 \\times 10^{12}-10^{13} M_\\odot$ \\citep{Lidz2006,Ross2009,Bonoli2009, Shen2009}, this is consistent with our limited simulation volumes which can only follow the growth of the faint-end of quasar population (DeGraf et al. 2010), and cannot follow formation of such massive groups. The typical host group mass shows some evolution with redshift, most significant below $z \\sim 3$, where typical host masses increase by up to a factor 10 (at $z=1$). This low-redshift increase is distinctly luminosity dependent, with the more luminous sources ($L_{\\rm{BH}} > 10^9 L_\\odot$) undergoing the most substantial increase in typical host mass. Overall the evolution of clustering with redshift and luminosity is minor and consistent with current observational constraints (albeit in low luminosity populations this is yet to be fully constrained). The relatively weak dependence found in our simulations is consistent with the complex lightcurves we derive from our direct modeling in which quasar luminosities vary over relatively short timescales for a given host (as regulated by hydrodynamical processes). This is also consistent with the models of \\citet{Lidz2006}. In addition to the large-scale clustering (the 2-halo regime), our simulations allow us to study the small scale clustering (the 1-halo term) of $\\xi_{\\rm{BH}}$. We found that $\\xi_{\\rm{BH,1h}}$ follows a power law behavior all the way to the smallest scales. The clustering of black holes at small scale is unlike that of galaxies (or dark matter). We showed that the 1-halo BH term can be subdivided into two components: 1-subhalo and 2-subhalo. The 1-subhalo term, $\\xi_{\\rm{subgroup,1h}}$, represents the clustering of BHs within a galaxy and 2-subhalo that of BHs occupying different galaxies. We have shown that the 1-subhalo is the one that provides the power law behavior, indicating that galaxies do contain multiple black holes as a result of mergers. These galaxies tend to be the central galaxy within relatively large groups (for our simulation), generally hosting at most a single massive BH with one or more smaller BHs, likely as a result of smaller satellite galaxies merging with the large, central galaxy within the group. In the absence of these multiply-occupied galaxies, $\\xi_{\\rm{BH,1h}}$ and $\\xi_{\\rm{subgroup,1h}}$ exhibit very close agreement, but the inclusion of these merger remnants causes a significant boost in the small-scale BH clustering. This merger-based boost is most significant at low redshift, where typical group size is largest, though we find it in sufficiently massive groups at all redshifts. Though observational limitations make observing these scales difficult, several recent studies have found a small scale excess at scales below $\\sim 100 \\: \\rm{kpc} \\: h^{-1}$ \\citep{Hennawi2006, Myers2008}. The observed excess is in remarkable agreement to the one predicted by our simulations coming from groups approaching $10^{12} M_\\odot$, which host mostly intermediate size black holes. This suggests that multiple black holes co-occupying a subgroup at low redshifts are likely faint(ish) AGNs hosted in Milky Way size halos that have recently undergone merging. We also note that galaxies hosting multiple AGN \\citep{Komossa2003, Gerke2007, Barth2008, Comerford2009b} or inspiralling supermassive black holes \\citep{Comerford2009} have been found in recent studies, further supporting our conclusion of multiply-occupied subgroups. Although we leave more detailed investigation of the small scale BH pairs in our simulations (particularly with regard to the luminosities of inspiralling black holes) for a future work, we note that our finding that multiply-occupied galaxies tend to host a single massive BH with one or more small BHs appears to be in keeping with the observation that most of the inspiralling BH pairs power only a single AGN \\citep{Comerford2009}. Given that, our agreement in small-scale merger-induced boost certainly reinforces the importance of galaxy mergers on the evolution of supermassive black holes. We also note this small-scale excess' sensitivity to the host mass suggests that future small-scale studies may provide a means to constrain the typical mass of merger events between galaxies hosting black holes, with current observational data combined with our simulations suggesting groups with typical masses comparable to those probed in our simulations (from a few $10^{11} M_\\odot$ to $10^{12} M_\\odot$) produce the multiply-occupied galaxies underlying the observed small scale excess. We would like to point out however that there are several aspect of our modeling approach, including numerical issues, in the simulations that potentially affect our results on the small-scale clustering. We have a very simplistic prescription to determine how BHs merge with one another (imposed by the limits on the resolution that can be achieved in these cosmological boxes). The current prescription has a BH pair merge when BHs are separated by less than their smoothing length and if the BHs relative velocity is small compared to the local sound speed. Changes to this prescription could accelerate (postpone) BH mergers, which would result in a suppression (increase) of our small scale clustering signals. It would be desirable to compare our results with other simulations which implement different prescriptions, or in the future to include more direct physical modeling of this region in higher resolution simulations. However, neither of these are currently possible. A numerical issue that may affect the results of our one-halo term is that black holes need to be fixed to potential minima (calculated among the neighboring particles within the smoothing length used for the accretion model) in order to avoid them leaving their subhalo due to numerical N-body noise (and the fact that dynamical friction is hard to calculate for sink particles). However, in some instances this may cause a BH particle in a small subhalo in orbit in a bigger group to 'hop' to the potential minimum of the larger group. This effect may be exacerbated in situations where the small subhalo may be stripped of gas by infalling into a larger one. These effects could artificially increase the number of BHs within large, central halos, thereby boosting small scale clustering. However, when we measure what fraction of BHs appear to 'hop' into the center of groups experiencing an unexpected jump in their position, we find that it is only $\\sim 1-2\\%$. Future simulations and comparison amongst different approaches (once they become available) should of course attempt to characterize these effects more specifically. We further note however, that observational studies have indeed found cases of galaxies hosting multiple BHs \\citep{Comerford2009}, so the existence of a one-subhalo term is expected. Additionally, as seen in Figure \\ref{projectedxi}, the projected clustering of subgroups has a fundamentally different form than the observed quasar clustering. Thus the BHs cannot simply trace their host subgroups/galaxies and still produce the observed small scale excess, but rather a significant one-subhalo term is required to produce the small scale power law behavior. In future work we also plan to simulate larger volumes (which we are starting to be feasible with the most advanced technology) to allow us to study clustering of AGN at larger (mass and length) scales while simultaneously investigating luminosity dependence for brighter sources more directly comparable to current and upcoming observational data, as well as providing increased statistics for the small scale clustering." }, "1005/1005.3656_arXiv.txt": { "abstract": "We present near-UV transmission spectroscopy of the highly irradiated transiting exoplanet WASP-12b, obtained with the Cosmic Origins Spectrograph (COS) on the Hubble Space Telescope (HST). The spectra cover three distinct wavelength ranges: NUVA (2539--2580\\,\\AA); NUVB (2655--2696\\,\\AA); and NUVC (2770--2811\\,\\AA). Three independent methods all reveal enhanced transit depths attributable to absorption by resonance lines of metals in the exosphere of WASP-12b. Light curves of total counts in the NUVA and NUVC wavelength ranges show a detection at a 2.5$\\sigma$ level. We detect extra absorption in the \\ion{Mg}{2} $\\lambda\\lambda$2800 resonance line cores at the 2.8$\\sigma$ level. The NUVA, NUVB and NUVC light curves imply effective radii of 2.69$\\pm$0.24\\,R$_J$, 2.18$\\pm$0.18\\,R$_J$, and 2.66$\\pm$0.22\\,R$_J$ respectively, suggesting the planet is surrounded by an absorbing cloud which overfills the Roche lobe. We detect enhanced transit depths at the wavelengths of resonance lines of neutral sodium, tin and manganese, and at singly ionised ytterbium, scandium, manganese, aluminum, vanadium and magnesium. We also find the statistically expected number of anomalous transit depths at wavelengths not associated with any known resonance line. Our data are limited by photon noise, but taken as a whole the results are strong evidence for an extended absorbing exosphere surrounding the planet. The NUVA data exhibits an early ingress, contrary to model expectations; we speculate this could be due to the presence of a disk of previously stripped material. ", "introduction": "Observations of the transiting extrasolar planets HD209458b and HD189733b revealed an enhanced transit depth at the wavelengths of several UV resonance lines \\citep{vidal03,vidal04,lecavelier10}. These UV lines from the ground state are sensitive probes of the presence of atomic and ionic species. Their presence enhanced the effective radius of the planet during transit, implying the planet is surrounded by an extended cloud of size comparable to or larger than its Roche lobe \\citep{vidal03,vidal04,jaffel07,vidal08}. This was attributed to a hydrodynamic `blow-off' of the planet's outer atmosphere caused by the intense irradiation suffered by this hot Jupiter exoplanet. An alternative explanation in which the planet is surrounded by a cloud of energetic neutral atoms caused by interactions with the host star's stellar wind has, however, been suggested \\citep{ena08,ena10}. WASP-12b is one of the hottest and most irradiated transiting exoplanets and orbits extremely close to a late F-type host star \\citep{hebb2009}. WASP-12b is, therefore, an attractive target to explore the properties of the phenomenon observed in HD209458b, and might yield evidence distinguishing between the suggested underlying causes. The initial UV observations of HD209458b were in the far UV around the Ly\\,$\\alpha$ emission line. The abundance of hydrogen makes this an attractive line to observe, but the temporal and spatial variability of stellar Ly\\,$\\alpha$ emission is a highly undesirable complicating factor. For this reason, and to obtain better signal to noise, we observed WASP-12 in the near-UV where there are many other resonance lines \\citep{morton1991,morton2000}, including the very strong \\ion{Mg}{2} UV resonance lines. This work became possible with the installation of the Cosmic Origins Spectrograph (COS) on the Hubble Space Telescope (HST) reinstating and enhancing our capabilities for UV spectroscopy. ", "conclusions": "\\label{sec:disc} We have performed three independent analyses, each of which suggests absorption in the resonance lines of metals from an extended atmosphere surrounding the transiting planet WASP-12b. In Section~\\ref{charb} we found a deeper transit in the core region of the \\ion{Mg}{2} doublet at the $2.8\\sigma$ level. In Sect.~\\ref{sec_lightcurve}, the transit depths in the NUVA, NUVB, and NUVC wavelength ranges respectively imply effective planet radii of 2.69$\\pm$0.24\\,R$_J$, 2.18$\\pm$0.18\\,R$_J$, and 2.66$\\pm$0.22\\,R$_J$. WASP-12b's optical radius is $R_{\\rm P} = 1.79 \\pm 0.09 \\,R_J$ while the mean Roche lobe radius is 2.36\\,R$_J$ using Paczy{\\'n}ski's (1971) prescription. Table~\\ref{wavelengths} shows that we detect enhanced transit depths at the wavelengths of resonance lines of neutral sodium, tin and manganese, and at singly ionised ytterbium, scandium, manganese, aluminum, vanadium and magnesium. Finally we detect an enhanced transit depth within $0.12$\\AA\\ of a resonance line of doubly ionised europium. We also find the statistically expected number of anomalous transit depths at wavelengths not associated with any known resonance line. Taken as a whole, these results constitute compelling evidence that WASP-12b is surrounded by an exosphere which over-fills the planet's Roche lobe, confirming predictions by \\citet{li2010}. This exosphere is likely composed of a number of elements/ions, including probably \\ion{Na}{1}, \\ion{Mg}{1}, \\ion{Mg}{2}, \\ion{Al}{1}, \\ion{Sc}{2}, \\ion{Mn}{2}, \\ion{Fe}{1}, and \\ion{Co}{1}. The phenomenon found in HD209458b \\citep{vidal03,vidal08} probably occurs generally for hot Jupiter exoplanets. By analogy with HD209458b, and as WASP-12b and its host star are almost certainly predominantly composed of hydrogen, we expect that this exosphere is hydrogen rich. Models by \\citet{yelle} suggest that elements other then H and He should not be present in the upper atmosphere due to the low vertical mixing rate, but this takes Jupiter as the starting point. WASP-12b is extremely close to the host star and consequently the stellar irradiation and tidal effects could induce prodigious mixing, affecting the chemistry of the planet atmosphere. Our detections of several metallic elements and/or ions is certainly consistent with a metal-rich atmosphere for WASP-12b. The most surprising result is provided by the juxtaposition of our data with the optical ephemeris. We took contemporaneous optical photometry with OU-OAM PIRATE \\citep{klb2009} which showed the ephemeris of \\citet{hebb2009} remains accurate. Figure~\\ref{lightcurve} shows the NUVA transit has an early ingress and an egress consistent with the optical ephemeris. In contrast, naive momentum considerations and hydrodynamic simulations would instead suggest that the effect of a diffuse cloud surrounding the planet would be to smear and delay egress while ingress is relatively unaffected, see e.g. Fig.~1 and 2 of \\citet{schneiter2007}. In detail, the shape of the diffuse cloud may well be element/ion dependent since different elements/ions behave differently in the presence of strong radiation pressure. This can explain why we observe different transit shapes in the NUVA region and the other regions. As Fig.~\\ref{spectra} shows, the stellar spectrum in the NUVA region is strongly absorbed by a plethora of lines, dominated those of neutral elements. The NUVC region is also strongly absorbed in the stellar photosphere but predominantly from the \\ion{Mg}{2} doublet. It is presumably the cumulative absorption from many relatively weak spectral lines in the planet's exosphere which creates the excess transit depth in the NUVA region, while Table~\\ref{wavelengths} and Eq.~\\ref{mg2depth} demonstrate that planet's absorption in the NUVC region is associated with the \\ion{Mg}{2} doublet. The \\ion{Mg}{2} ion will experience different forces to neutral atoms in an environment where there is certainly a strong radiation field, and strong and varying large-scale magnetic fields are also likely. The NUVB light curve is least deviant from the optical transit, and this is consistent with the relative dearth of strongly absorbing lines in this spectral window, c.f. Fig.~\\ref{spectra}. We do not have any detailed explanation for the observed early ingress in NUVA, but we speculate the effect could be produced if material is lost from the planet exosphere and forms a diffuse ring or torus around the star enveloping the planet's orbital path, as models suggest \\citep{li2010}. The orbital motion of the planet through this medium might compress the material in front of it. This could increase the opacity of the medium through which the star is viewed immediately before first contact. A void in the medium might be expected to form behind the planet, and consequently the egress is relatively unaffected by the diffuse ring. Our observations demonstrate that COS spectroscopy of transiting exoplanets has the potential to detect many species via transmission spectroscopy, and to measure velocities and deduce spatial distributions. There are now about 40 known transiting exoplanets with orbital periods shorter than that of HD209458b. Many of these transit stars significantly brighter than WASP-12b. COS spectroscopy of brighter examples will allow us probe the exosphere species-by-species examining their density, velocity and spatial distributions. This detailed information should allow us to determine whether these planets really are being photo-evaporated by their host stars, and, if so, to empirically deduce the mass loss rate. We encourage detailed element/ion dependent modeling of the exosphere in the highly irradiated environment of WASP-12b and similar systems, and observations of other similar extrasolar planets. There is a rich new parameter space to explore!" }, "1005/1005.3642_arXiv.txt": { "abstract": "{The Ultra Luminous InfraRed Galaxy (ULIRG) Mrk\\,231 reveals up to seven rotational lines of water (H$_2$O) in emission, including a very high-lying ($E_{\\mathrm{upper}}=640$ K) line detected at a 4$\\sigma$ level, within the Herschel/SPIRE wavelength range ($190 < \\lambda (\\mathrm{\\mu m})< 640$), whereas PACS observations show one H$_2$O line at 78 $\\mu$m in absorption, as found for other H$_2$O lines previously detected by ISO. The absorption/emission dichotomy is caused by the pumping of the rotational levels by far-infrared radiation emitted by dust, and subsequent relaxation through lines at longer wavelengths, which allows us to estimate both the column density of H$_2$O and the general characteristics of the underlying far-infrared continuum source. Radiative transfer models including excitation through both absorption of far-infrared radiation emitted by dust and collisions are used to calculate the equilibrium level populations of H$_2$O and the corresponding line fluxes. The highest-lying H$_2$O lines detected in emission, with levels at $300-640$ K above the ground state, indicate that the source of far-infrared radiation responsible for the pumping is compact (radius$=110-180$ pc) and warm ($T_{\\mathrm{dust}}=85-95$ K), accounting for at least $45$\\% of the bolometric luminosity. The high column density, $N(\\mathrm{H_2O})\\sim5\\times10^{17}$ cm$^{-2}$, found in this nuclear component, is most probably the consequence of shocks/cosmic rays, an XDR chemistry, and/or an ``undepleted chemistry'' where grain mantles are evaporated. A more extended region, presumably the inner region of the 1-kpc disk observed in other molecular species, could contribute to the flux observed in low-lying H$_2$O lines through dense hot cores, and/or shocks. The H$_2$O 78 $\\mu$m line observed with PACS shows hints of a blue-shifted wing seen in absorption, possibly indicating the occurrence of H$_2$O in the prominent outflow detected in OH (Fischer et al., this volume). Additional PACS/HIFI observations of H$_2$O lines are required to constrain the kinematics of the nuclear component, as well as the distribution of H$_2$O relative to the warm dust.} ", "introduction": "One key question in the study of composite infrared (IR) merging galaxies and quasi-stellar objects (QSOs) is what fraction of their luminosity is generated in the nuclear region ($<200$ pc) associated with the Active Galactic Nucleus (AGN) and a possible extreme nuclear starburst, and what fraction arises from a more extended kpc-scale starburst \\citep[e.g.][]{arm07,vei09}. The ULIRG Markarian 231 (Mrk 231) is the most luminous ($L\\sim4\\times10^{12}$ \\Lsun) galaxy in the local Universe ($z<0.1$), and thus provides a unique template for such studies. Since the bulk of the luminosity in ULIRGs arises at far-IR wavelengths, where sub-arc-second resolution observations are not available, an alternative technique is required to constrain the compactness of the far-IR emission and its physical origin. In a previous work based on observations with the \\emph{Infrared Space Observatory (ISO)}, \\citet[][hereafter G-A08]{gon08} have argued that the observation of molecular species such as OH and H$_2$O at far-IR wavelengths is ideal for such a purpose, because their high-lying rotational levels are pumped through absorption of far-IR radiation and the observable excitation is then sensitive to the far-IR radiation density that in turn depends on the compactness of the far-IR continuum source. In addition, these molecular observations shed light on the dominant chemistry in those nuclear regions. G-A08 reported the ISO detection of 3 high-lying H$_2$O lines, relevant upper limits over the entire ISO spectrum, and also high-lying OH lines, indicating the ocurrence of a compact-luminous far-IR component. With their high sensitivity, spectral resolution, and wavelength coverage, the \\emph{Herschel} \\citep{pil10} instruments are ideal for extending our previous study to additional key lines in the far-IR/submillimeter. As part of the HerCULES Key Programme \\citep[see][this volume, hereafter vdW10]{vdw10}, we report in this Letter the Herschel SPIRE/PACS \\citep{gri10,pog10} detection and first analysis of several H$_2$O lines in Mrk 231, which supports the conclusions of G-A08 and gives additional clues to the origin of H$_2$O in this ULIRG. We adopt a distance to Mrk 231 of 192 Mpc ($H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\\Omega_{\\Lambda}=0.73$, and $z=0.04217$). ", "conclusions": "The extreme nature of the nuclear region in Mrk 231 is well illustrated by comparing its SPIRE spectrum with that of the Orion Bar \\citep[][this issue]{hab10}, the prototypical Galactic PDR. The Orion Bar spectrum shows CO lines a factor of $\\gtrsim50$ stronger than the H$_2$O lines, while in Mrk 231 the H$_2$O and CO lines have comparable strengths (vdW10). This contrast will be still higher in the nuclear region, provided that a significant fraction of the CO emission in Mrk 231 arises from a more extended region. Thus the H$_2$O-to-CO line intensity ratios in the SPIRE wavelength range are an excellent diagnostic of extragalactic compact/warm far-IR continuum sources with unusually high amounts of H$_2$O. The above comparison also indicates that the nuclear region of Mrk 231 cannot be interpreted as an ensemble of classical PDRs. Three main scenarios are proposed to explain such high amounts of H$_2$O: $(i)$ widespread shocks/cosmic rays: although the H$_2$O lines peak around the systemic velocity, outflows of $\\sim100$ km s$^{-1}$ are not ruled out by our data, and indeed some indications in the H$_2$O line shapes of systematic motions have been found; an enhanced cosmic ray flux could also have an important impact on the nuclear chemistry. $(ii)$ XDR chemistry: our derived H$_2$O abundance of $\\sim10^{-6}$ is in very good agreement with XDR model results by \\citet[][their Fig. 3, Model 3]{mei05}, as well as with our preliminary estimate of the H$_2$O spatial distribution. $(iii)$ An undepleted chemistry, where H$_2$O that formed on grain mantles is released into the gas phase, as in Galactic hot cores; in support of this scenario, the derived $T_{\\mathrm{dust}}$ in $W_C$ is close to the evaporation temperature of solid H$_2$O. All three scenarios are probably taking place, and the identification of the dominant process requires a multi-species analysis. The nuclear region traced by the high-lying H$_2$O lines has a size similar to the nuclear disk (or outflow) observed at radio wavelengths and H I 21 cm by \\citet[][their Figs. 3 \\& 7]{car98}, suggesting a close physical correspondence. From $H_C$ and $W_C$, the nuclear surface brightness ($\\sim1.5\\times10^{13}$ \\Lsun\\ kpc$^{-2}$) exceeds the highest values attained in starburst on spatial scales $\\gtrsim100$ pc \\citep{meu97,dav07}, while the nuclear luminosity-to-mass ratio ($L/M\\sim4\\times10^3$ \\Lsun/\\Msun) exceeds the limit for a starburst estimated by \\citet{sco03}. From near-IR data, \\citet{dav07} estimated a starburst luminosity from a similarly sized region of $\\lesssim7\\times10^{11}$ \\Lsun; according to the joint luminosity of our $H_C$ and $W_C$, the AGN would account for at least 50\\% of the output power in Mrk 231." }, "1005/1005.1647_arXiv.txt": { "abstract": "\\noindent Numerical simulations show that box-shaped bulges of edge-on galaxies are not bulges: they are bars \\hbox{seen side-on.} Therefore the two components that are seen in edge-on Sb galaxies such as NGC 4565 are a disk and a bar. But face-on SBb galaxies always show a disk, a bar, and a (pseudo)bulge.~Where is the (pseudo)bulge in NGC{\\thinspace}4565? We use archival {\\it Hubble Space Telescope\\/} $H$-band images and {\\it Spitzer Space Telescope\\/} 3.6 $\\mu$m wavelength images, both calibrated to 2MASS $K_s$ band, to penetrate the prominent dust lane in NGC 4565. We find a high surface brightness, central stellar component that is clearly distinct from the boxy bar and from the disk. Its brightness profile is a S\\'{e}rsic function with index $n = 1.55 \\pm 0.07$ along the major axis and 1.33$\\pm$0.12 along the minor axis. Therefore it is a pseudobulge. It is much less luminous than the boxy bar, so the true pseudobulge-to-total luminosity ratio of the galaxy is $PB/T = 0.06 \\pm 0.01$, much less than the previously believed value of $B/T = 0.4$ for the ``boxy bulge''. We infer that published $B/T$ luminosity ratios of edge-on galaxies with boxy bulges have been overestimated. Therefore, more galaxies than we thought contain little or no evidence of a merger-built classical bulge. From a formation point of view, NGC 4565 is a giant, pure-disk galaxy. This presents a challenge to our picture of galaxy formation by hierarchical clustering: it is difficult to grow galaxies as big as NGC 4565 without also making big classical bulges. ", "introduction": "\\pretolerance=15000 \\tolerance=15000 Figure 1 compares the prototypical Sb galaxies NGC 3351 and NGC 4565. NGC 3351 is more nearly face-on and shows three main components -- a bulge, a bar,~and~a~disk. In contrast, NGC 4565 is edge-on; it shows only two components, a box-shaped bulge and disk. As long as we thought that boxy structure was a secondary property of normal bulges, a galaxy morphologist (Sandage 1961) would just use the bulge-to-total luminosity ratio $B/T \\simeq 0.4$ (Simien \\& de Vaucouleurs 1986) to classify NGC 4565 as an Sb. However, we now know that \\phantom{00000000} \\vskip 230pt \\centerline{\\null} \\vskip -6pt \\noindent ``boxy bulges'' are not bulges at all; rather, they are edge-on bars (Combes \\& Sanders 1981). So the SBb galaxy NGC 3351 shows a disk, a bar, and a bulge, but the SBb galaxy NGC 4565 shows only a disk and a bar. Where is the bulge in NGC 4565? Boxy bulges like that in NGC 4565 are a fundamental feature of edge-on disk galaxies (Sandage 1961; Buta \\etal 2007). Their identification as edge-on bars is a well known result. The Combes \\& Sanders (1981) $N$-body model demonstration that bars heat themselves vertically by a combination of buckling instabilities and resonant star scattering has been confirmed \\phantom{00000000} \\clearpage \\noindent and extended many times (Combes et al.~1990; Pfenniger~\\& Norman 1990; Pfenniger \\& Friedli 1991; Raha et al.~1991; Athanassoula \\& Misiriotis 2002; Athanassoula 2005, Shen et al.\\ 2010, and many others). Cylindrical rotation is observed in $N$-body bars and in boxy bulges (Kormendy \\& Illingworth 1982; Jarvis 1990; Shaw, Wilkinson, \\& Carter 1993; Bettoni \\& Galletta 1994; Fisher \\etal 1994; D'Onofrio et al.~1999; Falc\\'on-Barroso et al.~2004; Howard \\etal 2008) but not in classical, elliptical-galaxy-like bulges (Illingworth \\& Schechter 1982; Kormendy \\& Illingworth 1982; Binney \\etal 1990; Verolme et al.~2002; Copin \\etal 2004; Emsellem \\etal 2004); this further cements the connection between boxy bulges and edge-on bars. Finally, a splitting of gas rotation velocities in edge-on boxy bulges (a ``figure 8'' shape of spectral emission lines) also is a robust signature of gas flow in an edge-on bar (Kuijken \\& Merrifield 1995; Merrifield 1996; Merrifield \\& Kuijken 1999; Bureau \\& Freeman 1999). There is little doubt that the ``boxy bulge'' of NGC 4565 is not the real bulge of the galaxy. Does the galaxy contain a bulge at all? That is, does it contain a dense, central component that we would identify as a bulge in addition to the bar if the galaxy were seen face-on? We care for two reasons, both involving galaxy formation: {\\it Background:} Early galaxy evolution was dominated by hierarchical gravitational clustering of density fluctuations that resulted in galaxy collisions and mergers \\hbox{(White \\& Rees 1978);} these scrambled disks into ellipticals (Toomre~1977).~Enormous energy has been invested in studying hierarchical clustering; there is little danger that the picture is fundamentally wrong (Binney 2004). However, it is incomplete. Recent work has established that hierarchical clustering is gradually giving way to a complementary suite of evolution processes that shape {\\it isolated\\/} galaxies. They evolve by rearranging energy and angular momentum; one consequence is the growth of central components that masquerade as classical bulges but that, in general, formed slowly (``secularly'') out of disks (see Kormendy 1993; Kormendy \\& Kennicutt 2004 for reviews). We call them ``pseudobulges'' to distinguish them from merger remnants. They come in at least two varieties. As reviewed above, ``boxy bulges'' are believed to be edge-on bars. Our Galaxy contains one (Dwek \\etal 1995). Another variety is grown out of disk gas that was transported inward by nonaxisymmetries such as bars; we call them ``disky pseudobulges'' here, because they are often highly flattened, but we emphasize that they are not always flat (Kormendy 1993; Kormendy \\& Kennicutt 2004; \\S\\thinspace2 here). Then: {\\it Reason 1:} Confidence in our conclusion that the boxy center of NGC 4565 is an edge-on bar would be increased if we also observed a (pseudo)bulge as we do in face-on galaxies~(Fig.~1). As long as face-on and edge-on galaxies appear to show physical differences, we can't be sure that we understand them. {\\it Reason 2:} If the box in NGC 4565 is a bar, then it is part of the disk and $B/T$ is smaller than we thought. If in addition the galaxy contains a pseudobulge and not a hidden classical bulge, then $B/T$ is even smaller -- possibly zero. This is hard to understand in the context of hierarchical clustering, which essentially always makes substantial bulges in giant galaxies (see, e.{\\thinspace}g., Peebles \\& Nusser 2010; Kormendy \\etal 2010). ", "conclusions": "The interpretation of boxy bulges in edge-on Sb galaxies as bars is more believable if we also find (pseudo)bulges like those associated with bars in face-on Sb galaxies. Our discovery of a pseudobulge in NGC 4565 that is distinct from the boxy bar increases confidence in our picture of secular evolution. Furthermore, $B/T$ ratios in edge-on galaxies with boxy bulges are smaller than previously believed. In NGC 4565, the detection only of a central pseudobulge means that we see no sign of a major merger remnant. Moreover, NGC 4565 rotates at 255 km s$^{-1}$ interior to the outer warp (Rupen 1991). NGC{\\thinspace}4565 has grown very massive while remaining a pure-disk galaxy. We do not know how this can happen in a Universe dominated by the dynamical violence of hierarchical clustering. \\centerline{\\null} \\vskip -30pt \\centerline{\\null}" }, "1005/1005.4062_arXiv.txt": { "abstract": "Recent high resolution observations of the Galactic center black hole allow for direct comparison with accretion disk simulations. We compare two-temperature synchrotron emission models from three dimensional, general relativistic magnetohydrodynamic simulations to millimeter observations of Sgr A*. Fits to very long baseline interferometry and spectral index measurements disfavor the monochromatic face-on black hole shadow models from our previous work. Inclination angles $\\le 20^\\circ$ are ruled out to $3\\sigma$. We estimate the inclination and position angles of the black hole, as well as the electron temperature of the accretion flow and the accretion rate, to be $i={50^\\circ}^{+35{}^\\circ}_{-15{}^\\circ}$, $\\xi={-23^\\circ}^{+97{}^\\circ}_{-22{}^\\circ}$, $T_e=(5.4 \\pm 3.0) \\times 10^{10}$K and $\\dot{M}=5^{+15}_{-2}\\times10^{-9} M_\\odot \\mathrm{yr}^{-1}$ respectively, with 90\\% confidence. The black hole shadow is unobscured in all best fit models, and may be detected by observations on baselines between Chile and California, Arizona or Mexico at $1.3$mm or $.87$mm either through direct sampling of the visibility amplitude or using closure phase information. Millimeter flaring behavior consistent with the observations is present in all viable models, and is caused by magnetic turbulence in the inner radii of the accretion flow. The variability at optically thin frequencies is strongly correlated with that in the accretion rate. The simulations provide a universal picture of the $1.3$mm emission region as a small region near the midplane in the inner radii of the accretion flow, which is roughly isothermal and has $\\nu/\\nu_c \\sim 1-20$, where $\\nu_c$ is the critical frequency for thermal synchrotron emission. ", "introduction": " ", "conclusions": "" }, "1005/1005.1701_arXiv.txt": { "abstract": "We searched for ${\\rm H_{2}O}$ 6(1,6)-5(2,3) maser emission at 22.235 GHz from several Saturnian satellites with the Nobeyama 45m radio telescope in May 2009. Observations were made for Titan, Hyperion, Enceladus and Atlas, for which Pogrebenko {\\it et al.} (2009) had reported detections of water masers at 22.235 GHz, and in addition for Iapetus and other inner satellites. We detected no emission of the water maser line for all the satellites observed, % although sensitivities of our observations were comparable or even better than those of Pogrebenko {\\it et al.}. % We infer that the water maser emission from the Saturnian system is extremely weak, or sporadic in nature. Monitoring over a long period and obtaining statistical results must be made for the further understanding of the water maser emission in the Saturnian system. ", "introduction": "Maser emissions are widely found in celestial objects such as dense cores of molecular clouds and circumstellar envelopes of late-type stars \\citep{rei81}. Masers have been used as probes of gas with the H$_2$ number density of typically $10^4$--$10^{10}$ cm$^{-3}$. For solar system objects, several maser and laser phenomena have been found; $e.g.$, ${\\rm CO_{2}}$ (Venus and Mars: Mumma 1992) and ${\\rm OH}$ (for many comets: {\\it e.g.} Crovisier {\\it et al.} 2002). Each phenomenon would be induced by different physical processes. While a thermal 22.235 GHz water line was possibly detected for comet Hale-Bopp (Bird {\\it et al.} 1997), the first detection of ${\\rm H_{2}O}$ maser in the solar system was reported at the catastrophic impact of comet Shoemaker-Levy9 and Jupiter (Cosmovici {\\it et al.} 1996). This report suggests that such an incident can induce collisional pumping for water masers. Recently, Pogrebenko {\\it et al.} (2009) (abbreviated as POG hereafter) have reported the detections of ${\\rm H_{2}O}$ masers from the Saturnian satellites (Titan, Hyperion, Enceladus and Atlas) with the Medicina 32m and Mets${\\rm \\ddot{a}}$hovi 14m telescopes. This is interesting because, unlike a temporal phenomenon such as the break-up and disruption of a comet, we can perform long period monitoring of ${\\rm H_{2}O}$ emission using ground-based telescopes, space telescopes ({\\it e.g.} Herschel Space Telescope) and spacecrafts. So far, we do not have much knowledge about the maser mechanism in the solar system, although a lot of water maser phenomena are observationally studied for extra-solar objects. The combination of ground, space and in-situ observations would contribute to understand the nature of water maser emission if the presence in the Saturnian system is verified. Therefore, we must accumulate more data of the water maser lines in the Saturnian system. In this letter, we report our trial of detecting water maser satellites with the 45m radio telescope at Nobeyama Radio Observatory (NRO). We observed the major Saturnian satellites for which POG reported the detections, and in addition, we observed a few other inner satellites. ", "conclusions": "Figure 1 shows the obtained spectra (upper two and lower-left panels) and the illustration of the satellite positions (lower-left panel) at the time of observations for Titan, Hyperion and Iapetus. Figure 2 also shows the spectra (upper panels) and illustrations of positions (lower panels) for Enceladus and Atlas. The position of a satellite $\\theta$ is defined as an angle between the line of sight towards Saturn and the line of a satellite and Saturn ($i.e.$ $\\theta=0$ when the object transits on the Saturn seen from the Earth). Table 2 shows a typical 1$\\sigma$ upper limit of the signal considering the factor of a satellite position inside the telescope beam. The conversion efficiency of flux density to antenna temperature (${\\rm T_{A}^{*}}$), 2.8 Jy ${\\rm K^{-1}}$, was used for the calculations. The observational results show the data combined on each day and the entire days during the observation period. POG reported that the water line detections of Titan and Hyperion were $\\sim$30 mK ($3.8 \\sigma$ for 8 hour integration) and $\\sim$50 mK ($4.0 \\sigma$ for 6 hours integration), respectively, which correspond to 300 mJy and 500 mJy, respectively. They found that these emissions were seen on almost the same satellite positions for both objects, and based on these results, they inferred a common mechanism of the emission such as Saturnian magnetosphere bow shock. On 16-18 and 28 May 2009, Titan and Hyperion were near conjunction so that we could observe them simultaneously in the telescope beam. During 16-18 May, especially, these objects were at almost the same position. Therefore, if the maser emissions had been caused by the common mechanism, we would have detected them simultaneously for these two objects. However, our observations this time could not detect the signals of the emissions, and all the data combined also showed no symptom. We estimated typical values of the $3\\sigma$ upper limit for each daily data. We also obtained $3\\sigma$ upper limits during the observations using all the daily data combined. The typical(all) $3\\sigma$ upper limits were 200(60) mJy for Titan and 240(80) mJy for Hyperion, which means that we could certainly detect the line emissions if the levels of flux densities were as high as those of POG. Although POG did not report the water maser emission on Iapetus, we monitored this satellite during our observation period aiming this object at the OFF-point of the position switching observations. We could not find any signal stronger than $\\sim$210(75) mJy ($3\\sigma$ upper limit) . The ${\\rm H_{2}O}$ ice and vapor plume on Enceladus has been reported by the Cassini Ultraviolet Imaging Spectrometer (Hansen {\\it et al.} 2006 and Hansen {\\it et al.} 2008), and the hypothesis that there exists liquid water in the crust has been proposed (Porco {\\it et al.} 2006). As for the water maser emission, the mass ratio of water vapor to ice in the plume is important because a majority of water molecules must be in the form of vapor in order to have maser emission originated from the Enceladus plume. A recent theoretical study indicates that the plume would be dominated by vapor from the thermodynamics perspective (Kieffer {\\it et al.} 2009). The maser line intensity reported by POG was 500 mJy (4.2$\\sigma$), and they estimated the column density of water vapor from the observed maser intensity, which agrees with the column density of the water vapor plume observed in UV ($n{\\rm =1.5 \\times 10^{16}~cm^{-2}}$: Hansen {\\it et al.} 2006). Nevertheless, we did not find appreciable maser emission in our data. The $3\\sigma$ upper limits were 780(540) mJy. Our data were worse compared with those of POG because Enceladus was located around the edge of the telescope beam much of the observational time. However, these results might indicate that the flux of the plume is varying, or the plume is sporadic; like a geyser. POG reported the most certain detections for Atlas. The averaged spectrum showed a peak with 32 mK and S/N=7.0. From the satellite positions, they found that the maser emissions occurred on the trailing side which was several thousand km away from Atlas, and suggested that the disturbance of the Atlas's motion had caused the emission in the edge regions of Saturnian rings A and F. We attempted to verify this subject in our data. We divided the positions of Atlas into 3 parts; $i.e.$, position 1 and 3: positions passing before the geometrical edges of the Saturn's rings seen from the Earth, and Position 2: those after the rings (see the positions of Atlas in Figure 2). For each position, we combined the data, and furthermore, we tested the data on 19 May because Atlas passed by the geometrical edges of the rings ($\\theta=90^{\\circ}$) during the observations. However, we could not have positive results for each case. All the combined data did not show any prominent feature of the emission. The 3$\\sigma$ detection limits for each data set were in the range of 430--520(400) mJy. We also checked several inner satellites which have a diameter larger than $\\sim$10 km; Mimas, Janus, Epimeteus, Prometheus, Pandora and Pan. For all the satellites, the acquired data did not show indicative of the maser emission. The typical value of $3\\sigma$ upper limits were 500(210) mJy. We had an opportunity to observe the Saturnian water maser line at the time of the ring's disappearance; that is, we see the Saturn ring almost in the edge-on view, and we obtained the data which had comparable or even better sensitivities than those by POG for almost of the satellites, though our observations were made for a limited period in 2009 May. If maser emission is stationary in intensity at the levels as those reported by POG, we could detect them. However the results were negative for all of the observed Saturnian satellites. From these results, we conclude that the water maser in the Saturnian system may be sporadic in nature and it is strongly restricted to the time and position of satellites. We have to monitor the satellites for longer periods and to obtain statistical results. These studies would be useful to figure out the water maser emission reported in the Saturnian system. In addition, we should also perform monitoring observations for other icy bodies; Jovian satellites, comets, outer asteroids and Kuiper belt objects would be inside the scope. As suggested by Cosmovici {\\it et al.} (1996), the maser emission may be induced by catastrophic events. Such events like a disruption or eruption can occasionally be found among the solar system objects; ($e.g.$ C/1999 S4 Linear: disruption and dissipation, 29P/Schwassmann-Wachmann: outburst, 7968 Elst-Pizarro: impact or cometary activity (Toth 2000 and Hsieh et al. 2004), etc.). We should not miss events which will occasionally happen and carry out the observations to accumulate the data." }, "1005/1005.3829_arXiv.txt": { "abstract": "We report the first X-ray detection of \\lya\\ emitters at redshift z$\\sim$4.5. One source (J033127.2-274247) is detected in the Extended Chandra Deep Field South (ECDF-S) X-ray data, and has been spectroscopically confirmed as a $z = 4.48$ quasar with $L_X = 4.2\\times 10^{44} \\ergsec$. The single detection gives a \\lya\\ quasar density of $\\sim$ 2.7$^{+6.2}_{-2.2}\\times$10$^{-6}$ Mpc$^{-3}$, consistent with the X-ray luminosity function of quasars. Another 22 \\lya\\ emitters (LAEs) in the central Chandra Deep Field South (CDF-S) region are not detected individually, but their coadded counts yields a S/N=2.4 (p=99.83\\%) detection at soft band, % with an effective exposure time of $\\sim 36$ Ms. Further analysis of the equivalent width (EW) distribution shows that all the signal comes from 12 LAE candidates with EW$_{rest} <$ 400 \\AA, and 2 of them contribute about half of the signal. From follow-up spectroscopic observations, we find that one of the two is a low-redshift emission line galaxy, and the other is a Lyman break galaxy at z = 4.4 with little or no Ly$\\alpha$ emission. Excluding these two and combined with ECDF-S data, we derive a 3-$\\sigma$ upper limit on the average X--ray flux of $F_{0.5-2.0 keV}$ $<$ 1.6 $\\times 10^{-18}$ \\fluxunit, which corresponds to an average luminosity of $\\langle L_{0.5-2 keV} \\rangle$ $<$ 2.4 $\\times 10^{42}$ ergs $s^{-1}$ for z $\\sim$ 4.5 \\lya\\ emitters. If the average X-ray emission is due to star formation, it corresponds to a star-formation rate (SFR) of $<$ 180--530 M$_\\sun$ yr$^{-1}$. We use this SFR$_X$ as an upper limit of the unobscured SFR to constrain the escape fraction of \\lya\\ photons, and find a lower limit of f$_{esc,Ly\\alpha}$ $>$ 3--10\\%. However, our upper limit on the SFR$_X$ is $\\sim$7 times larger than the upper limit on SFR$_X$ on z$\\sim$ 3.1 LAEs in the same field, and at least 30 times higher than the SFR estimated from \\lya\\ emission. From the average X-ray to \\lya\\ line ratio, we estimate that fewer than 3.2\\% (6.3\\%) of our LAEs could be high redshift type 1 (type 2) AGNs, and those hidden AGNs likely show low rest frame equivalent widths. ", "introduction": "The LAE candidates were selected with narrowband imaging of the GOODS CDF-S (RA 03:31:54.02, Dec -27:48:31.5, J2000) at the Blanco 4m telescope at Cerro Tololo InterAmerican Observatory (CTIO) with the MOSAIC II camera. Three 80 \\AA\\ wide narrowband filters (NB656, NB665 and NB673) were utilized to obtain deep narrowband images (Finkelstein et al. 2008, 2009c). The LAE candidates are selected based on a 5 $\\sigma$ detection in the narrowband, a 4 $\\sigma$ significant narrowband flux excess over the broad band continuum image (here, an R band image from the ESO Imaging Survey [EIS], Arnouts et al. 2001), a factor of 2 ratio of narrowband flux to broadband flux density, and no more than 2 $\\sigma$ significant flux in the EIS-B band. Candidates with GOODS B-band coverage were further examined in the GOODS B-band image, and those with significant B-band detections were excluded. These conditions are satisfied by 113 LAE candidates with the \\textit{Chandra} CDF-S and ECDF-S coverage, including 4 in the NB656 filter\\footnote{The NB656 data was much shallower than the other two bands, thus the galaxies were selected in a different way (see Finkelstein et al. 2008) - we search for the NB656 candidates from the positions of galaxies which were detected in GOODS V-band but not in GOODS B-band. Thus, we were only able to select galaxies over the GOODS region, which is why only four objects were selected. The other two catalogs consist of all selected candidates over the overlap region between the MOSAIC image and the ESO Imaging Survey, which consists of a much larger area. } (% Finkelstein et al. 2008), 39 in NB665, and 81 in NB673 (% including 11 that were detected in both NB665 and NB673). The equivalent widths (EWs) of our LAEs were calculated from our narrowband and EIS-R broadband data. Finkelstein et al (2008, 2009c) have previously studied the 14 objects from this sample that lie within the GOODS \\textit{HST} field. For these sources, we choose the deeper GOODS V-band to calculate the EWs. The 2 Ms \\textit{Chandra X--Ray Observatory} ACIS (Advanced CCD Imaging Spectrometer) exposure of the CDF-S is composed of 23 individual ACIS-I observations. We downloaded the raw data from the \\textit{Chandra} public archive and reduced the data using the \\textit{Chandra} Interactive Analysis of Observations software version 4.0 (CIAO4.0). Each observation was filtered to include only standard \\textit{ASCA} event grades 0, 2, 3, 4, 6. Cosmic ray afterglows, ACIS hot pixels, and bad pixels were removed, along with all data taken during high background time intervals. All exposures were then added to produce a combined event file with a net exposure of 1.9 Ms. The \\textit{Chandra} exposure of the ECDF-S is composed of 9 individual ACIS-I observations obtained in 2004, covering $\\sim$ 0.3 deg$^2$ with four pointings. We reprocessed the X-ray raw data of the four pointings separately. The averaged net exposure per pointing at ECDF-S was 238 ks. The aspect offset\\footnote{http://cxc.harvard.edu/cal/ASPECT/fix\\_offset/fix\\_offset.cgi} of both CDF-S and ECDF-S data was examined and no offset above 0.1\\arcsec\\ was found in either field. We used the published X-ray source catalogs of the 2-Ms CDF-S (Luo et al. 2008) and the 240-ks ECDF-S (Lehmer et al. 2005) in the following source-match and source-mask processes. ", "conclusions": "Our work shows that X-ray observation is an effective method to identify AGN, as well as foreground objects in LAE samples. One X-ray detected LAE is spectroscopically confirmed as a type 1 quasar at z = 4.5. A stack of 22 other LAEs in the CDF-S field yields a marginal detection. However, two of these 22 sources contribute about half of the stacked X-ray signal, and these two were found to be a foreground interloper and a LBG at z=4.4 without strong \\lya\\ emission. The mean flux of the remaining 20 sources, while positive, is not significantly different from zero. Including the ECDFS data, we obtain a $3\\sigma$ upper limit on the average X-ray luminosity of 2.4 $\\times 10^{42}$ erg s$^{-1}$. Compared to their average \\lya\\ luminosity, we estimate that that fewer than 3.2\\% (6.3\\%) of our LAEs could be high redshift type 1 (type 2) AGNs, and those hidden AGNs might show low EW$_{rest}$. Using the relationship of X-ray emission and star-forming activity from low redshift star-forming galaxies, we obtained an upper limit on the unobscured SFR of SFR$<$ 180-530 M$_\\sun$ yr$^{-1}$. Compared to the SFR estimated from their average \\lya\\ luminosity, we find a lower limit on the escape fraction of \\lya\\ photons, f$_{esc,Ly\\alpha} >$ 3-10\\%. Doubling the depth of CDF-S X-ray observations is planned in 2010 and 2011 (see \\textit{Chandra} Electronic Bulletin 89). This will strengthen the power of X-ray diagnostics of LAEs, especially for revealing their unobscured SFR, for the new discovery of \\lya\\ selected quasars and weak AGN, and for excluding the low-redshift contamination." }, "1005/1005.2191_arXiv.txt": { "abstract": "We present new moderate-resolution, far-ultraviolet spectra from the {\\it Hubble Space Telescope}/Cosmic Origins Spectrograph (HST/COS) of the BL\\,Lac object 1ES\\,1553$+$113 covering the wavelength range $\\rm 1135~\\AA<\\lambda<1795~\\AA$. The data show a smooth continuum with a wealth of narrow ($b<100$ \\kms) absorption features arising in the interstellar medium (ISM) and intergalactic medium (IGM). These features include 41 \\Lya\\ absorbers at $00.395$ based on a confirmed \\Lya$+$\\OVI\\ absorber and $z_{\\rm em}>0.433$ based on a single-line detection of \\Lya. The current COS data are only sensitive to \\Lya\\ absorbers at $z<0.47$, but we present statistical arguments that $z_{\\rm em}\\la0.58$ (at a $1\\sigma$ confidence limit) based on the non-detection of any \\Lyb\\ absorbers at $z>0.4$. ", "introduction": "The current interpretation of BL\\,Lac objects \\citep{Ghisellini85} is that they are active galactic nuclei (AGN) with a strongly relativistic jet pointed toward our line of sight. As such, any line emission or accretion disk features seen in most other types of AGN could be masked by the bright jet if present in BL Lac objects. Their spectra usually show a featureless power-law continuum extending from radio to X-ray wavelengths. This spectral characteristic makes BL\\,Lac objects ideal for observing intervening absorption features arising in the interstellar medium (ISM) and intergalactic medium (IGM). Since their continuum is easily defined, they make excellent targets for studying weak metal-line systems and low-contrast, highly thermally broadened \\HI\\ absorbers \\citep[e.g.,][]{Richter04,Lehner07,Danforth10}. The BL\\,Lac object 1ES\\,1553$+$113 shows the characteristic featureless power-law spectrum and is one of the brightest known sources of extragalactic high-energy radiation from X-rays up to VHE (TeV) photons \\citep{CostamanteGhisellini02}. However, the featureless spectrum makes it difficult to determine the redshift of the object and hence its luminosity. Indirect methods have given a wide range of limits for the redshift of 1ES\\,1553$+$113; the nondetection of a host galaxy gave limits from $z_{\\rm em}>0.09$ to $z_{\\rm em}>0.78$ \\citep{HutchingsNeff92,Scarpa00,Urry00,Carangelo03,Sbarufatti06,Treves07}. The shape of the $\\gamma$-ray spectrum observed by the {\\it Fermi Observatory} and ground-based VHE detectors (HESS, MAGIC) constrains the redshift to values from $z_{\\rm em}<0.4$ to $z_{\\rm em}<0.8$ \\citep{Ahronian06,Albert07,MazinGoebel07,Abdo10} based on assumptions about the intrinsic spectral energy distribution (SED) and pair-production interactions with the cosmic infrared background. The only direct redshift determination \\citep[$z_{\\rm em}=0.36$;][]{MillerGreen83} was based on a spurious feature in low-resolution UV spectra from the {\\it International Ultraviolet Explorer} (IUE). The detection was later retracted \\citep{FalomoTreves90}, but the erroneous redshift value lives on. 1ES\\,1553$+$113 is of interest as a bright background continuum source for detecting intergalactic absorption along the sight line. Bright X-ray sources are especially valuable for potentially detecting the long-predicted \\OVII\\ and \\OVIII\\ tracers \\citep{Bregman07} of intergalactic gas at $T=10^6-10^7$~K. Even for a bright X-ray source, the required integration times would be very long. However, a sufficiently long IGM pathlength provided by a bright high-$z$ target would make the required observing time investment more attractive. In this paper, we present the first medium-resolution far-UV spectroscopic observations of 1ES\\,1553$+$113 including {\\it Hubble Space Telescope}/Cosmic Origins Spectrograph \\citep[HST/COS][]{Green10,Osterman10} observations ($\\lambda=1135-1795$~\\AA) as well as archival data at $905-1187$~\\AA\\ from the {\\it Far Ultraviolet Spectroscopic Explorer} \\citep[\\FUSE;][]{Moos00,Sahnow00}. We confirm the featureless power-law nature of the spectrum over this wavelength range. Absorption is seen in 42 intervening systems including 41 \\Lya\\ absorbers and six metal-line systems. The frequency of IGM absorbers is consistent with larger surveys using \\FUSE\\ and HST/STIS data (Danforth \\& Shull 2005, 2008; hereafter DS08), and the systems are spread across the entire redshift range covered by the combined COS/\\FUSE\\ dataset ($z\\la0.47$). The observations and data reduction techniques are discussed in \\S2, and we present a preliminary catalog of absorption lines in \\S3. Our conclusions are presented in \\S4. ", "conclusions": "The BL\\,Lac object 1ES\\,1553$+$113 is one of the brightest objects in the sky in $\\gamma$-rays, as well as being a notable UV and X-ray source. However, the AGN emission is that of a relativistic jet aligned closely with our line of sight and, like most such objects, has no intrinsic emission or absorption features at any wavelength. This featureless, power-law continuum is ideal for measuring intervening IGM features that are weak and broad, such as thermally-broadened \\Lya\\ systems. However, the lack of intrinsic features makes constraining the redshift of the object difficult. We present unprecedented high-quality far-UV HST/COS and \\FUSE\\ spectra of the BL\\,Lac object 1ES\\,1553$+$113 at spectral resolution 15-20 \\kms. These data show 42 intervening IGM absorbers, 41 of which are detected in \\Lya, and 15 in \\Lyb\\ and/or metal lines. The richest absorption system in the line of sight is a trio of \\Lya\\ absorbers at $z\\approx0.188$ covering $\\sim1000$ \\kms\\ of velocity space. Several metal ions are also detected in these systems, including \\OVI, \\NV, and \\CIII. However, neither \\SiIV\\ nor \\SiIII\\ is detected in any of the systems. The \\CIII/\\SiIII\\ ratio implies a (C/Si) abundance at least four times the solar value, while a high \\NV/\\OVI\\ value suggests an overabundance of N as well. A detailed analysis of the physical conditions in this system can be found in \\citet{Yao10}. The redshift of 1ES\\,1553$+$113 has never been determined directly, and the only limits placed on it come from indirect means such as the shape of the $\\gamma$-ray spectrum and the lack of an AGN host galaxy in deep optical images. A strong \\Lya$+$\\OVI\\ absorber at $z=0.3951$ gives the first direct lower limit to the redshift of the object. Two weaker \\Lya\\ absorbers at $z=0.4063$ and $z=0.4326$ give slightly higher estimates of the redshift, but these weak \\Lya\\ lines are not confirmed by additional line detections. These lower limits are consistent with most previous measurements via optical non-detections of host galaxies and $\\gamma$-ray SED constraints. \\citet{Abdo10} derive $z_{\\rm em}=0.75^{+0.04}_{-0.05}$ based on the latest {\\it Fermi} and TeV $\\gamma$-ray SED measurements, considerably higher than our intervening absorber upper limits. COS far-UV spectra are not sensitive to \\Lya\\ absorbers at $z>0.47$, but the G160M grating has some sensitivity to intervening \\Lyb\\ and \\OVI\\ absorbers out to $z\\sim0.75$. If the \\citet{Abdo10} redshift estimate were accurate, we would expect to find $\\sim8$ \\Lyb\\ absorbers at $0.471500$~\\AA\\ with ambiguous line identifications that could potentially be \\Lyb\\ systems at $z>0.47$. While these systems are individually suggestive, we find nowhere near the number of absorbers predicted statistically. We conclude that the redshift of 1ES\\,1553$+$113 is not much higher than $z\\approx0.45$. 1ES\\,1553$+$113 is one of the brightest X-ray sources on the sky and has been suggested as a sight line that could be efficiently probed for WHIM absorption in \\OVII. The combined \\OVI\\ column density in the three absorbers at $z\\sim0.19$ is $\\sim2\\times10^{14}\\rm~cm^{-2}$. Spectrographs on modern X-ray observatories are sensitive to $\\log\\,N_{\\rm OVII}\\ga15.5$. If the temperature of any of these \\OVI\\ systems is high enough, sufficiently long Chandra and/or XMM/Newton observations may reveal a \\OVII\\ counterpart to these \\OVI\\ absorbers that could constrain the long-sought X-ray WHIM \\citep{Bregman07}. However, at the observed Li-like (\\OVI) oxygen column density, $\\log\\,N_{\\rm OVI}\\approx14.3$ in the trio of absorbers, the expected column densities of He-like (\\OVII) and H-like (\\OVIII) oxygen are probably just below the detectability levels of {\\it Chandra} and {\\it XMM}. Recent analysis of stacked X-ray absorption data \\citep{Yao09} at the known IGM redshifts of \\OVI\\ absorbers finds no evidence for \\OVII\\ or \\OVIII\\ absorbers to a limit $N_{\\rm OVII}/N_{\\rm OVI}<10$. Therefore, the $z\\approx0.19$ absorbers might have X-ray column densities $\\log\\,N_{\\rm OVII}\\leq15.3$, just below the limits of current X-ray observatories. Finally, these observations showcase the powerful new tool available to astronomers for probing the low-redshift IGM. COS is $10-20$ times more sensitive in the far-UV to point sources than previous instruments on HST. An additional six orbits of COS observations are planned for 2010, which should improve the S/N of the combined dataset by a factor of $\\sim\\sqrt{3}$. Improving the data quality will help confirm or refute some of the tentative line identifications from this paper and will undoubtedly uncover additional weak absorbers. We will place further constraints on [C/Si] and [N/O] in the $z=0.188$ system, identify new broad, shallow \\Lya\\ absorbers, and investigate possible high-$z$ \\Lyb\\ systems with our new Cycle 18 observations. \\medskip \\medskip It is our pleasure to acknowledge the many thousands of people who made the HST Servicing Mission 4 the huge success that it was. We furthermore thank Steve Penton, St\\'ephane B\\'eland, and the other members of the COS ERO and GTO teams for their work on initial data calibration and verification. C. D. wishes to acknowledge a fruitful discussion with members of the KIPAC consortium. This work was supported by NASA grants NNX08AC146 and NAS5-98043 to the University of Colorado at Boulder." }, "1005/1005.2969_arXiv.txt": { "abstract": "{ High-resolution far-infrared and sub-millimetre spectroscopy of water lines is an important tool to understand the physical and chemical properties of cometary atmospheres. We present observations of several rotational ortho- and para-water transitions in comet \\garradd{} performed with HIFI on {\\it Herschel}. These observations have provided the first detection of the $2_{12}$--$1_{01}$ (1669 GHz) ortho and $1_{11}$--$0_{00}$ (1113 GHz) para transitions of water in a cometary spectrum. In addition, the ground-state transition $1_{10}$--$1_{01}$ at 557 GHz is detected and mapped. By detecting several water lines quasi-simultaneously and mapping their emission we can constrain the excitation parameters in the coma. Synthetic line profiles are computed using excitation models which include excitation by collisions, solar infrared radiation, and radiation trapping. We obtain the gas kinetic temperature, constrain the electron density profile, and estimate the coma expansion velocity by analyzing the map and line shapes. We derive water production rates of $\\range \\times10^{28}\\ \\mathrm{s}^{-1}$ over the range $r_\\mathrm{h}=1.83$--$1.85$ AU.} ", "introduction": "Comets spend most of their lifetime in the outer Solar System and therefore have not undergone considerable thermal processing. Line emission is useful to study the physical and chemical conditions of cometary atmospheres, and their relation to other bodies in the Solar System \\citep{2002EM&P...90..323B,2004come.book..391B}. Water molecules in cometary atmospheres are excited due to collisions with other molecules and radiative pumping of the fundamental vibrational levels by the solar infrared flux. The \\transi{} ortho-water transition at 557 GHz is one of the strongest lines in cometary comae, but it cannot be detected directly from the ground due to absorption in the Earth's atmosphere \\citep{1987A&A...181..169B}. Water vapour production has been estimated previously from the ground through measurements of its photodissociation product, the OH radical \\citep[see e.g.][]{1982come.coll..433A}, and water high vibrational bands \\citep{2004come.book..391B}. The \\transi{} rotational transition of ortho-water at 557 GHz has been observed using heterodyne techniques by the Submillimeter Wave Astronomical Satellite (SWAS) \\citep{2000ApJ...539L.151N,2001Icar..154..345C}, and later with Odin \\citep{2003A&A...402L..55L,2007P&SS...55.1058B,2009A&A...501..359B}. ESA's {\\it Herschel} Space Observatory was successfully launched on May 14, 2009 and entered a Lissajous orbit around the $L_2$ Lagrangian point \\citep{2010Herschel}. The Heterodyne Instrument for the Far-Infrared (HIFI) onboard {\\it Herschel} has continuous coverage in five frequency bands in the 480--1150 GHz range, and an additional dual frequency band covering the 1410--1910 GHz range that are not observable from the ground \\citep{2010HIFI}. HIFI's submillimetre frequency coverage is of great importance to observe water vapour in Solar System objects such as cometary comae with unique sensitivity and the required high spectral resolution to resolve the emission lines. Comet \\garradd{} is a long-period comet ($P = 1.9\\times10^5$) with a highly eccentric orbit ($e=0.99969$). It passed perihelion on June 23, 2009 at a distance of 1.7982 AU from the Sun and was observed with HIFI in July 2009 as part of the {\\it Herschel} guaranteed time key project ``Water and related chemistry in the Solar System'' \\citep{2009P&SS...57.1596H}. In this letter, we describe the observations of water and the analysis of part of the data set. Water production rates derived from our radiative transfer models are presented. ", "conclusions": "\\label{sec:conclusions} The {\\it Herschel} Space Observatory provides unique new capabilities for the detection of water in the Solar System. HIFI's spectral range and high resolution allowed for the direct detection of several water lines almost simultaneously. High spectral resolution is crucial to resolve the line shape and asymmetries due to self-absorption. On 20-27 July 2009, comet \\garradd{} was observed with HIFI. The high-resolution spectra of HIFI allows us to detect for the first time several rotational water lines in cometary spectra. A water production rate of $\\sim 2\\times10^{28}\\ \\mathrm{s}^{-1}$ was derived at heliocentric distance of 1.8 AU using radiative transfer numerical codes which include collisional effects and infrared fluorescence by solar radiation. In future studies, HIFI will be able to detect water isotopes, and determine the D/H ratio in active comets." }, "1005/1005.5252_arXiv.txt": { "abstract": "{Non-blazar AGN have been recently established as a class of gamma-ray sources. M87, a nearby representative of this class, show fast TeV variability on timescales of a few days.} {We suggest a scenario of flare gamma-ray emission in non-blazar AGN based on a red giant interacting with the jet at the base.} {We solve the hydrodynamical equations that describe the evolution of the envelope of a red giant blown by the impact of the jet.} {If the red giant is at least slightly tidally disrupted by the supermassive black hole, enough stellar material will be blown by the jet, expanding quickly until a significant part of the jet is shocked. This process can render suitable conditions for energy dissipation and proton acceleration, which could explain the detected day-scale TeV flares from M87 via proton-proton collisions. Since the produced radiation would be unbeamed, such an events should be mostly detected from non-blazar AGN. They may be frequent phenomena, detectable in the GeV-TeV range even up to distances of $\\sim 1$~Gpc for the most powerful jets. The counterparts at lower energies are expected to be not too bright.} {M87, and nearby non-blazar AGN in general, can be fast variable sources of gamma-rays through red giant/jet interactions.} ", "introduction": "\\label{intro} Active galactic nuclei (AGN) are believed to be powered by an accreting supermassive black hole (SMBH) in the center of a galaxy, a significant fraction of AGN show powerful jets, supersonic relativistic flows, on small (sub parsec) and large (multi hundred kpc) scales. \\citep[e.g.][]{bbr84}. These AGN are characterized by nonthermal emission extending from radio to high energy gamma-rays. This radiation comes from an accretion disc and from two relativistic jets that are launched close to the SMBH in two opposite directions. The emission associated to the accretion process can be generated by thermal plasma in the form of an optically-thick disc under efficient cooling \\citep[e.g.,][]{s72,ss73}, or as an optically-thin corona \\citep[e.g.,][]{bkb77,lt79}. The emission from the jets is non thermal and comes from a population of relativistic particles accelerated for instance in strong shocks, although other scenarios are possible as well \\citep[see, e.g.,][]{recon,na07,rbd07,ra08}. This non-thermal emission is thought to be produced through synchrotron and inverse Compton (IC) processes \\citep[e.g.][]{gmt85}, although hadronic models have been also considered in the past \\citep[e.g.,][]{m93,ah00,mp01,a02}. The existence of a stellar clustering in the central regions of AGN, possibly down to very small distances from the central SMBH \\citep[e.g.][]{pens88}, implies that the interaction between a star and the jet should eventually happen. The gamma-ray production due to the interaction between an obstacle and an AGN jet has been studied in a number of works. For instance, \\cite{dl97} suggested the high-energy radiation produced by a beam of relativistic protons impacting with a cloud of the broad-line region (BLR). The gamma-ray emission from one or many clouds from the BLR interacting with a hydrodynamical jet recently has bin analyzed by \\cite{abr10}. The radiation from the interaction between a massive star and an AGN jet was studied in \\cite{bp97}. Namely they suggested that the jet interacts with stellar winds of massive stars, in their model they assume that the source of gamma-rays is moving with a relativistic speed, therefore the radiation is Doppler boosted. The main radiation mechanism in this scenario is related to the development of the pair cascade in the field of the radiation of the massive star. In this work, we study the interaction of a red giant (RG) star with the base of the jet in AGN and their observable consequences in gamma rays. We focus here on the case of M87, a nearby non-blazar AGN that presents very high-energy recurrent activity with variability timescales of few days \\citep{ah06,mag,vvhm09,ver10}. In the framework presented here, the jet impacts the RG envelope, already partially tidaly disrupted by the gravitational field of the central SMBH. The RG envelope is blown up, forming a cloud of gas accelerated and heated by jet pressure. The jet base is likely strongly magnetized \\citep[e.g.,][]{kbvk07,bk08}. The jet flow affected by the impact with the RG envelope can be a suitable region for particle acceleration, and a significant fraction of the involved magnetic and kinetic energy of the jet can be transferred to protons and electrons. Although electrons may not able to reach TeV emitting energies because of the expected large magnetic fields, protons would not suffer from this constraint. These protons could reach the star blown material, and optically-thick proton-proton ($pp$) interactions could lead to significant gamma-ray production in the early stages of the cloud expansion. Unlike in \\cite{bp97}, we deal with solar-mass-type stars instead of the more rare high-mass stars, study the RG atmosphere-jet interaction, and follow the hydrodynamical evolution of the cloud. Finally, we do not introduce any beaming factor to the radiation, since in our scenario most of the emission is produced when the cloud has not been significantly accelerated, Doppler boosting being therefore negligible. ", "conclusions": "The total jet luminosity can be inferred from observations using Eq.~(\\ref{egcn}): \\begin{equation} L_{\\rm j}=8\\times 10^{44}\\,L_{\\gamma,41}^{5/8} M_{\\rm BH,9}^{2/3}\\theta_{-1}\\, \\eta_{-1}^{-5/8}t_{\\rm v,5}^{-1/8}M_{\\rm RG}^{-1/6}\\,\\mbox{erg s}^{-1}. \\label{etj} \\end{equation} This formula weakly depends on the observables, being almost insensitive to $M_{\\rm RG}$, on the other hand hard to estimate. This provides therefore a quite robust estimate of the jet luminosity with $\\eta$ as most unknown parameter. Actually, if $L_{\\rm j}$ were known, then $\\eta$ could be also estimated. For the most powerful jets, $L_\\gamma$ would be limited by the jet size becoming $L_\\gamma=\\chi\\eta L_{\\rm j}$. Taking for instance $L_{\\rm j}\\sim 10^{47}$~erg~s$^{-1}$, $L_\\gamma$ could be as high as $\\approx 2\\times 10^{45}\\,\\eta_{-1}$~erg~s$^{-1}$. An improvement of a factor of several in the VHE sensitivity (e.g. through the forthcoming the Cherenkov Telescope Array -CTA-) would test our gamma-ray predictions for the whole RG-jet interaction process, including the early cloud expansion phase, allowing for a detailed study of the involved (magneto)hydrodynamics, particle acceleration, and radiation. We remark that, if a detectable gamma-ray flare with a duration of few days were to be produced in M87, in particular through $pp$ interactions, the cloud should have a mass of $\\sim 10^{28}$~g. Such a massive cloud cannot acquire a large speed in the jet direction at the times when $pp$ collisions are an efficient gamma-ray emitting mechanism, and therefore the emission will not suffer significant Doppler boosting. In the case of a lighter cloud, large Lorentz factors can be achieved, but then $pp$ interactions will be inefficient producing gamma-rays, the probability to detect a flare lower due to beaming, and the duration of the event shorter than observed because of faster expansion and beaming. { Coming back to the question of cloud mass, we note that to extract a cloud with a mass $>10^{28}$ g, a more powerful jet than in M87, for similar jet-RG interaction conditions, would be required (see Eq.(\\ref{mb})). } An important question is whether there are enough RGs in M87 at the relevant jet scales. The model presented here would require few interactions per year to explain the observations in M87. Since the typical duration of the RG-jet interaction is of about 3--4~days, the RG filling factor should be $\\Upsilon\\sim 4/365\\approx 10^{-2}$. With a jet volume at the relevant scales of $\\sim \\pi\\theta^2 z_{\\rm T}^3/3$, the density of RGs in the region should be $\\sim \\Upsilon/V\\sim 2\\times 10^6$~pc$^{-3}$ for M87. Unfortunately, no direct information is available on the density of stars in the vicinity of the SMBH in M87. The stellar mass in a sphere with a radius of 80~pc is estimated in $2\\times 10^8 M_{\\odot}$ \\citep[e.g.][]{gt09}, and these observational data should be extrapolated four orders of magnitude down to $\\sim 0.01$~pc. Thus, depending on the assumed extrapolation law, the number of RGs in the vicinity of the SMBH may or may not be enough. It is worth noting that a dense stellar cluster near the SMBH could be behind the broad-line region in AGN as produced by the blown-up atmosphere of red dwarfs, which would imply the presence of numerous RGs in the center of AGN \\citep{pens88}. In addition, studies of the possible stellar density profiles in the vicinity of the SMBH in AGN \\citep{bkck82,mcd91} show that densities like the required one ($\\sim 2\\times 10^6$~pc$^{-3}$) could be achieved. The observation of VHE flares could be already an indication that enough RGs are present near the SMBH in M87. Interestingly, RG/jet interactions are expected to be transient phenomena. At higher jet heights, although many RGs could be simultaneously present in the jet rendering rather continuous emission, the much more diluted jet would not remove a significant amount of material from the star and the effective cross section of the interaction would be just $R_{\\rm RG}$, yielding a low energy budget for such a multiple interaction events. The scenario presented here, adopted to explain the day VHE flares observed from M87, could also be relevant in other non-blazar AGN. For blazar sources the beamed emission would overcome the RG-jet interaction, expected to be weakly beamed due to moderate $v_{\\rm z}$-values. For instance, the closest AGN, the radio galaxy Cen~A, at $\\sim 3.8$~Mpc distance \\citep{r04}, could also show detectable flare like emission. At present, persistent faint VHE emission has been detected \\citep{ah09} with $L_{\\gamma}=2.6\\times 10^{39}\\,\\mbox{erg~s}^{-1}$. Accounting for the black hole mass of this AGN, $M_{\\rm BH}=5.5\\times 10^7\\,M_{\\odot}$, taking the observed VHE luminosity as a reference, and assuming $t_{\\rm v}\\sim 1$~day, one derives implementing Eq.~\\ref{etj} a jet luminosity $L_{\\rm j}=1.2\\times 10^{42}\\,\\mbox{erg~s}^{-1}$, a rather modest value. Therefore, it cannot be excluded that RG-jet interactions may contribute to the VHE radiation detected from Cen~A, or that transient activity due to RG-jet interactions may be observed from this source. Another case, the radio galaxy NGC1275, at a distance of 73~Mpc \\citep{hwb09}, shows variable behaviour in GeV \\citep{fer09}. The GeV luminosity is of about $2\\times 10^{43}$~erg~s$^{-1}$. Using Eq.~(\\ref{etj}), we can estimate the power of the jet as $5\\times 10^{44}$~erg~s$^{-1}$ adopting an $M_{BH}=10^8\\,M_\\odot$. In the case of NGC~1275 the shocked cloud would be optically thick at the luminosity peak, implying significant attenuation of the TeV emission through photon-photon absorption with a cutoff around 50~GeV. At farther distances, the strong jet luminosity dependence $L_{\\gamma}\\propto L_{\\rm j}^{1.6}$ implies that FR~II sources with say $L_{\\rm j}\\sim 10^{46}$~erg~s$^{-1}$ may be still detectable up to distances of $\\sim 0.5$~Gpc (internal absorption should be included in these cases; see \\citealt{akc08}). Also, the luminosity in the range 0.1--100~GeV would also be significant unless there is a strong low-energy cutoff in the proton spectrum. Therefore, Fermi may detect day-long GeV flares originated due to RG-jet interactions from FR~II galaxies up to distances of few 100~Mpc. Summarizing, GeV and TeV instrumentation can potentially detect a number of RG-jet interactions per year taking place in nearby FR~II and very nearby FR~I galaxies, with the most powerful events being detectable up to 1~Gpc." }, "1005/1005.0641_arXiv.txt": { "abstract": "{We introduce two new methods that are designed to improve the realism and utility of large, active region-scale 3D MHD models of the solar atmosphere. We apply these methods to RADMHD, a code capable of modeling the Sun's upper convection zone, photosphere, chromosphere, transition region, and corona within a single computational volume. We first present a way to approximate the physics of optically-thick radiative transfer without having to take the computationally expensive step of solving the radiative transfer equation in detail. We then briefly describe a rudimentary assimilative technique that allows a time series of vector magnetograms to be directly incorporated into the MHD system. ", "introduction": "In this paper, we briefly summarize our efforts to improve our models of quiet Sun and active region magnetic fields in computational domains that include the upper convection zone, photosphere, chromosphere, transition region and low corona within a single computational domain. Our goal is similar to that presented in \\citet{Abbett07} --- that is, to develop the techniques necessary to efficiently simulate the spatially and temporally disparate convection zone-to-corona interface over spatial scales sufficiently large to accommodate at least one active region. The advantage of this type of single-domain modeling is clear. For example, evolving a turbulent convection zone and corona simultaneously in a physically self-consistent way allows for the quantitative study of important physical processes such as flux emergence, submergence and cancellation; the transport of magnetic free energy and helicity into the solar atmosphere; the generation of magnetic fields via a convective dynamo; and the physics of coronal heating. However, this approach is challenging. The computational domain is highly stratified --- average thermodynamic quantities change by many orders of magnitude as the domain transitions from a relatively cool, turbulent regime below the visible surface, to a hot, magnetically-dominated and shock-dominated regime high in the model atmosphere. In addition, the low atmosphere is where the radiation field transitions from being optically thick to optically thin. The chromosphere itself presents an additional challenge, since the radiation field is often decoupled from the thermal pool, particularly in some of the strongest, most energetically important transitions. There are a number of ways to model the energetics of the convection zone-to-corona system, ranging from approximate, parameterized descriptions of the thermodynamics (see e.g., \\citealt{Fan09,Hood09}), to highly realistic treatments of radiative transfer (see e.g., \\citealt{Martinez09a, Martinez09b}). Since our objective is to model the coupled system over large spatial scales, our goal is to find the most efficient treatment of the energetics possible that still provides a physically meaningful representation of the dynamic connection between the convection zone and corona. In order to describe the thermodynamics of the corona, a model should include the effects of electron thermal conduction along magnetic field lines and radiative cooling in the optically-thin limit. In addition, some physics-based or empirically-based source of coronal heating must be present if the model corona is to remain hot. In the convective interior well below the visible surface, radiative cooling can be treated in the diffusion limit. The trick is, how best to describe the effects of optically-thick radiative transfer in the region of the model atmosphere that lies between these two extremes. The most satisfying approach would be to couple the LTE transfer equation (or non-LTE population and transfer equations) to the MHD system to obtain cooling rates and intensities that could be compared directly to observations. Unfortunately, for large active region or global-scale problems, the computational expense of these techniques remains prohibitive. In \\citet{Abbett07} we tried the opposite approach --- ignore the transfer equation altogether, and develop an artificial, fully parameterized means of approximating surface cooling (in this case, we employed a modified form of Newton cooling). This worked relatively well, provided we carefully calibrated the adjustable parameters to match the average sub-surface stratification of previous, more realistic simulations of magneto-convection where the LTE transfer equation was solved in detail \\citep{Bercik02}. Of course, the principle drawback of this approach is that it is ultimately \\emph{ad hoc} and unphysical, and requires other, more realistic simulations as a basis for calibration in order to get meaningful results. We therefore have developed an approximation that is based on the macroscopic radiative transfer equation, and have incorporated this new treatment into our 3D MHD model, RADMHD. We describe this new method in Section 2. While it is important to treat the energetics of the system in a physically meaningful way, it is also important to remember that the utility of a given simulation ultimately depends on the statement of the problem. For an MHD simulation, this boils down to one's choice of initial states and boundary conditions. It is of great benefit, for example, to pose a simple, well-defined problem, and set up a numerical experiment that can shed light on what is believed to be the relevant physical processes in an otherwise complex system. For example, important progress has been made in understanding the physics of magnetic flux emergence by studying how idealized twisted flux ropes emerge through highly-stratified model atmospheres (see e.g., \\citealt{Cheung07,Fan04}). Yet the observed evolution of the photospheric magnetic field is often far more complex, particularly in and around CME and flare producing active regions. It is very difficult to set up a simple magnetic and energetic configuration that can initialize a simulation that will faithfully mimic the observed evolution of a real active region. It is desirable to do so, however, since we wish to quantitatively understand the physical mechanisms of energy storage and release, and the transport of magnetic energy and helicity between the convective interior and corona. To make progress, we could take a cue from meteorologists, and investigate a means to incorporate observational data directly into MHD models. This is not at all straightforward for solar models however, since data is obtained entirely through remote sensing, and not \\emph{in situ}. To address this challenge, we have developed a simple, rudimentary means of assimilating a time series of vector magnetograms into an MHD model of the photosphere-to-corona system. We briefly summarize this technique in Section 3, and apply it to the specific problem of finding a 3D magnetic field that is as force-free as possible given a single measurement of the vector magnetic field at the photosphere. ", "conclusions": "We have developed a rudimentary means of assimilating a time series of vector magnetograms into the interior volume of an MHD model in a manner that is stable, and does not over-specify the problem. We are currently using this assimilative technique to incorporate a timeseries of vector magnetograms into a 120 Mm$\\,^3$ RADMHD model atmosphere that contains a model photosphere, chromosphere, transition region and corona. The IVM data we are using is a four hour timeseries from NOAA AR 8210 --- a well-studied flare and CME-producing active region. The simulations are in their preliminary stages, and we hope to report on this work in the near future. In addition, we have presented a computationally efficient method of approximating optically-thick radiative cooling in our RADMHD quiet Sun models. The treatment improves upon the method of \\citet{Abbett07}, while still retaining the efficiency necessary to allow for large, active region-scale, convection zone-to-corona computational domains. Our simulations are progressing, and we are currently evaluating the efficacy and reliability of the new method. We are optimistic that each of these methods will improve the realism and utility of our current suite of numerical models." }, "1005/1005.1067_arXiv.txt": { "abstract": "Supermassive black holes are found at the centers of most galaxies and their inspiral is a natural outcome when galaxies merge. The inspiral of these systems is of utmost astrophysical importance as prodigious producers of gravitational waves and in their possible role in energetic electromagnetic events. We study such binary black hole coalescence under the influence of an external magnetic field produced by the expected circumbinary disk surrounding them. Solving the Einstein equations to describe the spacetime and using the force-free approach for the electromagnetic fields and the tenuous plasma, we present numerical evidence for possible jets driven by these systems. Extending the process described by Blandford and Znajek for a single spinning black hole, the picture that emerges suggests the electromagnetic field extracts energy from the orbiting black holes, which ultimately merge and settle into the standard Blandford-Znajek scenario. Emissions along dual and single jets would be expected that could be observable to large distances. ", "introduction": " ", "conclusions": "" }, "1005/1005.3548_arXiv.txt": { "abstract": "Massive metal-poor stars might form massive stellar black holes (BHs), with mass $25\\le{}m_{\\rm BH}/{\\rm M}_{\\odot{}}\\le{}80$, via direct collapse. We derive the number of massive BHs (N$_{\\rm BH}$) that are expected to form per galaxy through this mechanism. Such massive BHs might power most of the observed ultra-luminous X-ray sources (ULXs). We select a sample of 64 galaxies with X-ray coverage, measurements of the star formation rate (SFR) and of the metallicity. We find that N$_{\\rm BH}$ correlates with the number of observed ULXs per galaxy (N$_{\\rm ULX}$) in this sample. We discuss the dependence of our model on the SFR and on the metallicity. The SFR is found to be crucial, consistently with previous studies. The metallicity plays a role in our model, since a lower metallicity enhances the formation of massive BHs. Consistently with our model, the data indicate that there might be an anticorrelation between N$_{\\rm ULX}$, normalized to the SFR, and the metallicity. A larger and more homogeneous sample of metallicity measurements is required, in order to confirm our results. ", "introduction": "Ultra-luminous X-ray sources (ULXs, see Mushotzky 2004 for a review, and references therein) are defined as non-nuclear point-like sources with isotropic X-ray luminosity $L_{\\rm X}\\gtrsim{}10^{39}$ erg s$^{-1}$. The mechanism that powers the ULXs is still unknown, although various scenarios have been proposed. ULXs could be associated with high-mass X-ray binaries (HMXBs) powered by stellar-mass black holes (BHs) with anisotropic X-ray emission (e.g. King et al. 2001) or with super-Eddington accretion rate/luminosity (e.g. Begelman 2002; King \\&{} Pounds 2003; Socrates \\&{} Davis 2006; Poutanen et al. 2007) or with a combination of the two mechanisms (e.g. King 2008). ULXs could also be associated with HMXBs powered by intermediate-mass BHs (IMBHs), i.e. BHs with mass $100\\,{}{\\rm M}_\\odot{}\\le{}m_{\\rm BH}\\le{}10^5\\,{}{\\rm M}_\\odot{}$ (see van der Marel 2004 for a review). IMBHs with mass larger than $100\\,{}{\\rm M}_\\odot{}$ may be required to explain the brightest ULXs (i.e. the $\\lesssim{}4$ ULXs with $L_{\\rm X}\\gtrsim{}10^{41}$ erg s$^{-1}$), those ULXs showing quasi-periodic oscillations (M82 X-1, see Strohmayer \\&{} Mushotzky 2003, and NGC 5408 X-1, see Strohmayer et al. 2007) and some of those that are surrounded by isotropically ionized nebulae (Pakull \\&{} Mirioni 2002; Kaaret, Ward \\&{} Zezas 2004). However, IMBHs are not needed to explain the observational properties of most of the ULXs (e.g. Gon\\c{c}alves \\&{} Soria 2006; Stobbart, Roberts \\&{} Wilms 2006; Copperwheat et al. 2007; Roberts 2007; Zampieri \\&{} Roberts 2009). Thus, the objects classified as ULXs might actually be an inhomogeneous sample, including sources of different nature. Most of ULXs are located in galaxies with a high star formation rate (SFR, e.g. Irwin, Bregman \\&{} Athey 2004), although a small fraction (10--20 per cent, especially in the low-luminosity tail) might be associated with an old population (e.g. Colbert et al. 2004; Brassington, Read \\&{} Ponman 2005). The ULXs match the correlation between X-ray luminosity and SFR reported by various studies (Grimm, Gilfanov \\&{} Sunyaev 2003; Ranalli, Comastri \\&{} Setti 2003; Gilfanov, Grimm \\&{} Sunyaev 2004a,b,c; Kaaret \\&{} Alonso-Herrero 2008). Furthermore, the same studies indicate that the luminosity function of ULXs is the direct extension of that of HMXBs. Recent papers suggest a correlation between ULXs and low-metallicity environments, and propose that this may be connected with the influence of metallicity on the evolution of massive stars (Pakull \\&{} Mirioni 2002; Zampieri et al. 2004; Soria et al. 2005; Swartz, Soria \\&{} Tennant 2008). This scenario has been explored in detail by Mapelli, Colpi \\&{} Zampieri (2009, hereafter Paper~I) and by Zampieri \\&{} Roberts (2009), highlighting that a large fraction of ULXs may actually host massive ($\\sim{}30-80\\,{}{\\rm M}_{\\odot{}}$) stellar BHs formed in a low-metallicity environment. In fact, low-metallicity massive ($\\gtrsim{}40\\,{}{\\rm M}_\\odot{}$) stars lose only a small fraction of their mass due to stellar winds (Maeder 1992, hereafter M92; Heger \\&{} Woosley 2002, hereafter HW02; Heger et al. 2003, hereafter H03; Belczynski et al. 2010, hereafter B10) and can directly collapse (Fryer 1999; B10) into massive BHs ($25\\,{}{\\rm M}_\\odot{}\\le{}m_{\\rm BH}\\le{}80\\,{}{\\rm M}_\\odot{}$). Such massive BHs can power most of the known ULXs without requiring super-Eddington accretion or anisotropic emission. Furthermore, their formation mechanism can explain the correlation between ULXs and SFR, and the fact that ULXs are preferentially found in low-metallicity regions. In this Paper, we extend to a larger sample of galaxies the analysis reported in Paper~I, and we study the formation of massive BHs from the direct collapse of massive metal-poor stars and their possible connection with ULXs. In particular, we show that there is a correlation between the number of massive BHs formed in this scenario and the number of observed ULXs per galaxy. ", "conclusions": "Low-metallicity ($Z\\lesssim{}0.4\\,{}Z_\\odot{}$) massive ($\\gtrsim{}40\\,{}{\\rm M}_\\odot{}$) stars are expected to produce massive remnants ($25\\le{}m_{\\rm BH}/{\\rm M}_\\odot{}\\le{}80$, H03, B10) at the end of their evolution. Such massive BHs might power a large fraction of the observed ULXs in low-metallicity galaxies. In this Paper, we derived the number of massive BHs (N$_{\\rm BH}$) that are expected to form in a galaxy, via this mechanism, in the same time period that they could have a massive ($\\gtrsim{}15$ M$_\\odot{}$) donor companion. We find that N$_{\\rm BH}$ correlates well with the observed number of ULXs per galaxy (N$_{\\rm ULX}$). The slope of such correlation does not depend significantly either on the IMF or on the adopted stellar evolution model. The IMF and the stellar-evolution models affect only the normalization of N$_{\\rm BH}$ (the spread is generally less than a factor of 2). We stress that the stellar-evolution models adopted in this Paper neglect some important effects, such as the rotation and the possible influence of binary evolution. The final mass of the remnant is likely affected by the fact that the massive progenitor of the BH is fast rotating or that it resides in a binary, where additional mass loss is possible (see e.g. H03). Accounting for the probability that the progenitor of the massive BH is in a binary system likely introduces additional uncertainties. In addition, the model considered here does not include the possibility of pair instability supernovae (PISNs). PISNs are predicted to occur at very low metallicity, in the case of very massive stars ($\\ge{}140$~M$_\\odot${}), and lead to the complete disruption of the star (HW02; H03). Pulsational pair instability may occur also for smaller stellar masses ($100-140$~M$_\\odot{}$) and/or for larger metallicity, but it does not lead to a PISN, and a massive BH can form via direct collapse. Thus, PISNs probably do not play a role for stars with metallicity $Z\\gtrsim{}0.01\\,{}Z_\\odot$ and mass $<140$~M$_\\odot{}$ (H03). On the other hand, even assuming (as a strong upper limit) that all stars with mass $\\ge{}100\\,{}M_\\odot{}$ do not leave any remnant, due to a PISN, our estimates of N$_{\\rm BH}$ change by less than 10 per cent. The model described in this Paper is consistent with the observed correlation between the number of ULXs and the SFR, as well as with the fact that ULXs are found preferentially in low-metallicity environments. Furthermore, this model is a natural extension of the scenario described by Grimm et al. (2003). In fact, Grimm et al. (2003) find a correlation between the SFR and the number of X-ray sources (not necessarily ULXs) per galaxy, and they explain it with the correlation between the SFR and the number of HMXBs powered by stellar BHs. In this Paper we suggest that this correlation still holds for the ULXs, as the ULXs (or most of them) can be powered by massive BHs formed by the collapse of massive metal-poor stars. Our model predicts the existence of a dependence of N$_{\\rm ULX}$ on $Z$ (and the data suggest it, too), when the dominant effect due to the SFR is removed. Unfortunately, the statistical uncertainty of such dependence is still quite high, due to the dearth and to the inhomogeneity of the data. In particular, there is a large inhomogeneity in the metallicity measurements. Moreover, the metallicity needed in our model is that of the molecular clouds before the pollution from the first supernovae associated with the parent clusters, as very massive stars collapse into BHs before such supernovae. Thus, the metallicity measured today is likely higher than the value we should consider in our model. Only for some types of galaxies, such as the ring galaxies, where the star-formation history has a clear connection with the geometry of the system, is it possible to measure a pre-starburst value of $Z$, suitable for our purposes. A possible way to reduce this problem and to check our model is to take new measurements of the local metallicity in the neighbourhoods of the observed ULXs, or even in the nebula associated with the ULXs (Ripamonti et al., in preparation)." }, "1005/1005.4704_arXiv.txt": { "abstract": "We study the spectrum of high frequency radiation emerging from mildly dissipative photospheres of long-duration gamma-ray burst outflows. Building on the results of recent numerical investigations, we assume that electrons are heated impulsively to mildly relativistic energies by either shocks or magnetic dissipation at Thomson optical depths of several and subsequently cool by inverse Compton, scattering off the thermal photons of the photosphere. We show that even in the absence of magnetic field and non-thermal leptons, inverse Compton scattering produces power-law tails that extend from the peak of the thermal radiation, at several hundred keV, to several tens of MeV, and possibly up to GeV energies. The slope of the high-frequency power-law is predicted to vary substantially during a single burst, and the model can easily account for the diversity of high-frequency spectra observed by BATSE. Our model works in baryonic as well as in magnetically dominated outflows, as long as the magnetic field component is not overwhelmingly dominant. ", "introduction": "Understanding the origin of the prompt emission of Gamma-Ray Bursts (GRBs) has been hampered by two major challenges that have proven formidable obstacles despite more than four decades of studies. Radiating photons from an ultra-relativistic jet requires first a mechanism to convert at least some of the bulk kinetic energy into thermal motion of electrons, second a mechanism to radiate the electrons' energy into photons. These two processes are known as the problem of the dissipation mechanism and of the radiation process. The standard model that has been developed over the years assumes that internal shocks are responsible for the dissipation (Rees \\& M{\\'e}sz{\\'a}ros 1994), while electrons gyrating in a shock-generated magnetic field produce the radiation via the synchrotron process (M{\\'e}sz{\\'a}ros et al. 1994; Piran 1999; Lloyd \\& Petrosian 2000). Both mechanisms are fraught with numerous problems. Internal shocks assume that the outflow is released by the central engine with fluctuations in the Lorentz factor, so that different parts of the flow collide with each other, producing shocks that dissipate energy. Unfortunately, the internal shock mechanism has very little predictive power, since any behavior of the light curve can be explained by a suitable choice of the ejection history of the central engine. Since the physics of the emission from the engine is largely unconstrained, it is very hard to disprove internal shocks based on any observation. The only robust prediction of internal shocks as a dissipation mechanism is the low efficiency of the process (Kobayashi et al. 1997; Lazzati et al. 1999; Spada et al. 2000), due to the fact that internal shocks can dissipate only the energy associated with differential motions and not the energy associated to the bulk motion, which is much larger. Observations, instead, show that at least for a fraction of bursts, the efficiency of the prompt phase is 50 per cent, if not higher (Zhang et al. 2007). An alternative model that is becoming increasingly popular is magnetic dissipation in a Poynting dominated outflow (e.g., Thompson 1994; Spruit et al. 2001; Giannios \\& Spruit 2005). Even though magnetic dissipation could in principle provide very high efficiencies, it is, again, a very uncertain process and as such provides very little predictive power to allow for meaningful comparisons with observations. Synchrotron radiation, with possible self-Compton components, has long been considered the best candidate to explain the prompt emission of GRBs. While synchrotron radiation can easily explain the non-thermal nature of the observed spectrum, more thorough scrutiny reveals several important problems. First, optically thin synchrotron emission has a very well defined limiting slope $\\alpha\\ge-1/3$ (where $F(\\nu)\\propto\\nu^{-\\alpha}$) due to the synchrotron radiation spectrum of a single electron (the so-called line of death of synchrotron emission). However, at least several cases of GRBs with $\\alpha<-1/3$ have been detected (Crider et al. 1997; Preece et al. 1998; Ghirlanda et al. 2002, 2003). Second, the typical slope of the low-frequency part of the GRB spectrum is $\\alpha=0$ (Kaneko et al. 2006), which is not a natural slope of synchrotron radiation from shock-accelerated electrons. Third, the high frequency spectrum has a highly variable slope $\\beta$, not easily explained by synchrotron radiation from shock accelerated electrons that should produce a fairly standard spectrum with $\\alpha\\sim1$ (as in the X-ray afterglows). Finally, for the prompt emission to be efficient, electrons are expected to cool fast (Ghisellini et al. 2000), but the typical slope of cooling electrons ($\\alpha=1/2$) is not observed in GRB spectra (Kaneko et al. 2006). These well known difficulties of synchrotron emission have favored the study of various alternatives, including quasi-thermal Comptonization (Ghisellini \\& Celotti 1999), jitter radiation (Medvedev \\& Loeb 1999; Medvedev 2000; Workman et al. 2007; Morsony et al. 2008), bulk Compton (Lazzati et al. 2000); and photospheric emission (M{\\'e}sz{\\'a}ros \\& Rees 2000; M{\\'e}sz{\\'a}ros et al. 2002; Rees \\& M{\\'e}sz{\\'a}ros 2005; Pe'er et al. 2005, 2006; Giannios 2006; Giannios \\& Spruit 2007; Pe'er et al. 2007; Thompson et al. 2007; Ryde \\& Pe'er 2009, Lazzati et al. 2009). In this paper we expand the analysis of photospheric emission as a GRB prompt radiation mechanism, studying the conditions under which the photosphere can produce a high-frequency component with a non-thermal power-law shape up to tens of MeV. Photospheric radiation has two great advantages with respect to synchrotron emission: it does not require a dissipation mechanism and it naturally provides high efficiency. Photospheric radiation does not require a dissipation mechanism if the radiation is released before the full acceleration of the fireball and therefore at a stage when the fireball still contains a large fraction of internal energy. Numerical simulations have shown (Lazzati et al. 2009, hereafter LMB09) that this is indeed the case for long duration GRBs with a massive star progenitor. The photospheric efficiency of a typical GRB was found to be in good agreement with the observations (LMB09; Zhang et al. 2007). The obvious weakness of photospheric radiation is that it is customarily assumed to be thermal, therefore lacking the prominent non-thermal tails observed in GRBs. The possibility of adding non-thermal tails to the photospheric spectrum has been investigated before. Pe'er et al. (2006) showed that continuous dissipation and/or internal shocks at moderate optical depths give rise to a Comptonized spectrum with a flat energy spectrum $\\nu F(\\nu)\\propto\\nu^0$. Giannios (2006, see also Giannios \\& Spruit 2007) performed a similar study and reached analogous conclusions, focusing on magnetic reconnection as the dissipation mechanism. Here we argue that a dissipation mechanism that manages to be so continuous as to keep the electron temperature at a constant equilibrium value under the intense cooling of IC scattering is very unlikely. A more realistic assumption is that sub-photospheric electron heating takes place as a series of one or more episodes of impulsive acceleration and subsequent cooling. We show that in this case a non-thermal high-frequency tail is produced, characterized by a slope that is sensitive to a combination of various parameters and which is therefore expected to be highly variable during the prompt emission of GRBs. This paper is organized as follows: in \\S~2 we discuss the origin of the non-thermal tails, in \\S~3 we discuss the propagation of the non-thermal spectrum in the optically thick sub-photospheric plasma, and in \\S~4 we present Monte Carlo calculations of the model. We summarize and discuss our results in \\S~5. ", "conclusions": "We have presented a method by which a primary blackbody spectrum of a GRB outflow photosphere can be Comptonized into a non-thermal high-frequency power-law tail. In this scenario, the bulk of the GRB prompt emission is thermal photospheric and the non-thermal tail contains a relatively small fraction of the radiation energy. The majority of the energy of the radiation is therefore released before it is converted into bulk kinetic energy, solving the problem of identifying a dissipation mechanism with high efficiency (LMB09). The non-thermal tail is produced by thermal electrons accelerated to mildly relativistic energies by an intermittent dissipation mechanism. The electrons are allowed to cool through IC interactions with the photon field between acceleration events. This dissipation mechanism is not required to have an efficiency larger than a few per cent, since the power-law tail only contains a fraction of the total energy of the radiation.The spectral slope of the high energy radiation depends on the typical Lorentz factor of the accelerated electrons, on the peak frequency of the thermal photon spectrum, and on the ratio of the photon to lepton density. Differently from previous work (Pe'er et al. 2005, 2006; Giannios 2006) we assume that the electrons are accelerated intermittently and not continuously and that the electrons are allowed to cool among acceleration events. In addition, we assume that the energy density in radiation is larger than that in the relativistic electrons, a situation that is easily realized in jets born inside massive progenitor stars (LMB09). Compared to the standard optically thin synchrotron, this mechanism can explain the steep low-frequency slopes observed in the early phases of some bright bursts (Crider et al. 1997; Preece et al. 1998; Ghirlanda et al. 2002, 2003) and the transient thermal bumps detected is several events (Ryde 2005; Ryde \\& Pe'er 2009; Ryde et al. 2010). Bursts for which no dissipation takes place should not display any non-thermal features. A complete discussion of the low-frequency non-thermal tail is however beyond the scope of this paper and will be presented elsewhere. Explaining a typical low-frequency tail $F(\\nu)\\propto\\nu^0$ is indeed a challenge for any model and not only for the photospheric scenarios presented here and elsewhere (M{\\'e}sz{\\'a}ros \\& Rees 2000; M{\\'e}sz{\\'a}ros et al. 2002; Rees\\& M{\\'e}sz{\\'a}ros 2005; Pe'er et al. 2005, 2006; Giannios \\& Spruit 2007; Pe'er et al. 2007; Thompson et al. 2007; Ryde \\& Pe'er 2009, LMB09). Radiation from dissipative photospheres has been investigated before (Pe'er et al. 2005, 2006; Giannios 2006; Giannios \\& Spruit 2007; Beloborodov 2010). Giannios (2006) and Giannios \\& Spriut (2007) concentrate on the case of a continuous energy injection or ``slow heating'', different from our impulsive acceleration assumption. Pe'er et al. (2005, 2006) consider instead two possible scenarios: ``slow heating'' and impulsive heating by internal shocks. They find that the bulk of the photons are shifted in energy once a steady state electron population is attained. They calculate that, for a fairly wide parameter range, the comoving peak frequency is of few tens keV. Attaining such a steady state population requires however a large optical depth and as a consequence we do not find the same results. To understand the reason for this, consider an injection episode at an optical depth of a few. Due to IC scattering, the hot electrons are efficiently cooled and their energy given to a small fraction of photons that are shifted to high frequencies. For small optical depth, the energized photons reach the photosphere with only negligible losses and decouple from the electron population. The energy given to the electrons by the dissipation event is therefore given to a small number of photons that produce a power-law tail as shown in the figures. In order to obtain a steady state electron population able to exchange energy effectively with the radiation field, a large number of scattering per photon is required, so that the high-frequency photons can return their energy to the electron population. Large optical depths are therefore required for electron and photon populations to settle in a steady state configuration dictated by the balance of heating and cooling. Such optical depths are much larger than the ones envisioned in this work. In any case, a fraction of the dissipated energy is left in the electron population that thermalizes at a temperature higher than the one before the dissipation. For the parameters adopted in this paper the effect of the increased electron temperature is negligible because the increase of the electrons temperature is very small for the optical depths considered. Finally, Beloborodov (2010) investigated the radiation produced by a well-defined dissipation mechanism: nuclear and Coulomb collisions within a baryonic outflow with a substantial population of neutrons (see also Beloborodov 2003; Rossi et al. 2006). He finds that a typical spectrum with a high-frequency slope $F(\\nu)\\propto\\nu^{-1.4}$ is obtained. Differently to our result, the non-thermal spectrum is due to the presence of non-thermal leptons. A characteristic of any photospheric model is that the radiation is released at a small radius $r_{\\rm{ph}}\\sim10^{10-13}$~cm (e.g. Rees \\& M{\\'e}sz{\\'a}ros 2005). The main consequence of such a small radius is that the compactness of the region is large (Pe'er \\& Waxman 2004) and therefore photon-photon interactions result in the production of electron-positron pairs. This can have two consequences. If the newly created pairs do not outnumber the original electrons, the only consequence is the presence of a sharp cutoff in the spectrum at $h\\nu=511$~keV in the comoving frame ($\\sim100$~MeV observed). If, on the other hand, the pairs outnumber the original electrons, a new photosphere is created, typically at a radius 3 to 10 times larger than the original photosphere (Rees \\& M{\\'e}sz{\\'a}ros 2005). The scenario we developed in this paper is not dependent on whether the photosphere is due to the original electrons or to pairs, the only difference being that a pair photosphere has a lower photon to lepton ratio and produces therefore a steeper high-frequency spectrum. In addition, since the number of pairs and the number of photons are not independent, we showed that a pair photosphere with at least several energization events would produce spectra with a typical slope $1<\\beta_{\\rm{sat}}<1.5$, in good agreement with observations (e.g., Kaneko et al. 2006). The Fermi satellite has recently shown that at least some bright GRBs have spectra extending well into the GeV regime (Abdo et al. 2009abc; 2010; Ghisellini \\& Ghirlanda 2010; Granot et al. 2010). For any reasonable combination of parameters, GeV radiation cannot be produced at the photosphere, since it would be immediately absorbed by photon-photon collisions to produce pairs. In this scenario, the GeV emission can be either internal, i.e. produced by dissipation above the photosphere (see \\S~2.2; Toma et al. 2010) or external, due to the interaction of the outflow with the interstellar medium (Kumar \\& Barniol Duran 2010; Ghirlanda et al. 2010; Ghisellini et al. 2010). In terms of predictive power, the most notable prediction of this model is the presence of substantial variability of the high-frequency spectrum due to the dependence of the spectrum on parameters that are expected to be highly variable such as, e.g., the peak frequency of the thermal spectrum in the comoving frame (LMB09). Such variability might be however hard to detect since it is expected to have short time-scales and would therefore be averaged out in a time-integrated spectrum. The dependence of the spectral slope on the peak frequency of the spectrum opens also the possibility of correlations between the peak frequency and the spectral slope at high frequencies. Since, however, the slope depends on two more parameters (the ratio of photons to electrons and the typical Lorentz factors of the electrons) it is not obvious that such correlations should be visible in experimental data." }, "1005/1005.3915_arXiv.txt": { "abstract": "{We obtained \\emph{Herschel} PACS and SPIRE images of the thermal emission of the debris disk around the A5V star $\\beta$\\,Pic. The disk is well resolved in the PACS filters at 70, 100, and 160\\,$\\mu$m. The surface brightness profiles between 70 and 160\\,$\\mu$m show no significant asymmetries along the disk, and are compatible with 90\\% of the emission between 70 and 160\\,$\\mu$m originating in a region closer than 200\\,AU to the star. Although only marginally resolving the debris disk, the maps obtained in the SPIRE 250 -- 500\\,$\\mu$m filters provide full-disk photometry, completing the SED over a few octaves in wavelength that had been previously inaccessible. The small far-infrared spectral index ($\\beta = 0.34$) indicates that the grain size distribution in the inner disk ($<$200\\,AU) is inconsistent with a local collisional equilibrium. The size distribution is either modified by non-equilibrium effects, or exhibits a wavy pattern, caused by an under-abundance of impactors which have been removed by radiation pressure. } ", "introduction": "The $\\beta$\\,Pic disk, discovered by IRAS \\citep{aumann:1984}, was the first debris disk to be directly imaged in scattered light \\citep{smith:1984}. It is seen close to edge-on and extends in the optical out to 95\\arcsec, corresponding to 1800\\,AU \\citep{larwood:2001}. $\\beta$\\,Pic (A5V) is one of the closest \\citep[19.44\\,$\\pm$\\,0.05\\,pc,][]{vanleeuwen:2007} and youngest debris disks. The estimated age \\citep[12 Myr,][]{zuckerman:2001} significantly exceeds typical timescales for the survival of pristine circumstellar dust grains \\citep[e.g.,][]{fedele:2010}, hence continuous replenishment of the dust, presumably through collisions of planetesimals, is needed. The closeness of the object ensures that it can also be spatially resolved at long wavelengths: \\citet{holland:1998} resolved the disk at 850\\,$\\mu$m and \\citet{liseau:2003} at 1200\\,$\\mu$m. Optical and near-infrared observations of the inner part ($<$100AU) of the disk yield evidence of asymmetries such as warps and density contrasts, which may relate to the presence of planetesimals \\citep{kalas:1995,pantin:1997,mouillet:1997,heap:2000, telesco:2005}. \\citet{lagrange:2009B} imaged a possible companion at a projected distance of 8\\,AU from the star. Images of $\\beta$\\,Pic in scattered stellar light directly detect small grains and indirectly larger grains that produce the smaller ones through collisions. The grain-size distribution can be quantitatively constrained from the spectral energy distribution (SED) of the disk, in the infrared and (sub)mm domains. The spectral index of the SED at the longest wavelengths \\citep{liseau:2003,nilsson:2009} is inferred to be fairly low, which according to modeling \\citep{draine:2006,natta:2007,ricci:2010} can be interpreted as a deficit of small grains. In this paper, we present far-infrared imaging of the $\\beta$\\,Pic debris disk in six \\emph{Herschel} photometric bands between 70 and 500\\,$\\mu$m. These bands cover for the first time the long-wavelength side of the peak in the thermal emission of the disk, and the large aperture of the telescope enables us to resolve the disk at far-IR wavelengths for the first time. With these data, we measure the surface brightness profiles of the disk and readdress the issue of the grain-size distribution in the inner 200\\,AU. ", "conclusions": "We have presented images of the $\\beta$\\,Pic debris disk in six photometric bands between 70 and 500\\,$\\mu$m using the PACS and SPIRE instruments. We resolve the disk at 70, 100, 160, and 250\\,$\\mu$m. The images at 70--160\\,$\\mu$m show no evidence of asymmetries in the far-infrared surface brightness along the disk of $\\beta$\\,Pic. The observed profiles are compatible with 90\\% of the emission originating in a region within a radius of 200\\,AU from the star. The disk-integrated photometry in the six \\emph{Herschel} filters provides a far infrared SED with small spectral index $\\beta\\,\\approx\\,0.34$, which is indicative of a grain size distribution that is inconsistent with a local collisional equilibrium. The size distribution is modified by either non-equilibrium effects, or exhibits a wavy pattern, caused by the under-abundance of impactors that are small enough to be removed by radiation pressure." }, "1005/1005.2408_arXiv.txt": { "abstract": "We propose a novel method to estimate \\mm, the ratio of stellar mass (\\mstar) to black hole mass (\\mbh) at various redshifts using two recent observational results: the correlation between the bolometric luminosity of active galactic nuclei (AGN) and the star formation rate (SFR) in their host galaxies, and the correlation between SFR and \\mstar\\ in star-forming (SF) galaxies. Our analysis is based on \\mbh\\ and \\lbol\\ measurements in two large samples of type-I AGN at z$\\simeq$1 and z$\\simeq$2, and the measurements of \\mm\\ in 0.05$<$z$<$0.2 red galaxies. We find that \\mm\\ depends on \\mbh\\ at all redshifts. At z$\\simeq$2, \\mm$\\sim$280 and $\\sim$40 for \\mbh=$10^8$ and \\mbh=$10^9$\\msun, respectively. \\mm\\ grows by a factor of $\\sim$4-8 from z$\\simeq$2 to z=0 with extreme cases that are as large as 10--20. The evolution is steeper than reported in other studies, probably because we treat only AGN in SF hosts. We caution that estimates of \\mm\\ evolution which ignore the dependence of this ratio on \\mbh\\ can lead to erroneous conclusions. ", "introduction": "\\label{sec:intro} Study of the co-evolution of Active Galactic Nuclei (AGN) and their host galaxies provides important clues about the growth of super massive black holes (SMBHs) and the star formation (SF) history of the Universe. In the local universe one finds a tight correlation between the SMBH mass (\\mbh) and the mass of the bulge of its host, \\mbul\\ (Marconi \\& Hunt, 2003; H\\\"{a}ring \\& Rix 2004, hereafter HR04), or alternatively with the stellar velocity dispersion \\sigs\\ (Ferrarese \\& Merrit 2000; Gebhardt \\et 2000). Typically \\mbul/\\mbh$\\simeq$500--1000. $\\mbul/\\mbh$ must have been smaller at high redshift. For example, $\\mbh\\sim10^{9-10}\\,\\Msun$ are often observed at z=3--6 (e.g. Netzer 2003; Shemmer \\et 2004; Fan \\et 2006), yet galaxies that are 500-1000 times more massive are never observed at z$>$0.5 and are very rare even at z$\\simeq$0. The evolution of the $\\mbh-\\sigs$ relationship has been studied in numerous papers. Examples are Shields \\et (2003) who find no evolution up to z$\\simeq$2, and Woo \\et (2008) who suggest that $\\mbul/\\mbh$ has increased by a factor of $\\sim$3 since z=0.6. The uncertainties in all such measurements are very large due to the difficulties in measuring \\sigs\\ or \\mbul\\ in high redshift AGN hosts. While measuring \\mbul\\ in high-redshift galaxies is difficult, the total stellar mass, \\mstar\\ is easier to obtain. This is achieved by multi-band spectral energy distribution (SED) fitting, which is used to constrain \\ml. However, obtaining \\mm\\ for AGN hosts is severely limited by the problematic subtraction of the bright point-like continuum in type-I AGN and the estimation of \\mbh\\ in type-II AGN. Several recent studies used deep imaging to estimate \\mm\\ (or $L_{\\rm Host}/M_{\\rm BH}$) in z$\\sim$0.5--3 AGN by resolving the host galaxy emission and by careful PSF modelling (e.g. Kukula \\et 2001; Peng \\et 2006; Kotilainen \\et 2007; Bennert \\et 2010). The measured host luminosity was translated to \\mstar\\ by \\textit{assuming} a certain \\ml\\ ratio. A common assumption is that AGN hosts are ``red and dead'', with a stellar population which evolves passively from a high formation redshift, e.g. $z_{\\rm form}\\simeq$5. A detailed study of this type, including a summary of many earlier findings, is given in Decarli et al. (2010; hereafter D10) who find that \\mm\\ evolves following $\\mmwp\\propto\\,z^{-0.28}$. The assumption of non-SF AGN hosts may describe some objects, but many studies find that hosts of luminous AGN often contain much younger stellar populations (e.g., Kauffmann \\et 2003; Jahnke \\et 2004; Silverman \\et 2009; Merloni \\et 2010, and references therein). In this \\textit{Letter} we suggest a novel method to estimate the evolution of \\mm. Our approach makes use of the known relationships between star formation rate (SFR) and \\mstar\\ in SF galaxies (SFGs), and between the bolometric luminosity of AGN (\\lbol) and the SFR in their hosts. We describe these relationships in \\S\\ref{sec:sfr_corr} and use them to estimate \\mm\\ in SF galaxies at z$\\simeq$0.15, \\zandz\\ in \\S\\ref{sec:results}. In \\S\\ref{sec:discussion} we discuss the implications to the co-evolution of SMBHs and their hosts. Throughout this work we assume a standard $\\Lambda${\\sc CDM} cosmology with $\\Omega_{\\Lambda}$=0.7, $\\Omega_{M}$=0.3 and $H_{0}$=70\\,\\kms\\,Mpc$^{-1}$. ", "conclusions": "\\label{sec:discussion} The analysis presented here suggests strong evolution in \\mm\\ up to z=2, much steeper than what is suggested in other studies. In particular, our mean values of \\mm\\ at \\zandz\\ are smaller than the ones reported in D10, which is the most up-to-date compilation of such works. There are three main reasons for these differences: \\begin{figure} \\centering \\includegraphics[width=6.2cm]{Fig3_TN10.eps} \\caption{The distributions of \\mm\\ at different redshift, for the three \\mbh\\ sub-groups discussed in the text.} \\label{fig:MM_hist_3z_3m} \\end{figure} \\begin{enumerate} \\item As already mentioned, most earlier works estimate \\mstar\\ by assuming a passive evolution from a high formation redshift. This is obviously an over-simplification of the evolution of most galaxies and significant epochs of SF at z$<$4 (e.g. D07; E07; Drory \\& Alvarez 2008; van Dokkum \\et 2010, and references therein). Accounting for younger stellar populations would result in lower \\mstar\\ and thus lower \\mm\\ (see D10 and Peng \\et 2006). \\item The D10 sample represents the \\textit{minority} of the AGN population, as hinted by two biases. First, the majority of the D10 sources lie in the top 15\\% of the \\mbh\\ distributions corresponding to their redshift. Second, the selection criteria for the D10 HST observations could have been biased towards large, resolved galaxies with large \\mstar. \\item The mean \\mbh\\ in high redshift AGN samples is systematically larger than the corresponding low redshift \\mbh. For example, in our large sample of red galaxies, 99\\% of the sources show \\mbh$<10^{8.8}$\\msun\\ while 20\\% of the z$\\simeq$1 and 57\\% of the z$\\simeq$2 AGN have larger \\mbh. All the z$\\simeq$2 sources in D10 have \\mbh$\\gtsim10^{8.7}$\\msun. Such objects should only be compared with local galaxies which host BHs that are at least as massive. As Fig.~\\ref{fig:MM_vs_Mbh_all} shows, this corresponds to \\mm=100--200, instead of the commonly used $\\sim$700. \\end{enumerate} In conclusion, while our work applies to most AGN, the D10 sample probably represents the remaining sources. The results presented here point to a scenario where many galaxies have to increase their mass by factors of 4-8 (2-4) since z$\\simeq$2 (z$\\simeq$1). The growth factors for the most massive BHs are not well determined since the number of very massive galaxies in the local universe is not large enough to reliably extend the results of Fig.~\\ref{fig:MM_vs_Mbh_all} beyond $\\mbh\\simeq10^9\\msun$. These numbers represent the requirement for the galaxies to \\textit{over}-grow their SMBHs by the above factors. This seems to be consistent with models which suggest that the high mass SMBHs observed at z$\\simeq2$ could have accumulated most of their mass by that redshift (Marconi \\et 2004). On the other hand, it may be in contradiction with at least some scenarios linking AGN activity to the shut-down of SF in their host galaxies (see Somerville \\et 2008; Cattaneo \\et 2009 and references therein). While a full discussion of the various growth scenarios of \\mstar\\ is beyond the scope of this \\textit{Letter}, we comment briefly on some of these ideas. Major galaxy mergers would increase both \\mstar\\ and \\mbh, either through starbursts and gas accretion in ``wet mergers'' or the possible coalescence of the two SMBHs involved in ``dry mergers''. However, theoretical and observational arguments (e.g. Lotz \\et 2008; Genel \\et 2009) suggest a low rate of such events for z$<2$ galaxies. Thus, major mergers cannot change \\mm\\ by more than a factor of $\\sim2-3$ between z=2 and z=0. Small ``dry mergers'' may help. For example, Naab, Johansson \\& Ostriker (2009) show that present-day massive red ellipticals gain $\\sim40\\%$ of their mass through accretion of smaller companions since z$\\sim2$. Intense SF in outer parts of galaxies due to external source of cold gas which does not find its way to the centre (e.g. van Dokkum \\et 2010), is another possibility. More possibilities and more references are discussed in Benson \\& Devereux (2009). The \\mm\\ distributions presented here suggest an increase in \\mm\\ by factors beyond what is suggested in many theoretical studies. Finally, we comment on the possibility that the suggested evolution of \\mm\\ could be due to two wrong assumptions. First, many more AGN hosts may not lie on the SF sequence or may not obey the \\lbol-SFR correlation used here. This is unlikely to be the case at low redshift, where SDSS type-II AGNs are used. However, the selection of at least some of the most luminous, high redshift high-\\lbol\\ sources in Fig.~\\ref{fig:Lbol_Lsf} may be biased towards high FIR luminosity, high SFR AGN hosts (e.g. Zheng \\et 2009) in particular if these are found in mergers that are not part of the SF sequence. \\textit{Herschel} observations of well-defined AGN samples are likely to resolve this issue. Second, the Drory \\& Alvarez (2008) work shows a decline in SFR at the high-\\mstar\\ end for z$<$2 galaxies. Thus, some of our \\lbol-based SFR estimates might be associated with considerably larger values of \\mstar. The E07 and D07 SF sequences do not show such a decline. We conclude that there is a steep evolution in \\mm\\ from z$\\simeq2$ to z=0 for SF AGN hosts. This trend is barely consistent with some, but not all galaxy and BH evolution models. We have also demonstrated the crucial importance of considering different \\mbh\\ groups separately when evaluating the \\mm\\ evolution. \\vspace*{0.4cm} \\noindent We thank an anonymous referee for very useful comments on the manuscript. We thank Ido Finkelmann for fruitful discussions; Vincenzo Mainieri for allowing us to use zCOSMOS results ahead of publication; Roberto Decarli and Emanuele Daddi for sending us their data; Niv Drory and Samir Salim for their thoughtful comments; and the MPA/JHU team for making their SDSS catalogues available to the public. This study makes use of data from the SDSS (http://www.sdss.org/collaboration/credits.html). Funding for this work has been provided by the Israel Science Foundation grant 364/07. \\vspace*{ -0.6cm}" }, "1005/1005.0916_arXiv.txt": { "abstract": "SDSS\\,J094857.3+002225 is a very radio-loud narrow-line Seyfert 1 (NLS1) galaxy. Here, we report our discovery of the intranight optical variability (INOV) of this galaxy through the optical monitoring in the \\emph{B} and \\emph{R} bands that covered seven nights in 2009. Violent rapid variability in the optical bands was identified in this RL-NLS1 for the first time, and the amplitudes of the INOV reaches 0.5 mag in both the \\emph{B} and \\emph{R} bands on the timescale of several hours. The detection of the INOV provides a piece of strong evidence supporting the fact that the object carries a relativistic jet with a small viewing angle, which confirms the conclusion drawn from the previous multi-wavelength studies. ", "introduction": "Narrow-line Seyfert 1 galaxies (NLS1s) are a special and interesting group of active galactic nuclei (AGNs). They show narrow optical Balmer emission lines [FWHM(H$\\beta$) $<$ 2000 km s$^{-1}$], weak [\\ion{O}{3}]$\\lambda$5007 emission ([\\ion{O}{3}]/H$\\beta<3$), strong \\ion{Fe}{2} emission, and soft X-ray excess \\citep{pog00,ost85,sul00,bol96,bol97}. NLS1s show remarkable radio-loud/radio-quiet bimodality \\citep{lao00}. Only $7\\%$ of NLS1s are radio-loud objects \\citep{zho03,kom06}. The fraction is much smaller than that found in QSOs. Very radio-loud NLS1s (RL-NLS1s, $R>100$) are even much fewer ($\\sim2.5\\%$) \\citep{kom08}, where the radio loudness $R$ is commonly defined as the flux ratio of radio to optical at $\\lambda4400$ \\citep{kel89}. So far, it is still a puzzle why RL-NLS1s are so scarce. At present, the origin of RL-NLS1s is also still poorly understood. A few efforts have been made in the past few years to understand the nature of RL-NLS1s. Yuan et al. (2008) found that the broadband spectra of some RL-NLS1s are similar to those of high-energy-peaked BL Lac objects, and suggested that some of them may be BL Lac objects actually. Basing upon the recent observation taken by \\emph{Fermi} satellite, some RL-NLS1s display a hard X-ray component suggesting the presence of relativistic jets on the line of sight \\citep{fos09,abd09a}. The presence of the relativistic jets motivates us to search for intranight optical variability in some RL-NLS1s, because of the well-known beaming effect (e.g., Wagner \\& Witzel 1995). \\citet{kom06} argued that SDSS\\,J094857.3+002225 is a right candidate for searching for RL-NLS1s with beaming effect. The object is a very radio-loud NLS1 at $z=0.585\\pm0.001$. The reported radio loudness derived from the radio flux at 5 GHz ranges from 194 to 1982 \\citep{zho03,kom06,abd09a}. It is in the CRATES catalog as a flat-spectral radio source \\citep{hea07}. The simultaneous observations taken by both \\emph{Swift} and \\emph{Fermi} also suggest that the broadband spectral energy distribution is similar to those of flat-spectral radio quasars \\citep{abd09a,zho03}. Recent photometry from the Guide Star Catalogs 2.21 is $B_{\\mathrm{J}}(\\mathrm{GSC2.21})=18.83$ mag \\citep{zho03}. Previous studies revealed multi-wavelength variabilities in the object at timescales from day to year. Previous radio observations indicate its fluctuation in the radio band on the timescale from weeks to years \\citep{abd09a}. \\citet{zho03} also said that the object shows long-term variability in both the radio and optical bands. The amplitude of the variation in the radio can be $60\\%$ within a year. The long-term variability amplitude may be about 1 mag in the optical band. The latest multi-wavelength campaign carried out by \\citet{abd09b} discovered an optical variability on day timescales. Dramatic flux variabilities in both X-rays and radio 37 GHz were also found in the study. In this Letter, we report an optical monitor for the RL-NLS1 SDSS\\,J094857.3+002225. The monitor was designed to search for intranight optical variability (INOV) in the object. The INOV should be detected if the object indeed hosts a relativistic jet beaming toward the observers. ", "conclusions": "The particular RL-NLS1 galaxy SDSS\\,J094857.3+002225 was monitored in optical bands by NAOC 80 cm TNT telescope to search for its INOV phenomenon. Our optical monitoring indeed provides clear evidence for the presence of INOV in both the \\emph{B} and \\emph{R} bands in the object. The object exhibits optical variability not only on the timescale of a week, but also on several hours. The detection of the INOV indicates that the object contains a relativistic jet on the line of sight of an observer, which confirms the conclusion drawn from the high-energy observations (e.g., Abdo et al. 2009a, 2009b) and from the inverted radio spectrum and high brightness temperature (Zhou et al. 2003). SDSS\\,J094857.3+002225 is particular for its observational properties. On the one hand, its optical spectrum with strong \\ion{Fe}{2} emission is typical of NLS1s. The narrow H$\\beta$ emission yields a relatively low black hole (BH) mass $\\sim 4.0\\times10^7\\ M_\\odot$ and a high Eddington ratio (Zhou et al. 2003). On the other hand, some observational behaviors are characteristic of blazars with the relativistic jets close to the line of sight, such as the INOV detected here, flat radio spectrum, high brightness temperature, and variable $\\gamma$-ray emission (see citations in Section 1). So far, outstanding RL-NLS1s with blazer-like radio emission have been revealed by multi-wavelength observations in several cases including the object SDSS\\,J094857.3+002225. We refer the reads to Yuan et al. (2008) for a brief summarization. With the successful launch of \\emph{Fermi} satellite, $\\gamma$-ray emission was detected in four RL-NLS1s, including the object studied here, which suggests the presence of fully developed jets in these objects \\citep{abd09c}. The authors argued that the four RL-NLS1s may form a new class of $\\gamma$-ray AGNs because of their small BH masses, large Eddington ratios, and possibly disk-like morphology of the host galaxies. The intrinsic mechanism of RL-NLS1s is an attractive field. There are two possible models for interpreting RL-NLS1s. The first one is the inclination model \\citep{ost85,kom06,wha06,kom08}. The model suspects that the (at least a fraction of) RL-NLS1s are preferentially viewed pole-on. The observed narrow width of the Balmer emission lines could be resulted from small inclination if the broad-line region (BLR) is constrained to a plane \\citep{wil86,bia04}. In fact, there is some evidence supporting a flat BLR in some RL-AGNs \\citep{jar06,sul03}. In this scenario, the BH mass is largely underestimated in these objects since the current available estimation of the BH mass of AGN from single-epoch spectroscopic observation comes from an assumption of an isotropic distribution of the broad-line clouds with random orbital inclinations \\citep{kas05,pet04,ben07}. Although the inclination model sounds reasonable because it is able to shift the location of RL-NLS1s on the $R$--$M_{\\mathrm{BH}}$ plane to the massive BH end \\citep{lao00,lac01}, the massive BHs are not supported by the lack of massive bulges in several cases in which the host galaxies can be resolved. The second is the accretion mode model \\citep{kom06,wha06,kom08}. The RL-NLS1s with small BH masses are accreting close to or even above the Eddington limit. Low-mass BHs may lead to narrow emission lines when Keplerian velocities are considered mostly \\citep{wha06}. The accretion is thought related to the radio emission. Different accretion modes may result in different phenomena and maybe can explain the differences in the radio loudness. Accretion processes are known related to the spin of the accreting BHs. So the rapid spin of BHs may also affect the radio loudness of NLS1s \\citep{kom06}. It is possible that accretion mode combining with BH spin can explain the nature of the RL-NLS1s. Although our monitor indicates that the extremely high radio emission in RL-NLS1 SDSS\\,J0948\\\\57.3+002225 is mainly contributed from the beamed non-thermal jet with a small viewing angle (can also be found in the aforementioned other studies), more information is needed in the future to investigate the origin of the relativistic jet." }, "1005/1005.0127_arXiv.txt": { "abstract": " ", "introduction": "As an alternative to inflation, the curvaton model \\cite{Lyth:2001nq} is a well-motivated proposal for explaining the observed scale-invariant primordial density perturbation in the framework of inflation (and some pioneering works on this mechanism have been considered in Refs. \\cite{Mollerach:1989hu, Linde:1996gt, Enqvist:2001zp}). This model is based on the inflationary scenario of multiple fields, in which two fields are required at least. In this model the universe is composed of radiation decayed from the inflaton and the curvaton field. During inflation, the curvaton is subdominant and provides entropy (isocurvature) perturbation, and afterwards, the entropy perturbation can be converted to curvature perturbation as long as the curvaton decays into radiation before primordial nucleosynthesis. After the curvaton decays, the universe enters the standard thermal history, and then the primordial curvature perturbation leads to the formation of the large scale structure of our universe\\cite{Lyth:2002my, Lyth:2003ip}. Recently, a curvaton scenario realized in the frame of stringy inflationary models was presented in Ref. \\cite{Li:2008fma}. The idea of brane inflation proposed in \\cite{Dvali:1998pa}, is recently successfully realized by virtue of warped compactifications\\cite{Kachru:2003si}. In the stringy landscape a number of probe branes are allowed to move in warped throats, and under the description of supergravity (SUGRA) their dynamics is described by a Dirac-Born-Infeld (DBI) action in the relativistic limit. When applied to cosmology, it was found that a relativistically moving D3-brane in a deformed AdS throat is able to drive inflation without slow-roll at early universe\\cite{Silverstein:2003hf, Chen:2004gc}. This idea was later extended into the scenario of multiple brane inflation appeared in Refs. \\cite{Cai:2008if, Cai:2009hw} (see \\cite{Cline:2005ty, Piao:2002vf} for earlier studies under slow-roll approximation). A distinguish feature of the model in this type is that a large positive non-local non-Gaussianity can be obtained due to an enhancement of a small sound speed for the perturbation\\cite{Alishahiha:2004eh, Chen:2006nt}, which is in contrast to the prediction of a canonical single-field inflation model\\cite{Maldacena:2002vr}. Established on the scenario of multiple brane inflation\\cite{Cai:2008if, Cai:2009hw} and motivated by the model of {\\it spinflation} \\cite{Easson:2007dh}, an explicit construction of the curvaton brane model was recently realized in Ref. \\cite{Zhang:2009gw} in which angular degrees of freedom had been introduced to perform a relativistic rotation in a warped throat. The investigation of non-Gaussianities is crucial in distinguishing the curvaton scenario from usual inflation models. A key prediction of a general curvaton model is that a sizable local non-Gaussianity can be obtained of which the value is mainly decided by the occupation of curvaton energy $\\Omega_{\\chi}$ after inflation characterized by a transfer efficiency parameter $$r=\\frac{3\\Omega_{\\chi}}{4-\\Omega_{\\chi}}$$ which is usually calculated on the hypersurface of the curvaton decay. As a consequence, the decay mechanism of a curvaton field plays an important role in determining the detailed information of primordial perturbation seeded by the curvaton field \\cite{Sasaki:2006kq}. It was pointed out in Ref. \\cite{Sasaki:2006kq} that at third order, the primordial curvature perturbation under the assumption of sudden curvaton decay is nicely consistent with that obtained in the case of a non-instantaneous decay in the limits of small $r$ and $r\\rightarrow1$; however, their bispectra and trispectra deviate from the others in the middle band of the allowed value regime of $r$. The implications of the non-Gaussian features of curvature perturbation on the curvaton scenario was studied in \\cite{Huang:2008ze, Huang:2008bg, Gong:2009dh}. In the present paper we extend the investigation of Ref. \\cite{Sasaki:2006kq} relating non-Gaussian features of primordial curvature perturbation to the curvaton decay mechanism into a more general curvaton scenario. Specifically, we consider a model of curvaton brane. We study the generation of curvature perturbation and non-Gaussianity originating from the vacuum fluctuations on the curvaton brane in cases of a sudden decay on the slice of uniform curvaton density, and then compare their observational signatures with the standard curvaton model to illustrate the significance of the decay process of the curvaton field. Moreover, since the model of curvaton brane involves a sound speed parameter which describes the propagation of the curvaton fluctuations during inflation, it has interesting implications to cosmological observations as well. Note that in our scenario, the inflaton field can be realized by any string realization of inflation which is independent of the curvaton field. Thanks to the virtue of curvaton scenario, the detailed realization of inflation does not affect the calculation for curvaton perturbation, as long as the inflaton perturbation itself is negligible. Our paper is organized as follows. In Section 2, we point out that the non-Gaussian features in a curvaton mechanism crucially depend on the curvaton decay mechanism. We suggest the curvaton could decay on a slice of uniform curvaton density, and non-Gaussian fluctuation evaluated on this slice is mainly determined by a combination of the equation-of-state (EoS) of the curvaton and a generalized transfer efficiency parameter which will be shown in the main text. In Section 3, we study the model of curvaton brane as a specific example. We firstly describe the background evolution of a curvaton brane model in relativistic limit throughout the inflationary period and the process of curvaton decay. Afterwards, we develop an analytic investigation of the curvature perturbation and the non-Gaussianity generated in this model by solving the perturbation equation of the curvaton brane order by order. In Section 4 we consider a class of plausible curvaton brane dominant eras and derive the nonlinearity parameters characterizing non-Gaussian distribution of the primordial curvature perturbation in these cases respectively. Finally, we present concluding remarks in Section 5. ", "conclusions": "In this paper, we have calculated the nonlinear primordial curvature perturbation in the curvaton scenario using a generalized $\\delta{N}$ formalism. If curvaton decay does not occur on the slice of uniform total energy density but on the slice of uniform curvaton density instead, we find that the dynamics of the nonlinear fluctuations behaves different from the usual scenario. Specifically, we consider a model of curvaton brane which can provide a theoretical realization of the curvaton decay we expected. In the frame of this model, we have presented a full analysis on the sizes and shapes of its bispectra and trispectra respectively. We arrive at an important conclusion that the generation of nonlinearities is sensitive to the mechanism of curvaton decay, which is mainly determined by the EoS $w$, the density occupation $\\Omega_{\\chi}$ of the field $\\chi$ at curvaton decay and its sound speed $c_s$ during inflation. Explicitly speaking, our results show that the nonlinearity parameter characterizing the second order perturbation $f_{NL}$ is proportional to $1/\\tilde{r}$ in which the transfer efficiency relies on the EoS $w$ and the density occupation $\\Omega_{\\chi}$ of the curvaton. For the equilateral shape, the size of $f_{NL}$ is amplified by a large value of the factor $1/c_s^2$, and therefore when the curvaton brane moves relativistically the shape of non-Gaussianity is mainly of equilateral type; however, at the squeeze limit the nonlinearity parameter $f_{NL}$ decouples from the sound speed and its size is qualitatively consistent with the result obtained in a usual canonical curvaton model. Moreover, for the third order fluctuations, we find the nonlinearity parameter $g_{NL}$ is enhanced by the factor $1/c_s^4$ in equilateral shape and decouples in local limit, and in the simplest example its size is proportional to $1/\\tilde{r}^2$ which is quite different from the usual scenario. As an end, we would like to highlight the importance of our study in this paper. The mechanism of curvaton decay after inflation could determine the relation of the primordial curvature perturbation and the decay hypersurface. This process is rather robust and should be considered in any specific curvaton model, and the example of curvaton brane analyzed in the current work is a good illustration to emphasize its importance. Additionally, the signatures of curvaton decay imprinted on the primordial curvature perturbation could provide a new window to explore the combination of early universe physics and astronomical observations. \\medskip" }, "1005/1005.5228_arXiv.txt": { "abstract": "We investigate damping and growth times of the quadrupolar $f$-mode for rapidly rotating stars and a variety of different polytropic equations of state in the Cowling approximation. This is the first study of the damping/growth time of this type of oscillations for fast rotating neutron stars in a relativistic treatment where the spacetime degrees of freedom of the perturbations are neglected. We use these frequencies and damping/growth times to create robust empirical formulae which can be used for gravitational wave asteroseismology. The estimation of the damping/growth time is based on the quadrupole formula and our results agree very well with Newtonian ones in the appropriate limit. ", "introduction": "\\label{sec:introduction} During the birth of a proto-neutron star or the merging of two older compact stars, violent non-radial oscillations may be excited, resulting in the emission of significant amounts of gravitational radiation \\cite{Andersson:2011lr}. The detection of gravitational waves from oscillating neutron stars will allow the study of their interior, in the same way as helioseismology provides information about the interior of the Sun. It is expected that the identification of specific pulsation frequencies in the observational data will reveal the true properties of matter at densities that cannot be probed today by any other experiment. In this paper, we present new empirical relationships for mode-frequencies and damping times of the quadrupolar $f$-mode for rapidly rotating neutron stars, extending previous studies which deal with the non-rotating case \\cite{1996PhRvL..77.4134A,1998MNRAS.299.1059A,2001MNRAS.320..307K}. These original suggestions about gravitational wave asteroseismology have been supported by many complementary works which studied specific features of oscillation spectra for various compact objects, such as typical neutron stars \\cite{1999MNRAS.310..797B,2002PhRvD..65b4010S,2004PhRvD..70l4015B,2005PhRvL..95o1101T,2006PhRvD..74l4025T,2007PhRvD..76d3003C,2007MNRAS.381..151W,Lau:2010lr}, but also for strange \\cite{2003PhRvD..68b4019S,2004PhRvD..69h4008S,Benhar:2007lr} or superfluid stars \\cite{2001PhRvL..87x1101A,2002PhRvD..66j4002A}. More recently, it has also been suggested that one may use asteroseismology to find the imprints of scalar or even vector components of gravity \\cite{2004PhRvD..70h4026S,2005PhRvD..71l4038S,2009PhRvD..79f4033S,2009PhRvD..80f4035S}. It should be noted here that all previous studies \\cite{Andersson:2011lr,1996PhRvL..77.4134A,1998MNRAS.299.1059A,2001MNRAS.320..307K,1999MNRAS.310..797B,2002PhRvD..65b4010S,2004PhRvD..70l4015B,2005PhRvL..95o1101T,2006PhRvD..74l4025T,2007PhRvD..76d3003C,2007MNRAS.381..151W,Lau:2010lr,2003PhRvD..68b4019S,2004PhRvD..69h4008S,Benhar:2007lr,2001PhRvL..87x1101A,2002PhRvD..66j4002A,2004PhRvD..70h4026S,2005PhRvD..71l4038S,2009PhRvD..79f4033S,2009PhRvD..80f4035S} have been performed for non-rotating relativistic stars. The treatment of rotation was always a problem in general relativity and thus the majority of the studies for the oscillation spectra of fast-rotating compact stars was done mainly in Newtonian theory which gives only qualitative answers. Since stellar oscillations may become unstable in the presence of rotation, there was an increased interest during the last decade or so to study the dynamics of rotating stars, also thanks to the discovery of the $r$- and $w$-mode instability \\cite{1998ApJ...502..708A,1998ApJ...502..714F,1998PhRvL..80.4843L,1999ApJ...510..846A,Kokkotas:2004lr}. Still, the majority of these studies have been performed in Newtonian theory \\cite{1999PhRvD..59d4009L,1999ApJ...521..764L,1999PhRvD..60f4006L} while there are only a few works in which GR has been used; mainly in the so called {\\em slow-rotation approximation}. The slow-rotation approximation was successfully applied to study various aspects of the $r$-mode instability \\cite{2001IJMPD..10..381A,2001MNRAS.328..678R}, effects of uniform and differential rotation on the oscillation spectrum \\cite{Ferrari:2004qy,2007PhRvD..75f4019S,2008PhRvD..77b4029P} and on the crustal modes \\cite{2008MNRAS.384.1711V}. As it has been originally suggested by Chandrasekhar \\cite{1970PhRvL..24..611C} and verified by Friedman \\& Schutz \\cite{1978ApJ...221..937F,1978ApJ...222..281F} certain non-axisymmetric pulsation modes may grow exponentially in rotating stars; this is due to the emission of gravitational waves and is called CFS instability. Exploring this type of instability in rapidly rotating stars turned out to be very difficult. In linear perturbation theory for example, rapid rotation was never treated properly until recently; almost all formulations of the relevant perturbation equations were prone to numerical instabilities either at the surface or along the rotation axis of the neutron star. Thus, it was not surpising that the first results for the oscillations of rapidly rotating stars were derived using evolutions of the non-linear equations \\cite{2004MNRAS.352.1089S,2006MNRAS.368.1609D,2006PhRvD..74f4024K,2008PhRvD..77f4019K}. Still all these studies were purely axisymmetric and thus the effects of rotation on the spectra was present only for very high rotation rates. Rotational instabilities are driven by non-axisymmetric modes and thus these first 2D calculations where not of much use for their study. In the last two years there was significant progress in the study of non-axisymmetric perturbations of rapidly rotating neutron stars. For the first time it was possible to calculate in GR the oscillation spectra of fast rotating relativistic stars by using the linearized form of the fluid equations. Thus the effect of fast rotation on $f$- and $r$-modes has been demonstrated while the critical points for the onset of the $f$-mode (CFS) instability have been derived \\cite{2008PhRvD..78f4063G}. In addition it has been demonstrated that there is a way to derive empirical relations connecting the oscillation frequencies with the rotation of the stars. This study has been recently extended to $g$-modes \\cite{2009PhRvD..80f4026G} and even more recently has been expanded to study the oscillation spectra of fast and differentially rotating neutron stars \\cite{2010PhRvD..81h4019K}. It should be noted that the previous results have been derived using the so-called Cowling approximation where the spacetime is assumed to be frozen. This approximation is very good for the estimation of the spectra of $r$- and $g$-modes but it gives only qualitatively good results for the $f$-mode. Moreover, using non-linear codes it became possible for the first time to study the complete problem \\cite{PhysRevD.81.084055}, i.~e.~the non-axisymmetric stellar oscillations of fast rotating stars without the constraints of the Cowling approximation. The results are in qualitative agreement with those found in \\cite{2008PhRvD..78f4063G} and for the critical point for the onset of the $f$-mode instability with the studies presented in \\cite{1998ApJ...492..301S}. The next step for gravitational wave asteroseismology is to use additional information about the damping times to construct model-independent relations which allow for a robust determination of stellar key parameters. The damping time $\\tau$ of the potentially CFS-unstable branch in the high rotation regime for example can be approximated very accurately by using a simple relation of the form $\\tau_0/\\tau = \\sgn{(\\sigma_i)}\\,0.256(\\sigma_i/\\sigma_0)^4$, where $\\tau_0$ and $\\sigma_0$ are the damping time and mode frequency of the nonrotating model respectively, $\\sgn{(x)}$ is the signum function and $\\sigma_i$ is the actual mode frequency in the inertial frame. The structure of the paper is as follows. In Section \\ref{sec:problem_setup} we give an essential overview about our method of computing mode-frequencies and damping times of the $f$-mode. We then show the results of our simulations in Section \\ref{sec:empirical_relations}, where we present empirical relations which can potentially be used to estimate masses, radii and angular frequencies of rapidly rotating neutron stars. A more elaborate discussion about the numerical procedure, the equations of state and background models used in this study as well as a consistency check can be found in the Appendix. ", "conclusions": "\\label{sec:summary} In this work we demonstrated how one can do gravitational wave asteroseismology by using the frequencies and possibly the damping/growth times of the emitted waves from oscillating and rapidly rotating relativistic stars. This is possible by the empirical relations that we have derived and which connect the frequencies and the damping/growth times of the oscillation modes with the stellar characteristics, i.e. with the mass, radius and rotation rate. We have actually shown that for polytropic equations of state of varying stiffness one can create very robust formulae connecting the observable frequency and damping times with the quantities like rotation frequency, average density and/or compactness. We have shown on a few examples how one can use the empirical formulae in order to derive the stellar parameters. In a realistic situation when an $f$-mode will be excited, it will be possible to detect the signal at least from galactic sources if the mode is CFS-stable and at least from sources in the Virgo cluster if it is unstable \\cite{Lai:1995fk,Ou:2004qy,Shibata:2004uq,2010KWK}. This will be possible with the sensitivity of the advanced Virgo and LIGO detectors \\cite{Acernese:2006fk,Abbott:2009qy} and probably even more feasible with the next generation gravitational wave telescopes such as ET (Einstein Telescope) \\cite{2010CQGra..27h4007P,Andersson:2011lr}. The ``weak'' point of the whole procedure relies in the approximate calculation of both frequencies and damping times. As we already mentioned we have neglected the spacetime perturbations and thus there is a systematic quantitative but not qualitative error in all data. Thus it is expected that the coefficients in relations \\eqref{eq:sigmaStableBranch}, \\eqref {eq:sigmaUnstableBranch} , \\eqref{eq:fitForDampTimeUnstable} and \\eqref{eq:fitForDampTimeStable} will be affected by a proper treatment of the spacetime degrees of freedom. However, it is believed that these changes will not alter the results significantly since the relations for frequencies and damping times are normalized with their corresponding values in the non-rotating limit. These values will absorb most of the differences when compared with the correct results in the presence of spacetime perturbations but of course, this is an issue that has to be addressed properly in future work. An additional outcome of this analysis will be the frequencies and damping times of the $w$-modes. Finally, the empirical relations found in this study have been derived for polytropes of varying stiffness which are able to mimic the global properties of realistic equations of state; see for example EoS A and EoS II which are polytropic fits to tabulated EoS. Realistic hot equations of state are the best candidates for newly born neutron stars and have a higher chance of becoming unstable prior mutual friction completely suppresses any instability \\cite{Andersson:2009lr} but currently our code is unable to perform time-evolutions of rapidly rotating neutron stars for generic tabulated data. Preliminary studies regarding this issue are promising but in a very early stage." }, "1005/1005.2064_arXiv.txt": { "abstract": "We estimated absolute shifts of Fe~I and Fe~II lines from Fourier-transform spectra observed in solar active regions. Weak Fe~I lines and all Fe~II lines tend to be red-shifted as compared to their positions in quiet areas, while strong Fe~I lines, whose cores are formed above the level $\\log \\tau_5\\approx-3$ (about 425~km), are relatively blue-shifted, the shift growing with decreasing lower excitation potential. We interpret the results through two-dimensional MHD models, which adequately reproduce red shifts of the lines formed deep in the photosphere. Blue shifts of the lines formed in higher layer do not gain substance from the models. ", "introduction": "We continue the study initiated by us in the papers \\cite{4,5} concerned with the effect of small-scale magnetic fields on solar granulation. The study is based on the observations made by P. Brandt with a Fourier-transform spectrometer (FTS) at the McMath telescope (Kitt Peak National Observatory, USA). The advantages of FTS observations are their high quality and a possibility to observe almost concurrently a large number of lines. Areas close to the solar disk center were observed. Here we investigate in great depth absolute shifts of Fe~I and Fe~II lines in low-activity areas (plages) with different integral magnetic fluxes. This is the first time that such an investigation covers a large number of lines formed at various levels in the solar photosphere so that effects of small-scale magnetic fields on height distribution of photospheric velocity field may be studied in detail. The present state of the investigations on small-scale solar magnetic fields is elucidated in review \\cite{20} by Solanki. Determinations of absolute shifts of absorption lines in active areas are not numerous, they are overviewed in \\cite{1}. Paper \\cite{7} by Brandt and Solanki, similar to \\cite{4,5} and the present paper, is also based on the FTS observations, but only 32 lines were used there to study line parameter variations and 19 lines to determine absolute shifts. Those 19 lines were strong Fe~I lines only. The conclusion was made that the lines in active areas had red shifts with respect to their positions in the quiet photosphere. The shifts depend on magnetic field strength (filling factor) and are, on the average, 0.22 pm (about 120~m/s) for lower bisector sections. This result agrees qualitatively with the findings of the studies by Cavallini et al. \\cite{8}, Livingston \\cite{17}, Immerschitt and Schroter \\cite{15}. At the same time Cavallini et al. \\cite{9} dealing with absolute shifts of four lines formed at different photospheric levels obtained a quite different result --- shifts of line cores with respect to their positions in the quiet photosphere depend on the region of line core formation: the weak Fe~I line at $\\lambda$ 614.92~nm displays a red shift, while the strong Ca~I line at $\\lambda$ 616.22~nm is blue-shifted. A similar result was obtained by Keil et al. \\cite{16} for the strong Fe~I $\\lambda$ 543.45~nm line with a small lower excitation potential (1.01~eV), the core of this line is formed in the upper photospheric layers. Absolute shifts of the line were found to depend on magnetic field polarity, but when the shifts are averaged without regard for local field signs, the line core has a blue shift of about 150~m/s. Evidence of different behavior of velocity fields with height in the photosphere in active and quiet areas is found also in the analysis of spectral observations with high spatial resolution made by Hanslmeier et al. \\cite{13,14}. If this is the case, lines formed at different levels in the atmosphere should display different absolute shifts. Thus, the problem needs to be thoroughly investigated on the basis of an extended sample of spectral lines. This is the prime objective of the present study. ", "conclusions": "FTS observations were used to estimate possible absolute shifts of Fe~I and Fe~II lines in active regions. We conclude that weak Fe~I lines and all Fe~II lines are likely to be red-shifted with respect to their positions in the quiet photosphere, while strong Fe~I lines with low $EPL$s, their cores being formed above the level $\\log \\tau_5\\approx-3$ ($\\approx$425~km), have more violet shifts. The results were interpreted on the basis of 2-D MHD models that reproduce quite adequately the red shifts of the lines formed deep in the photosphere, but violet shifts of the lines formed in the upper photospheric layers are not confirmed within the scope of these models. \\vspace{1.cm} {\\bf Acknowledgements.} We wish to thank the National Optical Astronomical Observatory/National Solar Observatory (Tucson, Arizona, USA) and especially J. Brault and B.~Graves for support in the FTS observations; S. Solanki for filling factor calculations; R.~Rutten for assistance in the organization of this investigation. The study was partially financed by the Joint Foundation of the Government of Ukraine and International Science Foundation (Grant No. K11100)." }, "1005/1005.3987_arXiv.txt": { "abstract": "White dwarfs inspiraling into black holes of mass $\\MBH\\simgt 10^5M_\\odot$ are detectable sources of gravitational waves in the LISA band. In many of these events, the white dwarf begins to lose mass during the main observational phase of the inspiral. The mass loss starts gently and can last for thousands of orbits. The white dwarf matter overflows the Roche lobe through the $L_1$ point at each pericenter passage and the mass loss repeats periodically. The process occurs very close to the black hole and the released gas can accrete, creating a bright source of radiation with luminosity close to the Eddington limit, $L\\sim 10^{43}$~erg~s$^{-1}$. This class of inspirals offers a promising scenario for dual detections of gravitational waves and electromagnetic radiation. ", "introduction": "One of the goals of the Laser Interferometer Space Antenna (LISA) mission is to detect gravitational waves from compact stellar objects spiraling into massive black holes, a class of events called extreme-mass-ratio inspirals (EMRIs) (e.g., Hils \\& Bender 1995; Gair et al. 2004; Barack \\& Cutler 2004; see Hughes 2009 and Sathyaprakash \\& Schutz 2009 for recent reviews). Of particular interest are sources of coincident gravitational and electromagnetic radiation. Besides providing unique information on the nature of the event, such dual detections will lead to a new version of Hubble diagram that is based on the gravitational distance measurements (e.g. Bloom et al. 2009; Phinney 2009). Inspirals into black holes of masses $\\MBH\\sim (10^5-10^6)M_\\odot$ produce gravitational waves in the frequency band where LISA is most sensitive. Normal stars are tidally disrupted well before they approach such black holes, and therefore discounted as possible LISA sources (however, see Freitag 2003). Inspirals of compact objects are guaranteed sources of gravitational waves, however most of them are not promising for dual detections. In particular, stellar-mass black hole or neutron star inspirals are not expected to generate bright electromagnetic signals. Only inspiraling white dwarfs (WDs) offer a possibility for dual detection (Menou et. al. 2008; Sesana et al. 2008). WDs can be tidally disrupted very close to the black hole and then could create a transient accretion disc with Eddington luminosity. Estimated EMRI rates are high enough for observations with LISA (e.g. Phinney 2009). The expected fraction of WD inpirals among all EMRIs depends on the degree of mass segregation in galactic nuclei, which favors stellar-mass black holes over white dwarfs in the central cluster. The abundance of white dwarfs also depends on the details of stellar evolution, which are not completely understood. A fraction of WD inpirals as large as $\\sim 10\\%$ has been suggested (e.g. Hopman \\& Alexander 2006a). The orbital parameters of WD inspirals are also uncertain. EMRIs form when a compact object is captured onto a tight orbit whose evolution is controlled by gravitational radiation rather than random interactions with other stars in the central cluster around the massive black hole. Two main channels exist for EMRI formation: capture of single stars and capture of binary systems (e.g., Hils \\& Bender 1995; Sigurdsson \\& Rees 1997; Ivanov 2002; Gair et al. 2004; Hopman \\& Alexander 2005, 2006a,b, 2007; Miller et al. 2005; Hopman 2009). In the single-capture scenario, the shrinking orbit can retain a significant eccentricity until the end of inspiral. In contrast, the binary-capture scenario leads to nearly circular orbits (Miller et al. 2005). In this paper, we focus on WD inspirals that are not completed because the star is tidally disrupted before its orbit becomes unstable. We argue that the WD begins to lose mass very gently and, for thousands of orbital periods, this process resembles accretion through the $L_1$ point in a binary system rather than a catastrophic disruption. This offers a possibility of {\\it simultaneous} observation of the inspiral by LISA and traditional, optical and X-ray telescopes. Previous work on tidal deformation of a WD orbiting a massive black hole focused on two extreme regimes: (i) weak deformation was studied analytically using perturbation theory (e.g., Rathore et al. 2005; Ivanov \\& Papaloizou 2007), and (ii) strong deformation leading to immediate disruption was simulated numerically (e.g. Kobayashi et al. 2004; Rosswog, Ramirez-Ruiz \\& Hix 2009). The regime considered in the present paper is different from both cases explored previously. It involves an extended phase of strong deformation with small mass loss, which we call `tidal stripping' below. The mass of the WD remains almost unchanged during this phase and continues to emit gravitational waves. Even a small orbital eccentricity, e.g. $e=0.01$, implies that tidal stripping occurs only near the pericenter of the orbit, during a small fraction of each orbital period. Thus, the mass loss is expected to be {\\it periodic}, possibly leading to a periodic electromagnetic signal from the inspiral. ", "conclusions": "The standard picture of EMRI envisions an orbit that gradually shrinks due to gravitational radiation until its pericenter $r_p$ reaches $\\rmin$ where the orbit becomes unstable and plunges into the black hole. The main observational phase of the inspiral is when $r_p$ decreases from $\\sim 2\\rmin$ to $\\rmin$. For WDs with mass $M\\sim M_\\odot$ inspiraling into a black hole with $\\MBH\\sim 10^5M_\\odot$ the main observational phase lasts $\\sim 10^5$ orbital periods, which may be comparable to one year, depending on the orbital eccentricity. We argued in Section~2 that the inspiraling WDs can experience an extended period of slow mass loss during observations by LISA. To summarize, the tidal stripping begins very gently because the pericenter of the orbit drifts inward slowly, by a tiny fraction $\\alpha\\sim 10^{-5}$ in one orbital period. This leads to many repeated episodes of small mass loss. The process occurs at radius $r_0$ (eq.~\\ref{eq:r0}), comparable to $\\rmin$. The star is strongly deformed by the tidal forces near $r_0$ but it barely touches its Roche lobe for a small fraction of the orbital period. As a result, the star loses a small amount of mass through the $L_1$ point at each pericenter passage. To our knowledge, this regime of tidal stripping was not explored by direct hydrodynamical simulations. Our crude estimates suggest two phases of stripping: the first phase lasts $N_1\\sim 10^3$ orbits until the star loses $\\sim 1$ per cent of its mass. Then the mass loss accelerates: the decrease in $M$ (and the corresponding increase in the WD radius) implies that the star overfills the Roche lobe more and more with every pericenter passage. As a result, the star loses the remaining 99 per cent of its mass in a few hundreds of additional orbits. The mass lost in each individual episode during these last hundreds of orbits behaves as $\\dM\\propto (N_\\star-N)^{-\\xi}$ where $N_\\star-N\\ll N_\\star$ is the number of orbits remaining to complete disruption and $\\xi\\sim \\psi/(\\psi-1)$. According to our estimates, tidal stripping operates for days or weeks, creating a relatively long-lived source of gas. The gas can accrete onto the black hole and produce significant electromagnetic radiation together with the gravitational waves observed by LISA. The radiation source may become bright before the mass loss spoils the standard EMRI template for the gravitational-wave signal. The gas produced by tidal stripping moves on nearly Keplerian orbits that are initially close to the WD orbit. The gas probably leaves the star with relative velocity $\\sim v_{\\rm esc}=(2GM/R)^{1/2}$ and its orbital energy differs from that of the star by a small fraction $\\sim v_{\\rm esc}/v\\sim(Mr_0/\\MBH R)^{1/2} \\sim\\gamma^{-1/2}(M/\\MBH)^{1/3}\\sim 1/30$. After $\\sim 30$ orbital periods, the differential rotation of the gas has stretched it into a ring around the black hole. A non-zero eccentricity of the donor orbit will create an eccentric ring. It should viscously spread and accrete onto the black hole. A small mass-loss fraction $x$ can be a huge source of gas for accretion. If most of the stripped matter is accreted by the black hole, the accretion rate is $\\dot{M}\\sim \\dM/P$. It exceeds the Eddington value $\\dot{M}_{\\rm Edd}\\sim 10^{23}(\\MBH/10^5M_\\odot)$~g~s$^{-1}$ after $N\\sim 10$ orbits since the beginning of tidal stripping, well before the final disruption of the WD. The accretion timescale in the viscous ring can be estimated as $\\tacc\\sim \\alpha_v^{-1}(H/r)^{-2}P$ where $P$ is the WD orbital period, $\\alpha_v=0.01-0.1$ is the viscosity parameter and $H$ is the thickness of the ring (e.g. Shakura \\& Sunyaev 1973). $H/r\\sim 1$ is expected for super-Eddington accretion, which leads to $\\tacc\\sim (10-100)P$. This suggests that a bright source with Eddington luminosity $L_{\\rm Edd}\\sim 10^{43}$~erg~s$^{-1}$ is created quickly, in less than 1 day after the beginning of tidal stripping. For a typical distance to such LISA sources, $d\\sim 100$~Mpc, the accretion ring should be detectable with optical and X-ray telescopes provided its approximate location on the sky is known. LISA is expected to localize EMRIs within $\\sim 10$~deg$^2$ (Barack \\& Cutler 2004). For massive black-hole mergers, the localization information will be available weeks to months prior to the final coalescence (Kocsis et al. 2007, 2008; Lang \\& Hughes 2008), and the localization expectations for WD EMRIs are similar (S. Drasco, J. Gair, I. Mandel, E. Porter, private communications), giving sufficient time for simultaneous optical and X-ray observations during inspiral. Mass transfer in WD inspirals is special as it creates a long-lived source of gas very close to the black hole horizon, $r_0-r_g\\sim r_g$. As a result, the donor orbit is generally not confined to a plane, if the black hole rotation is significant. Besides, the orbit will experience fast precession. Therefore, the freshly stripped gas may collide with the previously released gas and generate shocks (e.g. Evans \\& Kochanek 1989). The resulting pattern of accretion may be complicated and needs careful study. An intriguing feature of tidal stripping is the periodic supply of gas. It may leave a fingerprint on the observed luminosity, modulating it with the WD orbital period $P$. The modulation of $\\dot{M}$ may create a detectable oscillation in the luminosity from the accreting ring, even though the accretion timescale $\\tacc\\sim (10-100)P\\gg P$, tends to reduce the amplitude of modulation. The shock emission from collisions between the periodic flow from the $L_1$ point and the gas accumulated around the black hole may be strongly modulated. The description of mass transfer in this paper is greatly simplified. The possibility of many repeated mass-transfer episodes is robust, but the exact rate of tidal stripping and the dynamics of accretion need to be explored with dedicated numerical simulations. The great potential that such events hold for joint detections of gravitational and electromagnetic radiation provides motivation for the effort. The joint detection would let us witness, in real-time and unprecedented detail, the slow tidal stripping of a WD followed by its complete disruption. Note that LISA observations are expected to provide the mass and spin of the black hole, as well as the details of the inspiral orbit. This can be used to model in detail the hydrodynamics of accretion. The scenario discussed in this paper assumes $r_0>\\rmin$ and uses a semi-Newtonian description for the WD orbit. A fully relativistic model will be needed to accurately evaluate the parameter space for such events. The relativistic effects are especially important if the black hole is rapidly rotating -- then $r_0$ and $\\rmin$ will depend on the black hole mass $\\MBH$, its spin parameter $a_s$, the eccentricity of the orbit, and the angle between the orbital angular momentum and the angular momentum of the black hole. The competition between tidal disruption and gravitational capture by rotating black holes was investigated for parabolic orbits in Beloborodov et al. (1992). In a broad range of $\\MBH$, the fate of a star approaching the black hole depends on the orbit orientation and can be either disruption or capture. For inspirals with $r_g 1200$, which determines the sound speed for acoustic fluctuations, (3) the evolution of the ionization fraction of the universe at $z < 1200$, which determines the thickness of the surface of last scatter and (4) the transition from radiation to matter domination. Moreover, while small-scale CMB fluctuations are initially pure $E$ mode, gravitational lensing rotates $E$ modes into $B$ modes \\cite{cmblensing}. By measuring the pattern of small scale $E$ and $B$ modes, cosmologists will be able to determine the large-scale convergence field, a direct measure of the integrated density fluctuations between redshift $z=1100$ and $z=0$ (see e.g. \\cite{hirata},\\cite{okamotohu}). The convergence power spectrum is particularly sensitive to density fluctuations at $z \\sim 2$, an important complement to planned optical lensing measurements that probe the evolution of density fluctuations in the $z < 1$ universe. The goal of this paper is to quantify the cosmological information that could come from these new datasets. This is important for several reasons. First, while there have been many studies of the future cosmological constraints from Planck, very few papers have investigated the constraining power of combinations of future CMB datasets from different sources. Second, as we will describe in the next sections, we will consider a large set of parameters focusing on those that mainly affect the \"damping tail\" of the CMB angular spectrum. We consider additional parameters such as the total neutrino mass $\\sum m_{\\nu}$ (which affects the growth of structure in the late universe), the number of additional relativistic neutrino species $N_{\\nu}^{eff}$ (which changes the matter-radiation epoch), and possible changes in the recombination process due to changes in the fractional helium abundance $Y_p$, dark matter self-annihilation processes, and variations in fundamental constants such as as the fine structure constant $\\alpha$ and Newton's gravitational constant $G$. We will not only show the constraints on each single parameter but also the degeneracies among them. We will consider $3$ experimental configurations: the Planck satellite \\cite{planck}, the combination of Planck with ACT fitted with polarization-sensitive detectors, ACTPol, \\cite{ACTPol} and, finally, the next CMBPol satellite \\cite{CMBPol}. Recent studies have already fully demonstrated the ability of next generation satellite missions to constrain inflationary parameters \\cite{baumann} and the reionization history \\cite{cmbpolreio} in the framework of the CMBPol concept mission study (see also \\cite{dode}). For this reason we will not consider primordial gravitational waves, more general reionization scenarios or experiments that will mainly probe large angular scale polarization in this paper. This paper will show that next generation CMB experiments can significantly improve constraints on cosmology and fundamental physics and could produce a detection of neutrino mass. The paper is structured as follows. Section II describes our analysis approach. Section III presents our analysis for improved constraints from the planned ACTPol experiment and for the proposed CMBPol experiment. In Section IV we present our conclusions. ", "conclusions": "In this paper we have performed a systematic analysis of the future constraints on several parameters achievable from CMB experiments. Aside from the $5$ parameters of the standard $\\Lambda$-CDM model we have considered new parameters mostly related to quantities which can be probed in a complementary way in the laboratory and/or with astrophysical measurements. In particular we found that the Planck experiment will provide bounds on the sum of the masses $\\Sigma m_{\\nu}$ that could potentially definitively confirm or rule out the Heidelberg-Moscow claim for a detection of an absolute neutrino mass scale. Planck+ACTPol could reach sufficient sensitivity for a robust detection of neutrino mass for an inverted hierarchy, while CMBPol should also be able to detect it for a direct mass hierarchy. The comparison of Planck+ACTPol constraints on baryon density, $N_{eff}$ and $Y_p$ with the complementary bounds from BBN will provide a fundamental test for the whole cosmological scenario. CMBPol could have a very important impact in understanding the epoch of neutrino decoupling. Moreover, the primordial Helium abundance can be constrained with an accuracy equal to that of current astrophysical measurements but with much better control of systematics. Constraints on fundamental constants can be achieved at a level close to laboratory constraints. Such overlap between cosmology and other fields of physics and astronomy is one of the most interesting aspect of future CMB research. We should note, however, that our forecasts rely on several technical assumptions. First, we assumed that the theoretical model of the recombination process is accurately known. This is not quite true as corrections to the recombination process are already needed for the Planck experiment (see e.g. \\cite{rubino}). However, this is mainly a computational problem that could be solved by the time of CMBPol launch, expected not before 2015. In addition, we assume that the foreground and beam uncertainties are smaller than the statistical errors. Nevertheless, the results clearly show the advantage of adding small scale data from the ACT telescope to the Planck satellite data. Adding the former to the latter will improve the constraints by a factor $\\sim2$ in most of the models considered." }, "1005/1005.4934_arXiv.txt": { "abstract": "Nearly all globular clusters (GCs) studied to date show evidence for multiple stellar populations, in stark contrast to the conventional view that GCs are a mono-metallic, coeval population of stars. This generic feature must therefore emerge naturally within massive star cluster formation. Building on earlier work, we propose a simple physical model for the early evolution (several $10^8$ yr) of GCs. We consider the effects of stellar mass-loss, type II and prompt type Ia supernovae, ram pressure, and accretion from the ambient interstellar medium (ISM) on the development of a young GC's own gas reservoir. In our model, type II SNe from a first generation of star formation clears the GC of its initial gas reservoir. Over the next several $10^8$ yr, mass lost from AGB stars and matter accreted from the ambient ISM collect at the center of the GC. This material must remain quite cool ($T\\sim10^2$K), but does not catastrophically cool on a crossing time because of the high Lyman-Werner flux density in young GCs. The collection of gas within the GC must compete with ram pressure from the ambient ISM. After several $10^8$ yr, the Lyman-Werner photon flux density drops by more than three orders of magnitude, allowing molecular hydrogen and then stars to form. After this second generation of star formation, type II SNe from the second generation and then prompt type Ia SNe associated with the first generation maintain a gas-free GC, thereby ending the cycle of star formation events. Our model makes clear predictions for the presence or absence of multiple stellar populations within GCs as a function of GC mass and formation environment. While providing a natural explanation for the approximately equal number of first and second generation stars in GCs, substantial accretion from the ambient ISM may produce fewer chemically peculiar second generation stars than are observed. Analyzing intermediate-age LMC clusters, we find for the first time evidence for a mass threshold of $\\sim10^4\\Msun$ below which LMC clusters appear to be truly coeval. This threshold mass is consistent with our predictions for the mass at which ram pressure is capable of clearing gas from clusters in the LMC at the present epoch. Recently, claims have been made that multiple populations within GCs require that GCs form at the center of their own dark matter halos. We argue that such a scenario is implausible. Observations of the young and intermediate-age clusters in the LMC and M31 will provide strong constraints on our proposed scenario. ", "introduction": "\\label{s:intro} Globular clusters (GCs) have historically been considered coeval, mono-metallic, gravitationally bound collections of stars. In the past several years, high precision photometric and spectroscopic observations have led to a radical revision of this picture. High resolution spectra of stars within nearly all GCs studied to date reveal internal spreads in light element abundances such as C, N, Na, O, Mg, and Al, beyond what can be explained by measurement errors \\citep[see review in][]{Gratton04}. The magnitude of the internal spread varies considerably from cluster to cluster, though there are noticeable trends with cluster mass and orbital properties \\citep{Carretta06, Carretta10b}. Intriguingly, internal spread in the Fe-peak elements and the type II supernovae product Ca is limited to only the most massive GCs \\citep[$\\omega$Cen, M22, M54, and Terzan 5;][]{Marino09, Carretta09a, Ferraro09, Carretta10a, Carretta10b}. The most massive clusters also show indirect evidence for a large spread in He abundances \\citep{Gratton10}. Abundance variations have been detected in main sequence stars \\citep{Gratton01}, indicating that the observed variation arises from stars forming out of different material, as opposed to being due to some unknown mixing process, which could only occur along the giant branch. While much attention has been paid to internal abundance spreads recently, such internal spreads have been known for over 30 years \\citep{Cohen78, Kraft79, Peterson80, Freeman81, Smith82a, Smith83}. Photometry from the {\\it Hubble Space Telescope (HST)} has demonstrated that many GCs contain multiple sub-giant branches and at least two ($\\omega$Cen and NGC 2808) contain multiple main sequences \\citep[see review in][]{Piotto09}. Variations in the light element abundances have been conclusively associated with the multiple sequences observed in the color-magnitude diagram \\citep[CMD;][]{Yong08, Marino08, Carretta09b, Milone10}, demonstrating that these two phenomena are intimately linked. The population with enhanced abundance patterns is more centrally concentrated than the `normal' population in the GCs NGC 1851 \\citep{Zoccali09} and $\\omega$Cen \\citep{Sollima07}, and perhaps many others as well \\citep{Carretta09b}. In the old Milky Way (MW) GCs the relative numbers of stars with `anomalous' and normal abundances ratios is approximately equal, with little dependence on metallicity \\citep[e.g.,][]{Smith82a, Carretta09b}. This fact imposes strong constraints on formation scenarios, as we explain in later sections. The relative numbers of anomalous and normal stars in GCs in other galaxies (e.g., the LMC, M31) is not known, but would provide new insight. Additional insight has come from the strong observed correlations between various light element abundances within individual GCs. The most striking is the anti-correlation between Na and O \\citep[e.g.,][]{Kraft93, Ivans99, Carretta09b}. This correlation arises naturally when material of standard (e.g., solar) abundance ratios is mixed with material that has been processed\\footnote{Throughout the text we will use `processed' and `un-processed' to refer specifically to the nuclear processing of material at the temperatures required to explain the observed light element abundance variations, i.e., at $T>10^7$K.} at temperatures $>10^7$K. At such temperatures the Na-Ne and CNO cycles are active, with the former producing Na and the latter depleting O. At similar temperatures the Mg-Al cycle is active, explaining observed correlations between these elements as well. The discovery of significant Li and F abundances in the second stellar generation \\citep[i.e., associated with stars of high Na and low O abundances;][]{Pasquini05, Smith05} has provided yet another puzzle and clue. These elements are fragile, especially Li which burns at $T\\gtrsim10^6$K. The existence of such fragile elements in the atmospheres of second generation stars that also show strong O depletion and Na enhancement suggests that these stars formed out of at least two kinds of material; i.e., from matter exposed to $T>10^7$K, and additional material that was never heated above $T\\sim10^6$K. The interpretation of Li is however complicated by the possibility that Li may, under special circumstances, actually be produced within AGB stars \\citep[see e.g.,][]{Ventura10}. Multiple stellar populations have also been detected in intermediate-age ($\\sim1$ Gyr) and old LMC clusters \\citep{Mackey08, Goudfroij09, Milone09, Mucciarelli09}, and possibly in the old GCs within the Fornax dSph galaxy \\citep{Letarte06}. These observations indicate that the multiple stellar population phenomenon is not specific to the MW. Observations of LMC clusters are of course hampered by the much larger distance modulus to the LMC. Despite this limitation, the intermediate-age clusters offer a new window into the internal age spreads because the main sequence turn-off point at these ages is a strong function of time. Small age differences are therefore readily noticeable in the CMD. Analysis of {\\it HST}-based CMDs have shown that the spread observed in the main sequence turn-off of intermediate-age LMC clusters can be explained with an internal age spread of a few $10^8$ yr. The only class of star clusters known {\\it not} to contain multiple stellar populations are the open clusters in the MW \\citep{deSilva09, Martell09}, which have typical masses of a few thousand solar masses, although some, such as NGC 7789, have masses of $\\sim10^4\\Msun$. These observations have led to the unavoidable conclusion that the majority of GCs studied to date harbor multiple stellar populations. For all but the most massive ones, GCs are still considered chemically homogeneous in Fe-peak and elements arising primarily from SNe type II such as Ca. The light element variations have been detected in both young and old clusters, both metal-poor and metal-rich \\citep{Martell09}, and are noticeably absent in the open clusters. From these observations the following timeline in the early evolution of GCs has emerged. Within GCs, a first generation of stars form. Type II SNe then remove any remaining gas from the GC. After the epoch of type II SNe, mass from evolved stars is cycled through temperatures of $T>10^7$K and then is returned to the gaseous reservoir of the GC. After a few $10^8$ yr, a second generation of stars forms from a mix of processed and un-processed material. Star formation permanently ceases after the formation of the second generation. This process does not occur in open clusters, nor in the field. This basic timeline has, in one form or another, been discussed by many authors \\citep[e.g.,][]{Cottrell81, Smith87, Carretta10c}. Any theory of GC formation must be embedded into our broader cosmological theories of structure formation. Many lines of evidence suggest that galaxy formation is an hierarchical process where dark matter halos serve as sites for assembling baryons and converting baryons into stars. One of the inevitable predictions of this bottom-up scenario is that MW GCs likely formed in environments very different from the $\\sim200\\kms$ dark matter dominated halo where they now reside. One possibility, which we discuss in $\\S$\\ref{s:dm}, is that GCs form in the centers of small dark matter halos. This scenario would imply that isolated GC are embedded in extended dark matter halos. This paper emphasizes another possibility: GCs form within small gas-rich dwarf galaxies --- the building blocks of the present MW. We emphasize that the MW is an evolving galactic system so that estimates of ram pressure and gaseous accretion must consider the likely environment in which a GC formed at $z\\sim2-10$, rather than its current location in the MW today. We have outlined above only the most basic sketch of what must occur to explain the observations. In the next section we critically assess previous, more detailed scenarios for the development of multiple stellar populations within GCs. Following this assessment, we describe our own model for the early evolution of GC stellar populations that includes several novel ingredients, and is, at least in certain respects, more plausible than other scenarios. We also present an analysis of intermediate-age clusters in the LMC that provides confirmation of a key aspect of our model. We conclude by commenting on various observations that may shed new light on this exciting observational puzzle. ", "conclusions": "\\label{s:res} In this work we have presented a comprehensive model for the early (several $10^8$ yr) evolution of massive star clusters. Our model considers the importance of type II and prompt type Ia SNe, accretion onto the GC from the ambient ISM, the ability of a GC to retain its internal gas supply in the face of ram pressure, and the effect of the Lyman-Werner photon flux density on the ability of the young GC gas to form molecules and, ultimately, stars. This model definitively addresses two of the three major issues in early GC evolution identified in $\\S$\\ref{s:prev}: how GCs can retain a gaseous reservoir in the face of ram pressure, and which stars are responsible for processing material to $T>10^7$K. From consideration of the formation environments of both old MW GCs and intermediate-age LMC clusters, we have shown that ram pressure stripping naturally explains the observed bifurcation between clusters that do and do not show evidence for multiple stellar populations. Based on several independent arguments, including the similar $\\sim10^8$ yr timescale of many physical mechanisms and the lack of internal spread in Fe and Ca in most clusters, we strongly favor massive AGB stars as the source of the processed material. The final open issue is in understanding how a second generation can form with a current total mass comparable to the first generation, at least for old MW GCs where data are available. We have considered accretion from the ambient ISM as a viable mechanism to provide copious amounts of gas to the GC. However, our solution to this last issue currently has little direct empirical verification, and may in fact have trouble reproducing the observed correlations in abundance ratios, as discussed below. We now discuss a variety of testable implications of our proposed solution, and comment briefly on several open issues. As a consequence of significant accretion from the ambient ISM, we expect several trends with GC mass (see e.g., Figure \\ref{fig:facc}). For example, as the fraction of pristine gaseous material increases, we expect a shorter Na-O correlation, a greater abundance of Li and F, and in general we expect the stars with anomalous abundances to be less anomalous. The sign of the trend with mass depends on the dominant accretion process, which in turn depends on the formation environment. Bondi accretion scales as $M^2$ while the mass provided by AGB winds scales with $M$ so the pristine fraction increases with mass. ISM sweeping scales as $R^2$ which is only weakly correlated with $M$ at the present epoch, and so if this process is dominant we would expect the pristine fraction to decrease with mass. Observations of young and intermediate-age clusters will allow identification of the relevant mechanism because the formation environment of young clusters is not so dissimilar from their present environment. The old GCs within dwarf spheroidals and dwarf irregulars may also shed light on this issue. The GCs within dwarf galaxies almost certainly formed where they are now observed since the stellar accretion/merger rate onto dwarfs is expected to be very low in a $\\Lambda$CDM cosmology, and so the present day conditions of the host dwarf galaxies cannot be very different from the formation environment of the GCs. \\citet{Carretta06} and \\citet{Carretta10c} have investigated the extent of the Na-O correlation as a function of global parameters and have found evidence that the correlation is more extended in higher mass GCs, and for GCs with more extended orbits. This result is consistent with ISM sweeping being the dominant accretion process, but we caution that the data show large scatter, and the present GC mass is poorly correlated with its mass at formation due to orbit-dependent mass-loss effects. In general, we expect that the extent of the Na-O correlation should depend on the formation environment at fixed GC mass, since the formation environment determines the amount of accreted material. Another effect might be the mass-dependent ability of a GC to retain the winds from AGB stars. In any event, it is clear that such trends hold the promise of isolating which physical process dominates the accretion of pristine material onto GCs. If accretion from the ambient ISM was important, then it is somewhat puzzling why the Fe and Ca abundances are so uniform between the first and second generations. The only plausible explanation is that at early times, during the formation of the ancient MW GCs, the spread in Fe and Ca abundances within the MW progenitor system was quite small. A clear prediction of the accretion scenario is that the intermediate-age LMC clusters showing multiple populations should show a much larger internal spread in Fe, Ca, etc. abundances, on the order of the spread in abundances of these elements within the LMC as a whole. The MW bulge GC Terzan 5 may provide additional insight. \\citet{Ferraro09} has convincingly shown that this GC has a split horizontal branch (HB) with the more luminous branch having a much higher Fe abundance ([Fe/H]$\\sim+0.3$) than the less luminous branch ([Fe/H]$\\sim-0.2$). \\citet{DAntona10} argue that the split HB is consistent with an internal age spread of several $10^8$ yr and enhanced He in the metal-enriched branch. Since this cluster is in the MW bulge and Fe-rich, it probably formed in the bulge. D'Antona et al. suggest that the second stellar generation may have acquired its high Fe abundance via accretion from the ambient, Fe-rich ISM. Self-enrichment is unlikely because the required Fe mass to produce the observed high Fe abundance would require so many type II SNe that the cluster would easily become unbound, unless Terzan 5 was significantly more massive in the past. Terzan 5 may therefore be an ideal cluster to look for further evidence for the importance of ambient ISM accretion in the formation of multiple stellar populations, e.g., by observing the Li and F abundances in this cluster. If the proposed scenario for Terzan 5 is generic then we might expect to find significantly separated Fe abundances within the other bulge metal-rich GCs as well. In the present work we have not attempted to make specific predictions for the extent of correlations amongst various elemental abundances. Such predictions would provide a powerful constraint on the model. Qualitative considerations suggest that substantial accretion from the ISM may yield a very short Na-O anti-correlation, owing to the ISM having abundance ratios similar to the first generation stars. Unfortunately, uncertainties in the AGB yields greatly complicate any attempted comparison to observed abundances. A fruitful avenue for future work will consist of a detailed comparison between predicted and observed abundance correlations in light of the uncertain AGB yields \\citep[see][for an initial attempt in this direction]{DErcole10b}. The most massive GCs, including $\\omega$Cen, NGC 2808, M22, and M54, deserve special mention. Each of these clusters show multiple, {\\it distinct} sequences in the CMD, and the former two are unique in that they display multiple main sequences. The CMD morphology of these clusters suggests very high He abundances in the second (and third) stellar generations. The discrete sequences and high He abundances argue against significant dilution from ambient ISM accretion. This can be accommodated in our model if ISM sweeping is the dominant accretion mechanism because this mechanism becomes increasingly less important at high GC masses (see the bottom panels of Figure \\ref{fig:facc}). Although we are then left with the original problem of explaining how so many second generation stars could form from the mass lost by the first generation. We speculate that perhaps the most massive GCs formed at the centers of their own dark matter halos, and contained even more stars in their past. The precursors of these systems could be nuclear star clusters, which are common in low-mass galaxies. Regardless of these details, we stress caution when attempting to incorporate the massive GCs into any framework for the early evolution of GCs since many of their observational characteristics differ qualitatively from the lower mass GCs. An exciting direction for future observational work is characterizing the incidence of multiple populations within the numerous young and intermediate-age clusters in M31. These clusters span a range in mass from $10^3\\Msun$ to $10^5\\Msun$ \\citep{Caldwell09}. Based on ram pressure arguments, we expect the higher mass clusters to show evidence for multiple populations, but not the lower mass clusters. Assuming $n=1\\cm$ and $V=250\\kms$ for disk of M31 \\citep{Widrow03}, we predict that the critical mass will be $\\approx10^{4.5}\\Msun$. However, as stressed in previous sections, the relevant velocity is the relative velocity between the ambient ISM and the GC. Since many of the young M31 GCs are orbiting within the disk of M31, their relative velocity will be considerably less than the circular velocity, and so the critical mass may be considerably lower. Confrontation of our model predictions with the properties of the M31 GCs must take these details into account. Since these M31 clusters have not experienced significant dynamical evolution, we may expect stronger correlations between the extent of any abundance anomalies and cluster mass. For the youngest clusters, with ages of several $10^8$ yr, we might even hope to {\\it directly observe} the formation of the second generation. Such an observation would provide definitive proof that GCs need not form within dark matter halos to produce multiple stellar generations, and would also demonstrate that AGB stars are the polluters, since the AGB polluter scenario is expected to produce a second generation on a timescale of several $10^8$ yr. In the past several years it has become clear that star clusters, once thought to be simple systems, in fact show an internal complexity that increases with increasing mass. Observations indicate that the lowest mass systems, i.e., the open clusters, appear to be truly coeval and mono-metallic. At higher masses one observes age spreads of several $10^8$ yr and internal variation in the light elements. At still higher masses ($\\gtrsim10^6\\Msun$) it appears that the cluster potential well is deep enough to retain SNe ejecta and hence self-enrich, perhaps because these massive clusters form at the center of their own dark matter halos. This last phenomenon is observed not only in the most massive MW GCs, but also appears to be present in the most massive GCs within other galaxies as well \\citep{Strader08, Bailin09}. At the very least, it is now abundantly clear that star clusters are not simple systems." }, "1005/1005.1720_arXiv.txt": { "abstract": "In this letter, we study in detail the evolution of the post-flare loops on 2005 January 15 that occurred between two consecutive solar eruption events, both of which generated a fast halo CME and a major flare. The post-flare loop system, formed after the first CME/flare eruption, evolved rapidly, as manifested by the unusual accelerating rise motion of the loops. Through nonlinear force-free field (NLFFF) models, we obtain the magnetic structure over the active region. It clearly shows that the flux rope below the loops also kept rising accompanied with increasing twist and length. Finally, the post-flare magnetic configuration evolved to a state that resulted in the second CME/flare eruption. This is an event in which the post-flare loops can re-flare in a short period of $\\sim$16 hr following the first CME/flare eruption. The observed re-flaring at the same location is likely driven by the rapid evolution of the flux rope caused by the magnetic flux emergence and the rotation of the sunspot. This observation provides valuable information on CME/flare models and their prediction. ", "introduction": "The developed systems of loops, named post-flare loops, usually occur in the decay phase of long duration events \\citep[e.g.,][]{bruzek64,kahler77,Harra98}. It is generally believed that magnetic reconnection plays an important role in the formation and evolution of the post-flare loops. The rising of post-flare loops in the corona, as well as the separation of flare ribbons at the footpoints, can be well explained by the classical reconnection model, i.e., the CSHKP model \\citep{carmichael64,sturrock66,hirayama74,kopp76}. In this model, the observed motions are caused by the systematic ascending of the reconnection site in the corona \\citep[e.g.,][]{cheng10a}. The post-flare loops comprise of a system of loops heated and formed consecutively through the ongoing reconnection process. Based on the CSHKP model, the post-flare loops rise with a gradually decreasing speed. However, \\citet{svestka96} found some exceptional cases, in which the post-flare loops rose with a constant speed for a long period of time, and was explained through the combination of two entirely different processes: the initial development of the magnetic reconnection proposed by \\citet{kopp76} and the subsequent expanding motion into the corona. It is generally believed that the post-flare loops evolve without further energy input after their formation. Therefore, they usually fade away into the background after a certain time. To our knowledge, there is no report on the re-activation of a post-flare loop system in previous studies. Coronal mass ejections (CMEs) are another kind of large-scale solar activity that may involve the magnetic field lines that twisted each other, named flux rope, in their core, as revealed by many coronagraph observations \\citep[e.g.,][]{pick06}. The role of flux ropes in the CME eruptions has also been studied extensively through numerical simulation. \\citet{amari00} simulated the evolution of a flux rope and found that it plays a crucial role in the process of the CME eruption. The flux rope was also used by \\citet{fan01} and \\citet{fan03,fan04} to investigate the dynamic evolution of the coronal magnetic field in response to the emergence of twisted magnetic structures. Moreover, \\citet{torok03,torok05} and \\citet{Kliem06} studied the instability of the flux rope and proposed that the CME eruption can be initiated through the kink and/or torus instability. \\citet{cheng10b} suggested that the flux rope may provide a magnetic structure favorable for the eruption and regarded the formation period of the flux rope prior to the eruption as the build-up phase of CMEs \\citep[see also][]{su09,Aulanier10}. \\citet{schrijver08} and \\citet{schrijver09} showed that the emergence of current-carrying flux ropes can lead to the occurrence of a series of major flares within a few hours in the same active region. Further, \\citet{canou09} constructed a flux rope structure above magnetic polarity inversion line (PIL) of an emerging sunspot by a nonlinear force-free field (NLFFF) model using the THEMIS vector magnetogram as the input. Recently, \\citet{guo10} also obtained a flux rope structure through NLFFF modeling and found that the flux rope coincided with the active region filament. However, up to now, there is still no study on the time evolution of extrapolated flux ropes. In this letter, we report the re-flaring of a post-flare loop system occurred on 2005 January 15, and further study the temporal development of associated flux rope. In section 2, we present the observations and data analysis method. Our results are shown in section 3 followed by summary and discussions in section 4. \\begin{figure*}[!ht] \\begin{centering} \\epsscale{0.9} \\plotone{cmeflare.eps} \\caption{(a--b) Running difference images of CMEs 1 and 2 observed by LASCO. (c) EIT 195 {\\AA} image of the CME-associated Flare 1. (d--f) Three images showing the post-flare loops after the peak time of flare 1. Slices 1 and 2 in panel (f) are used to trace the time evolution of the post-flare loop system. (g) EIT 195 {\\AA} image of the CME-associated flare 2. (h) Image showing the post-flare loops of flare 2. \\label{cmeflare}} (Animation of the post-flare loops is available in the online journal) \\end{centering} \\end{figure*} ", "conclusions": "In this study, we report the full evolution of a post-flare loop system, characterized by the initial rising and expanding motion and the final re-activation as another CME. We find that the rising and expanding of the post-flare loops were essentially due to the continual increase of the magnetic flux, which enhanced the magnetic pressure underneath the post-flare loop system. On the other hand, we speculate that the first CME/flare eruption may lead to a weakening of the constraining tension force of the overlying field to the post-flare loop system. We further study the development of the associated flux rope by the NLFFF extrapolation, and find that the twist and length of the flux rope were increased, which may be due to the continual flux emergence and/or the rotation of the sunspot. In fact, magnetic flux emergence \\citep{fan01,fan03,fan04} and sunspots rotation caused by photospheric plasma flows \\citep{amari00,torok03,torok05} are among the most plausible ways that have been invoked in numerical simulations to increase the magnetic non-potentiality leading to the final eruption. Through a quasi-linear force free method \\citep{wang01}, \\citet{zhao08} studied the magnetic topology skeleton of this active region. They found the rising of the magnetic null and its spine, which is consistent with the rising of the extrapolated flux rope and the strong twist of the flux rope as shown here. This strongly twisted flux rope provides a source for the shear of the post-flare loop system that enhances the non-potentiality of associated magnetic configuration. \\citet{jing09} studied the temporal evolution of the magnetic free energy for this active region from 21:00 to 23:00 UT on January 15 and found that it tends to increase prior to the onset of flare 2. The accumulation of the free energy was mainly caused by the emergence of magnetic flux and the rotation of the sunspots, which continually transfers energy below the photosphere into the corona \\citep{kurokawa02,cheng10b}. Therefore, the post-flare configuration may still accumulate magnetic free energy in a short time that is responsible for the second eruption. It is worthy noting that the homologous flares can take place in the same active region with similar morphologies and pattern of development \\citep[e.g.,][]{choe00}, and the successive flares usually occur in the magnetic loops that are closely related but not in exactly the same loops \\citep[e.g.,][]{liu09}. Whereas, the most significant difference, of the events in this studies from the general homologous and successive flares, is that the second CME/flare came directly from the post-flare loops of the first eruption but not the loops structures in the similar or other related locations. In conclusion, it is commonly viewed that a post-flare loop system evolves relatively uneventful after their formation following a CME/flare eruption. Except for some dynamic flows along the post-flare loops \\citep[e.g.,][]{1997SoPh..175..511B,1999SoPh..190..153Q}, the magnetic structure of the loops is closer to potential than that before the eruption, thus unable to re-erupt in a short time after a major eruption. The event studied here provides an exceptional case and may break such a point of view. It shows that a post-flare magnetic configuration can evolve quickly to a state that is favorable for another major CME/flare eruption, as long as there is continual emergence of flux rope and/or rotation of sunspots. Post-flare loops can re-flare in a short time. Models and prediction of solar flares (CMEs) should take into account such an observational fact." }, "1005/1005.5335_arXiv.txt": { "abstract": "{ Dynamos in the Sun and other bodies tend to produce magnetic fields that possess magnetic helicity of opposite sign at large and small scales, respectively. The build-up of magnetic helicity at small scales provides an important saturation mechanism. } { In order to understand the nature of the solar dynamo we need to understand the details of the saturation mechanism in spherical geometry. In particular, we want to understand the effects of magnetic helicity fluxes from turbulence and meridional circulation. } { We consider a model with just radial shear confined to a thin layer (tachocline) at the bottom of the convection zone. The kinetic $\\alpha$ owing to helical turbulence is assumed to be localized in a region above the convection zone. The dynamical quenching formalism is used to describe the build-up of mean magnetic helicity in the model, which results in a magnetic $\\alpha$ effect that feeds back on the kinetic $\\alpha$ effect. In some cases we compare with results obtained using a simple algebraic $\\alpha$ quenching formula. } { In agreement with earlier findings, the magnetic $\\alpha$ effect in the dynamical $\\alpha$ quenching formalism has the opposite sign compared with the kinetic $\\alpha$ effect and leads to a catastrophic decrease of the saturation field strength with increasing magnetic Reynolds numbers. However, at high latitudes this quenching effect can lead to secondary dynamo waves that propagate poleward due to the opposite sign of $\\alpha$. Magnetic helicity fluxes both from turbulent mixing and from meridional circulation alleviate catastrophic quenching. } {} ", "introduction": "The solar dynamo models developed so far and which agree with solar magnetic field observations tend to solve the $\\alpha\\Omega$ mean field dynamo equations. The turbulent $\\alpha$-effect first proposed by Parker (1955) is believed to be generated due to helical turbulence in the convection zone of the Sun. Since $\\alpha$ is generated due to quadratic correlations of the small-scale turbulence we need a closure in order to complete the set of mean field equations, e.g., the first order smoothing approximation (FOSA), and express the mean electromotive force in terms of the mean magnetic fields. This turbulent $\\alpha$ encounters a critical problem when the energy of the mean field becomes comparable to the equipartition energy of the turbulence in the convection zone and hence it becomes increasingly difficult for the helical turbulence to twist rising blobs of magnetic field. The solar dynamo modellers have traditionally used what is referred to as algebraic alpha quenching to mimic this phenomena. This involves replacing $\\alpha$ by $\\alpha_0/(1+\\ob{B}^2/B_{\\rm eq}^2)$, an expression used since Jepps (1975), or by $\\alpha_0/(1+R_{\\rm m} \\ob{B}^2/B_{\\rm eq}^2)$, where $\\alpha_0$ is the unquenched value and $R_{\\rm m}$ is the magnetic Reynolds number, $\\ob{B}$ is the mean magnetic field and $B_{\\rm eq}$ is the equipartition magnetic field. The latter expression has been discussed since the early work of Vainshtein \\& Cattaneo (1992). The $R_{\\rm m}$ in the denominator comes from the fact that the small-scale fluctuating magnetic field reaches equipartition long before the mean magnetic field does. This has been supported by several numerical experiments to determine the saturation behaviour of $\\alpha$ (e.g. Cattaneo \\& Hughes 1996, Ossendrijver et al. 2002). Given the large magnetic Reynolds numbers of Astronomical objects, such phenomena is referred to as catastrophic quenching. After the discovery of the layer of strong radial shear (called the tachocline by Spiegel \\& Zahn 1992) at the bottom of the solar convection zone, Parker (1993) proposed a new class of solar dynamo models called the interface dynamo. In these models the shear is confined to a narrow overshoot layer just beneath the convection zone, also the region of $\\alpha$ effect. The dynamo wave propagates in a direction given by the Parker--Yoshimura rule at the interface between the two layers defined by a steep gradient in the turbulent diffusivity. The toroidal field produced due to stretching by the shear is much stronger than the poloidal field and remains confined in the overshoot layer, away from the region where the $\\alpha$ effect operates. It may be noted that the interface dynamo model may have serious problems when solar-like rotation with positive latitudinal shear is included (Markiel \\& Thomas 1999). Similarly, in the Babcock-Leighton class of flux transport models (Choudhuri et al.\\ 1995; Durney 1995) the toroidal and the poloidal fields are produced in two different layers. Unlike in the interface dynamo models, the coupling between the two layers is mediated both by diffusion and the conveyer belt mechanism of the meridional circulation. It has been proposed that in interface and Babcock-Leighton type dynamos, the $\\alpha$ effect is not catastrophically quenched at high $\\Rm$ because the strength of the toroidal field is very weak in the region of finite turbulent $\\alpha$ (e.g. Tobias, 1996; Charbonneau, 2005). However, according to our knowledge, not much has been done to study the variation of the amplitude of the saturation magnetic field with the magnetic Reynolds number for these classes of $\\alpha\\Omega$ dynamos. Zhang et al (2006) made an attempt to reproduce the surface observations of current helicity in the Sun using a 2D mean field dynamo model in spherical coordinates coupled with the dynamical quenching equation. In a separate paper (Chatterjee, Brandenburg \\& Guerrero, 2010) we have demonstrated that interface dynamo models are also subject to catastrophic quenching. It has been identified a decade ago that the small-scale magnetic helicity generated due to the dynamo action back reacts on the helical turbulence and quenches the dynamo (Blackman \\& Field, 2000; Kleeorin et al. 2000). It has now been shown that this mechanism reduces the saturation amplitude of the magnetic field ($B_{sat}$) with increasing magnetic Reynolds number ($\\Rm$). Nevertheless this constraint may be lifted if the system is able to get rid of small scale helicity through several ways like open boundaries, advective, diffusive and shear driven fluxes (Shukurov et al. 2006, Zhang et al. 2006, Sur et al. 2007, K\\\"apyl\\\"a et al. 2008, Brandenburg et al. 2009, Guerrero et al. 2010). Even though the helicity constraint in direct numerical simulations (DNS) of dynamos with strong shear have been clearly identified, the results can be matched with mean field models having a weaker algebraic quenching than $\\alpha^2$ dynamos (Brandenburg et al.\\ 2001). It is possible to include this process in mean-field dynamo models through an equation describing the evolution of the small scale current helicity. We shall refer to this equation as the dynamical quenching mechanism. In this paper we perform a series of calculations with mean field $\\alpha\\Omega$ models in spherical geometry along with a dynamical equation for the evolution of $\\alpha$ for magnetic Reynolds numbers in the range $1\\le R_{\\rm m}\\le 2\\times10^5$. An important feature of the calculation is that the region of strong narrow shear is separated from the region of helical turbulence. This paper in addition to providing detailed results not mentioned in Chatterjee, Brandenburg \\& Guerrero (2010), is also aimed at studying somewhat more complicated models including meridional circulation. The role of diffusive helicity fluxes modelled into the dynamical quenching equation by using a Fickian diffusion term is also discussed for various models. It may be mentioned that helicity fluxes across an equator can indeed be modelled by such a diffusion term as shown by Mitra et al. (2010). In \\S2 we discuss the features of the $\\alpha\\Omega$ model used, and the formulation of dynamical $\\alpha$ quenching. The results are highlighted in \\S3 and conclusions are drawn in \\S4. ", "conclusions": "We have performed calculations for $\\alpha\\Omega$ dynamos in a spherical shell for spatially segregated $\\alpha$ and $\\Omega$ source regions. The two classes of models we have studied resemble the Parker's interface dynamo and the Babcock-Leighton dynamo. In agreement with earlier work, it is not possible to escape catastrophic quenching by merely separating the regions of shear and $\\alpha$-effect. The saturation value of magnetic energy decreases as $\\sim R_{\\rm m}^{-1}$ for both dynamical quenching and the algebraic quenching with $g_{\\alpha}=\\Rm$ for the simple two layer model without meridional circulation (Fig.~\\ref{fig:satrm}). However we find that a richer dynamical behaviour emerges for the cases with dynamical $\\alpha$ effect, in terms of parity fluctuations and appearance of `secondary' dynamos (Fig.~8, 9). We do not see evidence for chaotic behaviour in the time series of magnetic energy since the dynamo period and the saturation energy remains fairly constant. However this may not be the case in presence of diffusive helicity fluxes which introduce further complexity to the system. Addition of diffusive helicity fluxes relaxes the catastrophic $\\Rm^{-1}$ dependence of the saturation magnetic energy (Fig.~10a, 12). An interesting `side-effect' of diffusive helicity fluxes is the appearance of poleward propagating secondary dynamos. However, because of the lack of scale separation between the mean field and the forcing scale of the helical turbulence we refrain from interpreting this in terms of the poleward migration seen in the Sun. It remains to explore the role of the solar wind, coronal mass ejections which might help in throwing out the small scale helicity from the Sun and thus alleviate catastrophic quenching. The effects of Vishniac \\& Cho fluxes have been investigated and were found to be of secondary importance compared to diffusive helicity fluxes for $\\alpha\\Omega$ mean field dynamos (Guerrero, Chatterjee \\& Brandenburg 2010). When both the meridional circulation and the diffusive helicity fluxes are artificially shut off in the helicity evolution equation, the dynamo fails to reach significant saturation values, as expected (Fig.~12). It is interesting that the Babcock-Leighton dynamos, where $\\alpha$ is concentrated only in a narrow layer at the surface, also produce considerable helicity inside the convection zone when the dynamical quenching (Eq.~\\ref{eq:alphaeq}) is employed (Fig.~13, 14). We have to be cautious about using dynamical quenching equation for dynamo numbers not very large compared to the critical dynamo number. For highly supercritical $\\alpha$, the behaviour of the system begins to be governed by $\\alpha_{\\rm M}$. We would expect that the magnetic field should affect all the turbulent coefficients including both $\\alpha$ and $\\eta$. However for this analysis we have not included an equation for the variation for $\\eta_{\\rm t}$. This is justified for the simple two layer model with a lower $\\eta_{\\rm t}$ in the region of production of strong toroidal fields and a higher $\\eta_{\\rm t}$ in the region of weaker poloidal fields. It may also be noted that by quenching the diffusivity inversely with the magnetic energy in a nonlinear dynamo model, Tobias (1996) was able to produce a bonafide interface model where the magnetic field was restricted to a thin layer at an interface between a layer of shear and cyclonic turbulence. However none of the previous interface models have used the dynamical quenching equation. Unfortunately the direct numerical simulations have not yet reached the modest Reynolds numbers used in this paper ($\\sim 10^4$) which are still much lower than the astrophysical dynamos. To verify if the equation for dynamical quenching works in the same way as in $\\alpha^2$ dynamos, we need to embark upon systematic comparisons between DNS with shear and convection and mean field modelling for $\\alpha\\Omega$ dynamos." }, "1005/1005.0383_arXiv.txt": { "abstract": "{Significant progress has been made in the last years in the understanding of the jet formation mechanism through a combination of numerical simulations and analytical MHD models for outflows characterized by the symmetry of self-similarity. Analytical radially self-similar models successfully describe disk-winds, but need several improvements. In a previous article we introduced models of truncated jets from disks, i.e. evolved in time numerical simulations based on a radially self-similar MHD solution, but including the effects of a finite radius of the jet-emitting disk and thus the outflow.} {These models need now to be compared with available observational data. A direct comparison of the results of combined analytical theoretical models and numerical simulations with observations has not been performed as yet. This is our main goal.} {In order to compare our models with observed jet widths inferred from recent optical images taken with the Hubble Space Telescope (HST) and ground-based adaptive optics (AO) observations, we use a new set of tools to create emission maps in different forbidden lines, from which we determine the jet width as the full-width half-maximum of the emission.} {It is shown that the untruncated analytical disk outflow solution considered here cannot fit the small jet widths inferred by observations of several jets. Furthermore, various truncated disk-wind models are examined, whose extracted jet widths range from higher to lower values compared to the observations. Thus, we can fit the observed range of jet widths by tuning our models.} {We conclude that truncation is necessary to reproduce the observed jet widths and our simulations limit the possible range of truncation radii. We infer that the truncation radius, which is the radius on the disk mid-plane where the jet-emitting disk switches to a standard disk, must be between around 0.1 up to about 1 AU in the observed sample for the considered disk-wind solution. One disk-wind simulation with an {{\\em inner}} truncation radius at about 0.11 AU also shows potential for reproducing the observations, but a parameter study is needed.} ", "introduction": "Astrophysical jets and disks \\citep{Liv09} seem to be inter-related, notably in young stellar objects (YSOs), where jet signatures are well correlated with the infrared excess and accretion rate of the circumstellar disk \\citep{CES90, HEP04}. Disks provide the plasma which is outflowing in the jets, while jets in turn provide the disk with the needed angular momentum removal so that accretion onto the protostellar object takes place \\citep{Har09}. On the theoretical front, the most widely accepted description of this accretion-ejection phenomenon \\citep{Fer07} is based on the interaction of a large scale magnetic field with an accretion disk around the central object. Then, plasma is channeled and magneto-centrifugally accelerated along the open magnetic field lines threading the accretion disk, as first described in \\citet{BlP82}. Several works have extended this study either by semi-analytic models using radially self-similar solutions of the full magnetohydrodynamics (MHD) equations with the disk treated as a boundary condition \\citep{VlT98}, by self-consistently treating the disk-jet system semi-analytically \\citep[e.g.][]{Fer97,CaF00}, or, by self-consistently treating numerically the disk-jet system \\citep[e.g.][]{ZFR07, TFM09}. The original \\citet{BlP82} model, however, has serious limitations for a needed meaningful comparison of its predictions with observations. First, singularities exist at the jet axis, the outflow is not asymptotically super-fast, and most importantly, an intrinsic scale in the disk is lacking with the result that the jet formally extends to radial infinity, to mention just a few. First, the singularity at the axis can be easily taken care of by numerical simulations extending the analytical solutions close to this symmetry axis \\citep[][GVT06 hereafter]{GVT06}. Next, the outflow speed at large distances may be tuned to cross the corresponding limiting characteristic, with the result that the terminal wind solution is causally disconnected from the disk and hence perturbations downstream of the super-fast transition (as modified by self-similarity) cannot affect the whole structure of the steady disk-wind outflow \\citep[][V00 hereafter]{VTS00}, a state which has also been shown to be structurally stable \\citep[][M08 hereafter]{MTV08}. The next step of introducing a scale in the disk has been done in a previous paper \\citep[][paper I hereafter]{STV08}, wherein we presented numerical simulations of truncated flows whose initial conditions are based on analytical self-similar models. \\begin{table*}[!htb] \\caption{List of numerical science models} \\label{tbl_models} \\centering \\begin{tabular}{l l l l l} \\hline\\hline Name & Domain $[ R_0$ x $R_0]$ & Resolution & Description & $R_{\\rm trunc}$ $[ R_0 ]$ \\\\ \\hline model SC1a & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.4$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 5.375 \\\\ model SC1b & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.2$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 5.125 \\\\ model SC1c & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.1$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 4.875 \\\\ model SC1d & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.01$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 3.625 \\\\ model SC1e & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.001$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 2.625 \\\\ model SC2 & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.4$, external analytical solution $\\lambda_1 = 100$, $\\lambda_2 = 0.1$ & 5.375 \\\\ model SC3 & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & same as model SC2, but solutions are swapped & 5.375 \\\\ model SC4 & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.4$, external analytical solution $\\lambda_1 = 1$, $\\lambda_2 = 0.1$ & 5.375 \\\\ model SC5 & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & same as model SC4, but solutions are swapped & 5.375 \\\\ \\hline \\hline model SC1f & [0,50] $\\times$ [6,100] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 0.0005$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 2.375 \\\\ model SC1g & [0,10] $\\times$ [6,20] & 200 $\\times$ 400 & $\\alpha_{\\rm trunc} = 1\\times10^{-5}$, external analytical solution $\\lambda_1 = 10^3$, $\\lambda_2 = 10^{-3}$ & 0.575 \\\\ \\hline \\end{tabular} \\end{table*} In order to test our truncated models, we will now apply our simulations to observations. In recent years, many NIR and optical data have become available exploring the morphology and kinematics of the jet launching region \\citep[e.g.][and references therein]{DCL00,RDB07,Dou08}. Hubble Space Telescope (HST) and adaptive optics (AO) observations give access to the innermost regions of the wind, where the acceleration and collimation occurs \\citep{RMD96,DCL00,WRB02,HEP04}. Because YSO jets emit in a number of atomic (and molecular) lines, we used a set of tools described in \\citet{G??10} to create emission maps in different forbidden lines which were used by other authors to extract the jet width from images. The observed jet widths will be compared with those extracted from our synthetic images. A similar study has been done by \\citet{CFR99}, \\citet{GCF01} and \\citet{DCF04} using a different set of semi-analytical self-similar disk-jet solutions from \\citet{Fer97} and \\citet{CaF00}. Observed jet widths could be reproduced by manually truncating the solutions inside 0.07 AU and outside 1 AU, but the modification of flow streamlines induced by truncation was ignored. The remainder of the paper is organized as follows: we briefly review the initial setup of the numerical simulations in Sect. \\ref{sec_num_models} and describe our procedure for the comparison with observations in Sect. \\ref{sec_obs}. The results of our studies are presented in Sects. \\ref{sec_untrunc} - \\ref{sec_inner}. Finally, we conclude with the implications of the results in terms of the structure of the disk and the respective launching radii of the jets in YSOs. ", "conclusions": "We showed as a proof of concept that jet widths derived from numerical simulations extending analytical MHD jet formation models can be very helpful for understanding recently observed jet widths from observations with adaptive optics and space telescopes. However, further aspects have to be investigated in more detail. An intrinsic feature in all our models with outer truncation is the first bump in the extracted jet width, which complicates the comparison of synthetic and observed jet widths. Only for DG Tau and CW Tau, we could unambiguously find models with outer truncation which fitted the observed jet widths at larger distances. For HN Tau and UZ Tau E, we have no observed jet widths at larger distances, only in the region which is contaminated by the bump. RW Aur shows very small observed jet widths at all scales, which cannot be reproduced by any of our models and runs. The derived truncation radii ($>0.25$ AU) are several times larger than the inner radius of the gaseous disk in T Tauri stars \\citep[0.02-0.07 AU,][]{NCG07}, thus our use of a self-similar disk-wind solution is consistent. The model with inner truncation gives a synthetic jet width, which is constant and is within the range of observed widths. Another advantage of this model is that its flow velocities of the order of 100 km s$^{-1}$ are closer to observed values. All models with outer truncation have very high velocities ($> 600$ km s$^{-1}$ at our scaling point of $R_{\\rm jet} = 10$ AU and $Z_{\\rm jet} = 100$ AU), several times higher than in that with inner truncation. Because we originally focused only on the effect of outer truncation, we kept the inner truncation radius constant in our models SC3 and SC5. In order to further explore the ability of inner truncation to reproduce the observations, we have to vary the inner truncation radius in a parameter study. Naively, one would expect that the derived jet widths are not decreasing for decreasing truncation radii as for the outer truncation, but are {\\em increasing}. This, however, has to be tested with new simulations. In our study, we assumed an inclination of 90$^\\circ$ of the jet, thus projection effects may slightly change our results. We will present the results of such a study in a forthcoming paper. \\citet{GCF01} showed in their Fig. 1 that the measured jet widths are mainly characterized by the ejection index $\\xi$, defined by \\citet{Fer97}. This is related to the model parameter $x$ of the solution of V00, $\\xi = 2\\,(x - 3/4)$, and because in our simulations $x = 0.75$, we get $\\xi = 0$, which is intrinsic for a standard disk. In the solution of \\citet{Fer97}, $\\xi$ also controls the opening of the field lines, because it is connected to the lever arm $\\lambda$ by $\\lambda = 1 + 1 / (2\\,\\xi)$. \\citet{GCF01} favored cold solutions with values of $\\lambda$ between 50--70. If heating along streamlines is allowed, the relation is broken and also warm solutions of e.g. \\citet{CaF00b} with smaller $\\lambda$ values of 8 and opening of streamlines (the ratio of the maximum radius to the initial launch radius) of about 30 can reproduce the observation. In the solution of V00, we have $\\lambda = G ( \\pi / 2 )^{-2} \\approx 39$ and the maximum opening $G_{\\rm max} / G ( \\pi / 2 ) \\approx 893$. The use of other solutions will therefore highly influence our results in terms of the outer truncation radius. \\appendix" }, "1005/1005.2981_arXiv.txt": { "abstract": "{ {The origin and possible universality of the stellar initial mass function (IMF) is a major issue in astrophysics. One of the main objectives of the $Herschel$ Gould belt survey is to clarify the link between the prestellar core mass function (CMF) and the IMF. We present and discuss the core mass function derived from $Herschel$ data for the large population of prestellar cores discovered with SPIRE and PACS in the Aquila Rift cloud complex at $d \\sim260$~pc. We detect a total of 541 starless cores in the entire $\\sim$11~deg$^2$ area of the field imaged at 70--500~$\\mu$m with SPIRE/PACS. Most of these cores appear to be gravitationally bound, and thus prestellar in nature. Our $Herschel$ results confirm that the shape of the prestellar CMF resembles the stellar IMF, with much higher quality statistics than earlier submillimeter continuum ground-based surveys.} ", "introduction": "The question of the origin and possible universality of the IMF, which is crucial for both star formation and galactic evolution, remains a major open problem in astrophysics. The $Herschel$ Space Observatory (Pilbratt et al. 2010), equipped with its two imaging cameras SPIRE (Griffin et al. 2010) and PACS (Poglitsch et al. 2010), provides a unique tool to address this fundamental issue. This problem and other key questions about the earliest phases of star formation and evolution are central to the scientific motivation for the {\\it Herschel} Gould belt survey, which will image the nearby ($d \\leq 0.5$~kpc) molecular cloud complexes of the Gould belt using SPIRE at 250--500~$\\mu$m and PACS at 70--160~$\\mu$m (cf. Andr\\'e \\& Saraceno 2005 and Andr\\'e et al. 2010). Starting with the 1.2~mm continuum study of the Ophiuchus main cloud by Motte et al. (1998), several ground-based (sub)-millimeter dust continuum surveys of nearby, compact cluster-forming clouds such as $\\rho$~Ophiuchi, Serpens, and Orion~B have uncovered 'complete' (but small) samples of prestellar cores whose associated core mass functions (CMF) resemble the stellar IMF (e.g., Motte et al. 1998; Johnstone et al. 2000; Enoch et al. 2006; Nutter \\& Ward-Thompson 2007; -- see also Alves et al. 2007 and Andr\\' e et al. 2007). Albeit limited by small-number statistics at both ends of the CMF, these findings favor theoretical scenarios according to which the bulk of {\\it the IMF of solar-type stars is largely determined by pre-collapse cloud fragmentation}, prior to the protostellar accretion phase (e.g., Hennebelle \\& Chabrier 2008). The problem of the origin of the IMF may thus largely reduce to a good understanding of the processes responsible for the formation and evolution of prestellar cores within molecular clouds. Apart from limited statistics, another limitation of ground-based submillimeter continuum determinations of the CMF is that they rely on uncertain assumptions about the dust properties (temperature and emissivity). $Herschel$ is now making it possible to dramatically improve the quality of the statistics and to reduce the core mass uncertainties by performing direct measurements of the dust temperatures. In this Letter, we use the initial results of the Gould belt survey, obtained toward the Aquila Rift complex as part of the science demonstration phase (SDP) of {\\it Herschel}, to present the first CMF derived from $Herschel$ data. We discuss the robustness of the resulting CMF based on several tests and simulations. The Aquila molecular cloud complex is located at the high Galactic-longitude end of the Aquila Rift, in the neighbourhood of the Serpens star-forming region. Here we adopt a distance of 260~pc for the Aquila star-forming complex. For a detailed introduction of the region and a discussion of its distance, we refer the reader to Bontemps et al. (2010). \\begin{figure*} \\begin{center} \\begin{minipage}{1.0\\linewidth} \\resizebox{0.54\\hsize}{!}{\\includegraphics[angle=270]{./Fig1a_4_00_EpsPngJpgEps.eps}} \\hspace{2mm} \\resizebox{0.44\\hsize}{!}{\\includegraphics[angle=270]{./Fig1b_4_00_EpsPngJpgEps.eps}} \\end{minipage} \\end{center} \\caption{{\\bf(a)} Column density map derived from SPIRE/PACS observations of Aquila. The subregion referred to as the main subfield in the text is marked by the white rectangle. The cyan triangles mark the positions of the 541 starless cores identified in the entire field. The locations of the HII regions W40 and MWC297/Sh62 are shown. The PDR region around W40 is framed by white polygon, while the dashed square outlines the small region shown in more detail in online Fig.~\\ref{Fig_zooms}a. {\\bf(b)} Same as (a) for the main subfield, with a total of 452 starless cores, marked by cyan triangles.} \\label{Fig_coldens}% \\end{figure*} ", "conclusions": "\\subsection{Prestellar nature of the Aquila starless cores} In this paper, we follow the naming convention that a dense core is called {\\it prestellar} if it is starless {\\it and} gravitationally bound (cf. Andr\\'e et al. 2000, Di Francesco et al. 2007). In other words, prestellar cores represent the subset of starless cores that are most likely to form (proto)stars in the future. Strictly speaking, spectroscopic observations would be required to derive virial masses for the cores and determine whether they are gravitationally bound or not. However, millimeter line observations in dense gas tracers such as N$_2$H$^+$ show that thermal motions generally dominate over non-thermal motions in starless cores (e.g., Andr\\'e et al. 2007). Assuming that this is indeed the case for the Aquila cores observed here, we may use the critical Bonnor-Ebert (BE) mass, $M_{{\\rm BE}}^{{\\rm crit}}~\\approx~2.4~R_{{\\rm BE}}~a^2 / G$, as a surrogate for the virial mass, where $R_{{\\rm BE}}$ is the BE radius, $a$ is the isothermal sound speed, and $G$ is the gravitational constant. The critical BE mass may also be expressed as $M_{{\\rm BE}}^{{\\rm crit}} \\approx 1.18 {a^4 \\over G^{3/2}} P_{{\\rm ext}}^{-1/2} $, where $P_{{\\rm ext}}$ is the external pressure, which may be estimated as a function of the column density of the local background cloud, $\\Sigma_{cl}$, as $P_{{\\rm ext}} \\approx 0.88~G~\\Sigma_{cl}^2$ (McKee \\& Tan 2003). For each object, we derived two estimates of the BE mass: (1) $M_{BE}(R_{obs})$ as a function of the observed core radius assuming a gas temperature of 10~K, and (2) $M_{BE}(\\Sigma_{cl})$ as a function of the local background column density measured from both our source-subtracted $Herschel$ images and the near-IR extinction map of Bontemps et al. (2010 -- see also Fig.~\\ref{Fig_coldens_compare}b). We then calculated BE mass ratios of $\\alpha_{\\rm BE} \\equiv \\rm{max}[M_{BE}(R_{obs}), M_{BE}(\\Sigma_{cl})] / M_{obs} $ and selected candidate prestellar cores to be the subset of starless cores for which $\\alpha_{\\rm BE} \\lesssim 2$. Based on this criterion, 314 (or $\\sim69\\% $) of the 452 starless cores of the main subfield and 341 (or $\\sim63\\% $) of the 541 starless cores of the entire field were found to be bound and classified as good candidate prestellar cores. These high fractions of bound objects are consistent with the locations of the Aquila starless cores in a mass versus size diagram (see online Fig.~\\ref{Fig_mass_size} for the main subfield and Fig.~4 of Andr\\'e et al. 2010 for the entire field). The self-gravitating character of most $Herschel$ cores in Aquila is supported further by an independent analysis of their internal column density contrasts. We here define the internal column density contrast of a core as $\\Sigma_{peak}/$$\\bar{\\Sigma}_{core}$, where $\\Sigma_{peak}$ and $\\bar{\\Sigma}_{core}$ are the peak and mean column densities of the core, respectively. Assuming optically thin dust emission at $Herschel$ wavelengths and neglecting any dust temperature/opacity gradient, the internal column density contrast can be estimated from the core intensity values in the same form, as $I_{\\nu}^{\\rm peak}/\\bar{I}_{\\nu}$. The internal column density contrast is closely related to the degree of central concentration of a core defined by Johnstone et al. (2000) as $C =1-(\\bar{\\Sigma}_{core}/\\Sigma_{peak})$. Based on their radial intensity profiles (cf. online Fig.~\\ref{Fig_zooms}b), the Aquila starless cores have a median internal column density contrast $\\sim4$. For comparison, the internal column density contrast is larger than 3.6 for supercritical self-gravitating BE spheres, while it is only 1.5 for a uniform-density, non-self-gravitating object (Johnstone et al. 2000). As a final check, we also estimated the typical column density contrast of the $Herschel$ cores over the local background. For a critically self-gravitating BE core, the mean column density $\\bar{\\Sigma}_{\\rm BE}^{{\\rm crit}}~\\approx~1.56~(P_{{\\rm ext}}~/ G)^{1/2}$ is expected to exceed the column density of the local background cloud $\\Sigma_{cl}$ by a factor of 1.46, if $P_{{\\rm ext}}~\\approx~0.88~G~\\Sigma_{cl}^2$ (McKee \\& Tan 2003). The candidate prestellar cores identified with $Herschel$ have a median column density contrast of $\\sim1.5$ over the local background, which is consistent with the conclusion that they are self-gravitating. \\subsection{Prestellar core mass function in Aquila} Figure~\\ref{Fig_mass_distr}a shows the mass distribution (CMF) of the 452 starless cores identified in the Aquila main subfield. A power-law fit to the high-mass end of this CMF gives d$N$/dlog$M$ $\\propto$ $M^{-1.5 \\pm 0.2}$ for $M_{\\rm core} > 2~M_\\odot$, which is very close to the Salpeter power-law IMF (d$N/$dlog$M$ $\\propto$ $M^{-1.35}$ -- Salpeter 1955). The CMF of the 541 starless cores identified in the entire field, shown by Andr\\'e et al. (2010), has a very similar shape. The robustness of the derived CMF was tested by selecting several subsets of sources, based e.g., on their physical nature or locations in the SPIRE/PACS maps. In particular, we compared the mass spectrum of starless and prestellar cores. Figure~\\ref{Fig_mass_distr}a was then compared with Fig.~\\ref{Fig_mass_distr}b, which shows the differential mass function of 314 bound cores in the main subfield. We obtained the same best-fit power law fit to the high-mass end (d$N$/dlog$M$ $\\propto$ $M^{-1.45 \\pm 0.2}$) as for Fig.~\\ref{Fig_mass_distr}a. Only the low-mass end changed when selecting the subset of candidate prestellar cores. We also selected subsets of starless cores based on their locations in the Aquila field. Along lines of sight to the HII region W40, the associated photon-dominated region (PDR, Shuping et al. 1999) is the source of very strong extended background emission at all infrared wavelengths. The strong background emission in the {\\it Spitzer} 24~$\\mu$m and PACS 70~$\\mu$m images makes it more difficult to discriminate between YSOs and compact starless cores, implying that our census of starless cores is less reliable in this area. We thus constructed another version of the Aquila main subfield CMF, excluding the 83 cores identified toward the PDR region. The excluded region of unusually high infrared background emission (see in Fig.~\\ref{Fig_coldens}a) was defined using the dust temperature map shown in Bontemps et al. (2010). The high-mass end of this CMF can be fitted by a very similar power law to that of Fig.~\\ref{Fig_mass_distr}a: d$N$/dlog$M$ $\\propto$ $M^{-1.5 \\pm 0.3}$. Thanks to the large number of starless cores identified with $Herschel$ in Aquila (541 cores in the entire field), we have been able to consider several core subsamples and construct a statistically meaningful CMF in each case. We confirm that the shape of the prestellar CMF resembles the stellar IMF, using data of far higher quality statistics than earlier submillimeter ground-based surveys and more accurate core masses. Based on simulations, we conclude that our mass distributions are robust and do not depend strongly on different sets of extracted sources. The column density maps shown in Fig.~\\ref{Fig_coldens} and online Fig.~5 illustrate the tight correlation between the spatial distribution of the prestellar cores and the filamentary structure revealed by the $Herschel$ images (Men'shchikov et al. 2010). The difference between our two SDP fields (Aquila Rift and Polaris Flare), in terms of filamentary structure and core mass distribution, is highlighted and discussed in Andr\\'e et al. (2010)." }, "1005/1005.0984_arXiv.txt": { "abstract": "{Stellar rotation is a crucial parameter driving stellar magnetism, activity and mixing of chemical elements. Measuring rotational velocities of young stars can give additional insight in the initial conditions of the star formation process. Furthermore, the evolution of stellar rotation is coupled to the evolution of circumstellar disks. Disk-braking mechanisms are believed to be responsible for rotational deceleration during the accretion phase, and rotational spin-up during the contraction phase after decoupling from the disk for fast rotators arriving at the ZAMS. On the ZAMS, stars get rotationally braked by solar-type winds.} {We investigate the projected rotational velocities $v\\sin i$ of a sample of young stars with respect to the stellar mass and disk evolutionary state to search for possible indications of disk-braking mechanisms. Furthermore, we search for signs of rotational spin-up of stars that have already decoupled from their circumstellar disks.} {We analyse the stellar spectra of 220 nearby (mostly $<100$\\,pc) young (2--600\\,Myr) stars for their $v\\sin i$, stellar age, H$\\alpha$ emission, and accretion rates. The stars have been observed with FEROS at the 2.2m MPG/ESO telescope and HARPS at the 3.6m telescope in La Silla, Chile. The spectra have been cross-correlated with appropriate theoretical templates. We build a new calibration to be able to derive $v\\sin i$ values from the cross-correlated spectra. Stellar ages are estimated from the \\ion{Li}{i} equivalent width at 6708\\,\\AA. The equivalent width and width at 10\\% height of the H$\\alpha$ emission are measured to identify accretors and used to estimate accretion rates $\\dot{M}_\\mathrm{acc}$. The $v\\sin i$ is then analysed with respect to the evolutionary state of the circumstellar disks to search for indications of disk-braking mechanisms in accretors.} {We find that the broad $v\\sin i$ distribution of our targets extends to rotation velocities of up to more than 100\\,kms$^{-1}$ and peaks at a value of $7.8\\pm1.2$\\,kms$^{-1}$, and that $\\sim70\\%$ of our stars show $v\\sin i<30$\\,kms$^{-1}$. Furthermore, we can find indications for disk-braking in accretors and rotational spin-up of stars which are decoupled from their disks. In addition, we show that a number of young stars are suitable for precise radial-velocity measurements for planet-search surveys.} {} ", "introduction": "\\label{sec_intro} Rotation is one of the most important kinematic properties of stars, % giving rise to stellar magnetism and mixing of chemical elements. Stars form from rotating molecular cloud cores, preserving only a very minor fraction of their initial angular momentum (e.g., Palla 2002, Lamm et al. 2005). The initial angular momentum of rotating molecular cloud cores is about 4--5 orders of magnitude higher than that of the stars that eventually form in this cloud core (e.g., Bodenheimer 1989). Stellar formation models must account for that and several mechanisms are discussed, e.g., magnetic braking or magnetocentrifugally driven outflows. However, there is still no consens as on whether there is one dominant process for dispersing angular momentum during the entire star formation process or which process dominates at what evolutionary stage after the formation of a star-disk system (e.g., Palla 2002). The mechanisms believed to be responsible for efficient loss of stellar angular momentum after the formation of a star-disk system involve transfer of angular momentum along magnetic field lines that connect the stellar surface with the disk, either onto the disk (so-called 'disk-locking', e.g. Camenzind 1990, K\\\"onigl 1991) or into stellar winds originating at that boundary (Matt \\& Pudritz 2005 and 2008). Both models predict that the stellar magnetic field threads the circumstellar disk, accretion of disk material onto the star occurs along the field lines, and magnetic torques transfer angular momentum away from the star. In the disk-locking model, the angular velocity of the star is then locked to the Keplerian velocity at the disk boundary. For reasonable magnetic field strengths and accretion rates $\\sim$\\,10$^{-9}$\\,M$_{\\odot}$/yr, both models account for the observed rotational periods of classical T~Tauri stars (e.g., Armitage \\& Clarke 1996). Both models can be observationally distinguished only by the absence or presence of stellar winds originating from the boundary. When the accretion process ends and there is no longer an efficient way to disperse angular momentum, the star can spin-up during the contraction phase (see, e.g., Lamm et al. 2005, Bouwman et al. 2006). Once arrived on the ZAMS, stars undergo additional rotational braking by magnetic winds, irrespective of whether they arrived as slow or fast rotators (e.g., Skumanich 1972, Palla 2002, Wolff et al. 2004). If this picture holds true, very young stars are expected to rotate slower than slightly older stars that have already decoupled from their disks. Three different classes of young stars are expected to be distinguishable, which can be interpreted as a kind of evolutionary sequence: (i) slow rotating stars which still accrete, (ii) slow to intermediate rotating stars which do not longer accrete and are about to gradually spin-up due to contraction and (iii) fast rotators without disks. However, one has to keep in mind, that all star forming regions show an enormous spread in rotation periods and that differences in rotation velocities between different star forming regions can also be caused by the initial conditions of star formation and the braking time-scale of the star-disk system (e.g., Stassun et al. 1999, Lamm et al. 2005, Herbst et al. 2007, Nguyen et al. 2009). Several surveys of stellar rotation periods and projected rotational velocities of very young stars indeed found relatively slow rotators (e.g., Herbst et al. 2002; Rebull et al. 2004, 2006; Sicilia-Aguilar et al. 2005), which points to effective braking mechanisms, whereas other surveys did not find evidence for rotational braking of very young stars (e.g., Makidon et al. 2004; Nguyen et al. 2009). Nevertheless, rotational velocity measurements of stars with circumstellar disks and ongoing accretion in associations such as $\\eta$\\,Cha, TW~Hydrae, and NGC\\,2264 support a disk-locking mechanism for the removal of angular momentum from the star (Lamm et al. 2005; Bouwman et al. 2006; Jayawardhana et al. 2006; Fallscheer \\& Herbst 2006). This rotational braking due to disk-locking is mainly visible in very young clusters with an age $<$\\,3\\,Myr. These stars are expected to spin-up by a factor of $\\approx$\\,3 due to contraction after being magnetically disconnected from the circumstellar disk. In fact, a large fraction of PMS stars have been observed that evolve at nearly constant angular velocity through the first 4\\,Myr (Rebull et al. 2004). In clusters at the age of $\\sim$\\,10\\,Myr, signs of stellar velocity spin-up can be seen (Rebull et al. 2006). For our radial-velocity survey of very young stars, SERAM (\\emph{Search for Exoplanets with Radial-velocity At the MPIA}, Setiawan et al. 2008), we started to measure the $v\\sin i$ of stars with ages of 2--600\\,Myr and selected slow rotators for high-precision radial velocity measurements to search for planets. These $v\\sin i$ measurements are a necessary prerequisite for such an RV survey, because the projected rotational velocity of a star limits the accuracy of the radial velocity measurement due to broadening of the spectral lines. Furthermore, rotational velocities play a crucial role in describing and understanding chromospheric and coronal activity, which is related to radial velocity jitter due to stellar spots and scales with rotational velocities (e.g., Pallavicini et al. 1981, Noyes et al. 1984, Saar \\& Donahue 1997). Thus, measuring rotational velocities and periods of young stars is very important to understand stellar formation, evolution and activity. This paper is structured as follows. In section \\ref{obs}, we describe the target selection, observations, and data reduction. The parameter estimation is described in Section~\\ref{sec_dataana} and the calibration of $v\\sin i$ is described in Appendix~\\ref{sec_vm}. In Section~\\ref{res}, we present the results of our analysis and search for indications of disk-braking and rotational spin-up. Finally, we summarize our results in Section \\ref{con}. ", "conclusions": "\\label{con} We analysed 229 young and nearby stars for their projected rotational velocity $v\\sin i$, stellar age, and accretion signatures. The stars were part of our initial RV planet search target list in the SERAM project and have been observed with FEROS at the 2.2m MPG/ESO telescope and HARPS at the 3.6m telescope in La Silla, Chile. For stars showing broad H$\\alpha$ emission, we checked other accretion indicators, like the EW of \\ion{He}{i} at 5876\\,\\AA, to identify those stars that still accrete material from a circumstellar disk. For these stars, the veiling has also been measured. The age of the stars has been derived from the EW of \\ion{Li}{i} at 6708\\,\\AA{}. We calculated $v\\sin i$ of our 220 target stars and 9 spectroscopic companions by measuring the width of the CCF of the stellar spectra. To do this, we cross-correlated the stellar spectra with theoretical templates of similar spectral type. The CCF has been fitted with a Gaussian and the conversion to $v\\sin i$ was done with our new calibration (Equation~\\ref{eq_sigma0}) for both FEROS and HARPS spectra. The main conclusions from this work can be summarized as follows: \\begin{itemize} \\item{The mean stellar age of our sample is 30\\,Myr with a spread (width of Gaussian fit) of 20\\,Myr. 43 stars are younger than 10\\,Myr.} \\item{The distribution of $v\\sin i$ for all target stars peaks at 7.8$\\pm$1.2\\,kms$^{-1}$, with the vast majority having projected rotational velocities $<30$\\,kms$^{-1}$.} \\item{We found indication for rotational braking due to disk-locking, because the accreting stars in our sample rotate significantly slower ($\\langle v\\sin i\\rangle =10\\pm1$\\,kms$^{-1}$) than the non-accreting stars with more evolved disks ($\\langle v\\sin i\\rangle =15\\pm2$\\,kms$^{-1}$). The only 2 fast rotating ($v\\sin i \\sim 38$\\,kms$^{-1}$) accretors in our sample are DI\\,Cha and CR\\,Cha. This might point to differences in time spent for disk-braking or to different initial conditions in different star formation regions, but the significance of this statement is hampered by low number statistics.} \\item{This $v\\sin i$ evolution also provides indication that stars undergo rotational spin-up after being decoupled from their circumstellar disk and evolving along the Pre-Main-Sequence track. For more evolved stars with debris or no disks and an age\\,$>30$\\,Myr, $\\langle v\\sin i\\rangle$ decreases again to 10\\,$\\pm$\\,1\\,kms$^{-1}$. We conclude that these stars have been rotationally braked on the ZAMS, e.g., by magnetic winds. } \\item{We estimated the maximum rotational period of our target stars. The period distribution of the whole sample peaks at 2.5\\,$\\pm$\\,0.3\\,days, a similar result as found for young stars by Cieza \\& Biliber (2007).} \\item{Due to disk-braking, young and accreting stars can rotate sufficiently slow to serve as suitable targets for RV planet searches, since the accuracy of radial velocity measurements depends on the stellar $v\\sin i$. When accretion ends after $\\sim$5--10\\,Myr and the stars decouple from their disk, they tend to spin up as they contract, such that they are usually no longer suitable for precise RV measurements. This changes only when the stars arrive on the ZAMS and are slowed down again by wind braking at ages of $\\sim$30\\,Myr.} \\end{itemize}" }, "1005/1005.3569_arXiv.txt": { "abstract": "\\noindent On 1998 November 14, Saturn and its rings occulted the star GSC 0622-00345. We observed atmospheric immersion with NSFCAM at the National Aeronautics and Space Administration's Infrared Telescope Facility on Mauna Kea, Hawaii. Immersion occurred at 55{\\decdegree}5 S planetocentric latitude. A 2.3 {\\micron}, CH\\sb{4}-band filter suppressed reflected sunlight. Atmospheric emersion and ring data were not successfully obtained. We describe our observation, light-curve production, and timing techniques, including improvements in aperture positioning, removal of telluric scintillation effects, and timing. Many of these techniques are known within the occultation community, but have not been described in the reviewed literature. We present a light curve whose signal-to-noise ratio per scale height is 267, among the best ground-based signals yet achieved, despite a disadvantage of up to 8 mag in the stellar flux compared to prior work. \\if\\submitms y \\else \\comment{\\hfill\\herenote{DRAFT of {\\today} \\now}.} \\fi ", "introduction": "\\label{intro} Stellar occultations by planets offer among the highest spatial resolution of any astronomical observing technique, competing even with atmospheric descent probes for certain measurements. From Earth, spatial resolution is limited by the projected stellar beam size and Fresnel diffraction, the latter giving \\sim1 km resolution for solar system giant planets for visible and near-IR wavelengths. This is sufficient to reveal fine structure in a planetary ring or propagating gravity (buoyancy) waves in an atmosphere. However, the occultation technique is demanding, and failing to optimize even a single parameter of an observation can significantly reduce the quality of the derived light curve. Occultation observers have thus developed numerous specialized techniques to improve their data. Many of these have never been described in the literature. On 1998 November 14, Saturn and its rings occulted GSC 0622-00345, as predicted by \\citet{BoshMcDonald1992ajsatocs}. We obtained a light curve for atmospheric immersion, based on infrared imaging observations at the NASA Infrared Telescope Facility (IRTF) on Mauna Kea, HI. Figure \\ref{geometry} shows our viewing circumstances and Table \\ref{occtab} gives parameters of the event. Occultation cadences are typically from a few to 10 Hz, and in the case of lunar occultations they can be much higher. These rates are sufficient for the temporal resolution of flux variations caused by scintillation (variable refractive focusing and defocusing of the stellar beam) in Earth's atmosphere. This is typically the limiting noise source for occultations by bright stars. During our observations, we monitored Rhea, which experienced approximately the same scintillation effects as the star, just arcseconds away. We were thus able to use measurements of Rhea's (assumed constant) flux to compensate for scintillation. This, in combination with other techniques, produced a light curve of very high quality. Here we present a full accounting of the methods we employed to observe the event and derive the light curve, including those previously undocumented. Section \\ref{obslcsec} describes observing techniques, data reduction methods, and the light curve. It also discusses an attempt to apply optimal photometry to light curve production and thoughts on improving instrument interfaces to avoid failures such as our lost egress observation. Section \\ref{timingsec} presents a method for independently verifying the system timing solution. Section \\ref{conclsec} offers brief conclusions. We include the light curve and software for analyzing occultation timing data as electronic supplements. Paper II \\citep{HarringtonEtal2010apjSatoc98II} uses the derived light curve to investigate the atmosphere of Saturn. ", "conclusions": "\\label{conclsec} We have presented a light curve based on IRTF observations of the 1998 November 14 occultation of GSC 0622-00345 by Saturn. The light curve has a per-frame S/N of 66 and a per-scale-height S/N of 267, placing it among the best ground-based atmospheric light curves. We have described the observational and analytic methods used to derive the light curve, in some cases for the first time in the literature, although these techniques are known among experienced occultation observers. Methods that improved the photometry include using Rhea as a scintillation and pointing standard, spatially subsampling the data by a factor of 10 for accurate aperture placement, and removing a template of scattered light. Separate, identical, measurements of a flashing LED placed in the telescope beam provided an independent measure of the frame rate, when fit by a detailed model of the chip's sensitivity throughout the imaging cycle. Optimal photometry shows promise at reducing the photometric noise level as the star dims, but requires a PSF standard that is substantially brighter than the occultation star. Occultation cameras with wider fields of view (but still able to operate with accurate timing at high frame rates) would facilitate finding bright PSF standards." }, "1005/1005.1877_arXiv.txt": { "abstract": "{We present the observations of the starburst galaxy \\object{M82} taken with the \\textit{Herschel} SPIRE Fourier Transform Spectrometer. The spectrum (194--671 $\\mu$m) shows a prominent CO rotational ladder from $J$ = 4--3 to 13--12 emitted by the central region of \\object{M82}. The fundamental properties of the gas are well constrained by the high $J$ lines observed for the first time. Radiative transfer modeling of these high-S/N $^{12}$CO and $^{13}$CO lines strongly indicates a very warm molecular gas component at $\\sim 500$~K and pressure of $\\sim 3 \\times 10^{6}$~K~cm$^{-3}$, in good agreement with the H$_{2}$ rotational lines measurements from \\textit{Spitzer} and \\textit{ISO}. We suggest that this warm gas is heated by dissipation of turbulence in the interstellar medium (ISM) rather than X-rays or UV flux from the straburst. This paper illustrates the promise of the SPIRE FTS for the study of the ISM of nearby galaxies.} ", "introduction": "Starburst galaxies provide us with the opportunity to study star formation and its effect on the interstellar medium (ISM) in extreme environments. These galaxies combine large central gas concentrations and high ionizing radiation fields, resulting in bright molecular, neutral and ionized gas emission lines. At a distance of 3.9 Mpc \\citep{sakai99}, \\object{M82} is the most well-studied starburst galaxy in the local universe, and it is widely used as a starburst prototype in cosmological studies. Its infrared luminosity \\citep[$5.6\\times 10^{10}$ L$_\\odot$,][]{sanders03} corresponds to a star-formation rate of 9.8 M$_\\odot$ yr$^{-1}$, which has almost certainly been enhanced by its interaction with \\object{M81} and \\object{NGC 3077} \\citep{yun93}. With a reported molecular gas content of $1.3 \\times 10^9$ M$_\\odot$ \\citep{walter02}, its bright emission lines of CO and other molecules allow us to study its ISM in great detail \\citep{shen95,walter02,ward03}. Far-infrared fine structure lines were used to constrain the physical properties of the ionized gas and photo-dissociation regions (PDRs) in \\object{M82}. \\citet{colbert99} found that the ionized gas emission can be reproduced with a 3--5 Myr old instantaneous starburst and a gas density of 250 cm$^{-3}$, while the PDR component is best fit with a density of 2\\,000 cm$^{-3}$, in pressure equilibrium with the ionized phase. Stellar evolution and photoionization models \\citep{forster03} indicate a series of a few, Myr-duration starbursts with a peak of activity 10 Myr ago in the central regions, and 5 Myr ago in the circumnuclear ring. Models of the PDR and molecular emission as a set of non-interacting hot bubbles driving spherical shells of swept-up gas into a surrounding uniform medium also predict a starburst age of 5--10 Myr, but fail to match the observed far-infrared luminosity \\citep{yao09}. The strengths of the CO lines place fundamental constraints on the physical properties of the molecular gas. \\cite{tilanus91} fitted \\coone\\ and \\cotwo\\ lines from the central starburst up to $J$ = 3--2 with a single-component model with temperatures of 30--55 K and densities of 3--7 $\\times 10^3$ cm$^{-3}$. \\citet{wild92} used lines up to the CO $J$ = 6--5 transition to refine these parameters to 40--50 K and $\\sim 10^4$ cm$^{-3}$, while HCN and HCO$^+$ lines suggested densities greater than $3\\times 10^5$ cm$^{-3}$ are present. \\citet{petitpas00} showed evidence for a temperature or density gradient across the starburst region. \\citet{weiss05} showed that CO emission up to $J$ = 3--2 is dominated by more extended regions while higher $J$ transitions originate in the central disk. In this paper, we present observations of \\object{M82} with \\textit{Herschel} \\citep{pilbratt10} using the SPIRE Fourier Transform Spectrometer (FTS) \\citep{griffin10}, which measures the complete far-infrared spectrum from 194 to 671 $\\mu$m. This spectral region is particularly interesting for probing the peak of the CO spectral line energy distribution (SLED) in gas-rich galaxies. The wealth of lines across a continuous spectral region allows for unprecedented precision in modeling the physical and chemical properties of the molecular ISM. Here, we focus on the measurement and analysis of the CO rotational transitions from the central starburst in \\object{M82}. ", "conclusions": "We have presented the \\textit{Herschel}-SPIRE spectroscopic observations of the starburst galaxy \\object{M82}. The spectra show a prominent CO emission-line ladder along with \\ion{C}{I} and \\ion{N}{II} lines. Radiative transfer modeling of CO lines clearly indicates a warm gas component at $\\sim$500~K in addition to the cold ($\\sim$30~K) component derived from ground-based studies. The properties of the warm gas are strongly constrained by the high $J$ lines, observed here for the first time. The temperature and mass of warm gas agree with the H$_2$ rotational lines observations from \\textit{Spitzer} and \\textit{ISO}. At this temperature H$_{2}$ is the dominant coolant instead of CO, and we argue that turbulence from stellar winds and supernovae may be the dominant heating mechanism." }, "1005/1005.1931_arXiv.txt": { "abstract": "According to the no-hair theorem, all astrophysical black holes are fully described by their masses and spins. This theorem can be tested observationally by measuring (at least) three different multipole moments of the spacetimes of black holes. In this paper, we analyze images of black holes within a framework that allows us to calculate observables in the electromagnetic spectrum as a function of the mass, spin, and, independently, the quadrupole moment of a black hole. We show that a deviation of the quadrupole moment from the expected Kerr value leads to images of black holes that are either prolate or oblate depending on the sign and magnitude of the deviation. In addition, there is a ring-like structure around the black-hole shadow with a diameter of $\\sim$10 black-hole masses that is substantially brighter than the image of the underlying accretion flow and that is independent of the astrophysical details of accretion flow models. We show that the shape of this ring depends directly on the mass, spin, and quadrupole moment of the black hole and can be used for an independent measurement of all three parameters. In particular, we demonstrate that this ring is highly circular for a Kerr black hole with a spin $a\\lesssim0.9M$, independent of the observer's inclination, but becomes elliptical and asymmetric if the no-hair theorem is violated. Near-future very-long baseline interferometric observations of Sgr A* will image this ring and may allow for an observational test of the no-hair theorem. ", "introduction": "According to the no-hair theorem, the external spacetimes of black holes are uniquely characterized by their masses and spins (Israel 1967, 1968; Carter 1971, 1973; Hawking 1972; Robinson 1975). This theorem requires that the cosmic censorship conjecture (Penrose 1969) holds and that the exterior spacetime is free of closed time-like curves. Under these assumptions, all astrophysical black holes should be fully described by the Kerr metric (Kerr 1963). It is widely accepted that the universe contains an abundance of black holes as inferred from the observations of the centers of nearby galaxies (e.g., Tremaine et al.\\ 2002), of our own galactic center (Ghez et al. 2008; Gillessen et al. 2009), and of many galactic binaries (e.g., McClintock \\& Remillard 2006). Nonetheless, the factual existence of an event horizon is yet unproven and has only been inferred indirectly (e.g., Narayan, Garcia, \\& McClintock 1997, 2001; Narayan \\& Heyl 2002; McClintock, Narayan, \\& Rybicki 2004; see also Psaltis 2006). Alternative explanations for the nature of these objects have been suggested, which include naked singularities (Manko \\& Novikov 1992), exotic stellar objects (Friedberg, Lee, \\& Pang 1987; Mazur \\& Mottola 2001; Barcel\\'{o} et al. 2008), as well as a breakdown of general relativity itself on horizon scales (e.g., Yunes \\& Pretorius 2009; c.f. Psaltis et al. 2008). A test of the no-hair theorem can both identify the observed dark compact objects with Kerr black holes and verify the validity of general relativity in the strong-field regime. Indeed, within general relativity, if a compact object is not a Kerr black hole, then its external spacetime will not satisfy the no-hair theorem. Alternatively, if general relativity is not valid in the strong-field regime, the external spacetime of a compact object that is surrounded by a horizon may violate the no-hair theorem (see, however, Psaltis et al.\\ 2008). Mass and spin are the first two multipole moments of a black-hole spacetime. If the no-hair theorem is correct, then all higher multipole moments only depend on the mass and spin, and any deviation from the Kerr moments has to be zero. Consequently, the no-hair theorem can be tested by measuring (at least) three multipole moments of such a spacetime (Ryan 1995). In part I of this series of papers (Johannsen \\& Psaltis 2010; hereafter Paper I), we investigated a framework for testing the no-hair theorem with observations of compact objects in the electromagnetic spectrum. Based on a quasi-Kerr spacetime that contains an independent quadrupole moment (Glampedakis \\& Babak 2006), we analyzed in detail the spacetime properties that are critical for such observations as a function of the mass, spin, and quadrupole moment. We showed that already very moderate changes of the quadrupole moment lead to significant alterations of various quantities that determine observables. In particular, we explored the effect of changing the quadrupole moment on the locations of the innermost stable circular orbit (ISCO) and of the circular photon orbit, as well as on the lensing and redshift of photons. There has been, already, substantial work on potential tests of the no-hair theorem with observations of gravitational waves from extreme mass-ratio inspirals (Ryan 1995, 1997a, 1997b; Barack \\& Cutler 2004, 2007; Collins \\& Hughes 2004; Glampedakis \\& Babak 2006; Gair et al.\\ 2008; Li \\& Lovelace 2008; Apostolatos et al.\\ 2009; Vigeland \\& Hughes 2010). In this series of papers, we show that observations of black holes in the electromagnetic spectrum may also allow for a clean test of the no-hair theorem. In particular, we identify different observables that probe the quadrupole moments of the spacetimes but depend very weakly on the usual astrophysical complications, such as the flow geometry, the mode of emission, and the variability of the accretion flows that generate the photons we detect from black holes. Imaging observations of accreting black holes at (sub$-$)mm wavelengths using very-long baseline interferometry (VLBI) promise to enable unprecedented views of the vicinities of black-hole horizons. Recent VLBI-observations along only three baselines resolved Sgr~A$^*$, the black hole in the center of the Milky Way, on a scale comparable to its event horizon and provided evidence for sub-horizon scale structures (Doeleman et al.\\ 2008) as well as for the presence of an event horizon (Broderick, Loeb, \\& Narayan 2009). Far greater resolution can be achieved by adding either existing or planned telescopes located at various places on the Earth (Fish \\& Doeleman 2009). The black hole in the center of M87 (Broderick \\& Loeb 2009) as well as a small number of other nearby supermassive black holes (Psaltis 2008) offer additional targets for horizon-scale imaging that can be utilized in the near future. Imaging observations are expected to be able to resolve the shadows of black holes and lead to the determination of their spins and inclinations (Falcke et al. 2000; Broderick \\& Loeb 2005, 2006; Fish \\& Doeleman 2009). Templates for images of accretion flows around Kerr black holes within general relativity that are suitable to these observations have been reported by a number of authors (Bardeen 1973; Speith et al. 1995; Fanton et al. 1997; Falcke et al. 2000; Takahashi 2004; Beckwith \\& Done 2004, 2005; Dexter \\& Agol 2009; Broderick \\& Loeb 2005, 2006; Yuan et al. 2009). Recently, Bambi \\& Freese (2009) explored the possibility of using black-hole images to test whether black holes violate the Kerr bound $a\\le M$. In this paper, we study the properties of the images of compact objects that violate the no-hair theorem using the quasi-Kerr formalism we developed in Paper I. We calculate numerically the mapping between locations in the vicinity of a black hole and positions in the observer's sky using the mass, spin, and quadrupole moment of the spacetime as independent parameters. We investigate the impact of varying the quadrupole moment on the properties of this mapping and show that the images of the accretion flows around compact objects that violate the no-hair theorem are expected to have prolate or oblate geometries. Measuring the spacetime moments from the images of an accretion flow will be, of course, very model dependent and limited by our lack of understanding of the intrinsic geometry of the flow itself. For example, prolate images of the inner accretion may be the result of resolving the formation region of a jet and not of a violation of the no-hair theorem (see, e.g., Broderick \\& Loeb 2009). Moreover, a measurement of the spin from an image of the shadow alone is difficult (e.g., Falcke et al. 2000; Takahashi 2004) and might require complementary observations such as a multiwavelength study of polarization (Broderick \\& Loeb 2006; see also Schnittman \\& Krolik 2009, 2010). Additionally, accretion flows are very turbulent and variable at timescales much shorter than the rotation period of the Earth, which sets the characteristic integration time for an interferometric imaging observation. If this time variability is produced by a highly coherent, orbiting inhomogeneity in the accretion flow, it may allow measuring the properties of the compact object via non-imaging techniques (Doeleman et al.\\ 2009). The variability, however, will limit and may prohibit altogether the ability of obtaining a clean image of the accretion flow. The images of optically thin accretion flows around black holes, however, reveal a characteristic bright ring at the projected radius of the circular photon orbit along null geodesics (Beckwith \\& Done 2005) with properties that remain constant even as the underlying accretion flows are highly variable. This bright ring is the result of the light rays that orbit around the black hole many times before they reach the distant observer, and, therefore, have a much larger path length through the optically thin accretion flow. These photons can make a significant contribution to the total disk emission and produce higher-order images (Cunningham 1976; Laor, Netzer, \\& Piran 1990; Viergutz 1993; Bao, Hadrava, \\& {\\O}stgaard 1994; ${\\rm \\check{C}ade\\check{z}}$, Fanton, \\& Calvani 1998; Agol \\& Krolik 2000; Beckwith \\& Done 2005). We use our formalism to show that the bright emission ring is circular for a Schwarzschild black hole and remains nearly circular for Kerr black holes. On the other hand, if the quadrupole moment is left as an independent parameter, the ring shape changes significantly and becomes asymmetric. The degree of asymmetry is a direct measure of the violation of the no-hair theorem. We show that the diameter of the ring depends only very weakly on the spin and quadrupole moment of the black hole and can be used to directly measure the mass of the object. In addition, the ring is displaced off center in the image plane in the case of rotating black holes (Beckwith \\& Done 2005; see, also, Takahashi 2004), and we show that the displacement is a direct measure of the object's spin, modulo the disk inclination. In Section~2 we briefly review the framework for testing the no-hair theorem with observations in the electromagnetic spectrum. We simulate images of black holes and of the photon rings in Sections~3 and 4, respectively, and show how they depend on the mass, spin, and quadrupole moment of a given black hole. In Section~5 we quantify these dependencies before we discuss our results and their implications for future observations in Section~6. ", "conclusions": "In Paper~I, we proposed a new framework for testing the no-hair theorem with observations of black holes in the electromagnetic spectrum. We formulated our tests based on a quasi-Kerr metric (Glampedakis \\& Babak 2006), which deviates smoothly from the Kerr metric in the quadrupole moment. Since the no-hair theorem admits exactly two independent multipole moments for a black hole, a measurement of these three moments will allow us to test the no-hair theorem. In this paper, we calculated numerically the mapping between different locations in the accretion flow around a quasi-Kerr black hole and positions in the image plane of a distant observer. Our calculations allowed us to study the potential of using imaging observations of black holes, that will become available in the near future, in order to test the no-hair theorem. We argued that the expected image of an accretion flow will be characterized by a bright emission ring generated by light rays that circle multiple times around the event horizon before emerging towards the observer. We identified the ring diameter as a direct measure of the mass of the black hole, and we quantified the dependence of the displacement and the asymmetry of the ring on the spin and the quadrupolar parameter as well as on the disk inclination. For a given inclination angle, a measurement of the displacement and the asymmetry directly measures the spin and the quadrupolar parameter of the system, respectively. The asymmetry itself provides a direct measure of the violation of the no-hair theorem. It is important to emphasize here that only the relative displacement and asymmetry of the ring (i.e., measured in units of the ring diameter) and not their absolute values are necessary in inferring the spin and quadrupole moment of the black-hole spacetime. As a result, the outcome of such an observation does not depend on the distance to the black hole, which might not be known accurately. On the other hand, the angular diameter of the photon ring is proportional to the mass of the black hole. This can lead to an accurate measurement of the mass of the black hole if the distance is known, or else of the distance to the black hole, if its mass is known from, e.g., dynamical observations. Sgr A*, the black hole in the center of the Milky Way, is the ideal candidate for a test of the no-hair theorem due to its high brightness, large angular size, and relatively unimpeded observational accessibility. Recent VLBI observations (Doeleman et al. 2008) resolved Sgr A* on horizon scales. Incorporating additional baselines to the VLBI network will lead to the first images of Sgr~A$^{*}$ within the next few years (Fish \\& Doeleman 2009). The emission from Sgr~A$^{*}$ at sub-mm wavelengths is optically thin, the size of the scattering ellipse is $\\lesssim1M$, and the resolution of a VLBI image at this wavelength is also comparable to $\\simeq1M$ (Doeleman et al.\\ 2008; Fish \\& Doeleman 2009). The smearing of the images due to scattering and the finite resolution of the array will, therefore, not preclude measuring the position and asymmetry of the photon ring to an accuracy that is adequate in providing a quantitative test of the no-hair theorem. Observations of the orbits of stars in the vicinity of the black hole have provided an independent measurement of its mass (Ghez et al.\\ 2008; Gillessen et al.\\ 2009). Perhaps more importantly, the same observations may lead in the future to an independent measurement of the spin and orientation of the black hole, as well as to a complementary test of the no-hair theorem (Will 2008; Merritt et al.\\ 2010)." }, "1005/1005.3105_arXiv.txt": { "abstract": "Roche-lobe overflow and common envelope evolution are very important in binary evolution, which is believed to be the main evolutionary channel to hot subdwarf stars. The details of these processes are difficult to model, but adiabatic expansion provides an excellent approximation to the structure of a donor star undergoing dynamical timescale mass transfer. We can use this model to study the responses of stars of various masses and evolutionary stages as potential donor stars, with the urgent goal of obtaining more accurate stability criteria for dynamical mass transfer in binary population synthesis studies. As examples, we describe here several models with the initial masses equal to $1~M_\\odot$ and $10~M_\\odot$, and identify potential limitations to the use of our results for giant branch stars. ", "introduction": "\\label{sec1} The Fourth Meeting on Hot Subdwarf Stars and Related Objects was convened in Shanghai, China from 19th to 24th, July. Recent discoveries and developments in both theory and observation of hot subdwarfs and related objects were reported, and many unsolved problems were discussed. As we all know, hot subdwarf stars are extreme horizontal-branch stars or related objects. They may dominate the UV-upturn of early-type galaxies and they exist in both the field of our Galaxy and its globular cluster system. About half of subdwarf B (sdB) stars are binaries \\citep{max01,saf01}; binary evolution is obviously important in their formation, as the observed systems are too compact to have avoided past mass transfer. Binary population synthesis models \\citep{han02,han03,han08} explain naturally the sdB binary fractions and the UV-upturn of early-type galaxies via: (a) one or two phases of common envelope (CE) evolution, (b) stable Roche-lobe overflow (RLOF), and (c) the merger of two He-core white dwarf stars (WDs). RLOF and CE evolution are also very important in the formation of other binary systems, e.g., cataclysmic variables, X-ray binaries, double white dwarfs and binary neutron stars. Those systems containing compact objects are among the most energetic and rapidly variable sources known. Unfortunately, we know very little about the details of CE evolution, but dynamically unstable RLOF appears to be the trigger that launches it. The threshold conditions for dynamical mass transfer now in common use in binary population synthesis calculations are based on polytropic models for rapid mass loss process \\citep[and references therein]{hjel87}. These polytropic studies provide useful qualitative insights into RLOF and CE evolution, but they omit much relevant physics, and fail to address many advanced evolutionary stages of interest. Detailed studies of binary evolution \\citep{han02,chen08} reveal a need for more realistic determinations of the threshold conditions for dynamical mass transfer. With the motivation and cautions above, we set out to study stellar rapid mass loss based on the pioneering work of \\citet{hjel89a,hjel89b}. We describe the basic assumptions and numerical techniques employed in modeling stellar adiabatic mass loss in section~\\ref{sec2}, with initial results and their possible application to binary population synthesis in section~\\ref{sec3}. In section~\\ref{sec4}, we offer a short discussion of some remaining problems. ", "conclusions": "\\label{sec4} RLOF process and CE evolution are very important in the formation of hot subdwarf stars and other binary systems. Earlier published studies of \\citet[and references therein]{hjel87} provided useful qualitative insights in RLOF and CE evolution, but left considerable room for improvement, especially with regard to the range of evolved phases that need to be addressed. These deficiencies were in part redressed in later work \\citep{hjel89b}, but only a fragment of that work was ever published \\citep{hjel89a}. The stellar adiabatic mass loss models described in this paper considerably extend the scope of Hjellming's work. They allow us not only to study the interior structure of donor stars undergoing dynamical timescale mass transfer, but also to evaluate the stability criteria for dynamical mass transfer in binary population synthesis. An initial application of our models to $1~M_\\odot$ and $10~M_\\odot$ stars has been described in this paper. The reader should beware that the these models have limitations to their usefulness, as they depend on being able to separate the process of dynamical relaxation to hydrostatic equilibrium from that of thermal relaxation to thermal equilibrium. On the main sequence, for example, the dynamical time scale of the Sun is roughly $10^{-10}$ that of its thermal time scale, making the adiabatic mass loss model an excellent description of the asymptotic behavior of a solar-type star to very rapid mass loss. However, at the opposite extreme in the HR diagram, stars near the tip of the RGB or AGB have thermal time scales approaching their dynamical time scales, and the entire mass transfer process may require full-blown time-dependent modeling. This convergence of thermal and dynamical time scales is broadly related to the abrupt expansion seen in our models of $1~M_\\odot$ and $10~M_\\odot$ stars at the tip of the giant branch (TGB), when their outermost superadiabatic layers are stripped away. In reality, Roche lobe overflow is far from spherically-symmetric, as treated here (by necessity), but that superadiabatic expansion may nevertheless reflect a real physical phenomenon. Convection itself is not spherically symmetric, and with convective velocities in the superadiabatic zone approaching sound speed, it may be possible for rising flows near the inner Lagrangian point to bridge the potential to the companion's Roche lobe while the donor still lies well within its own Roche limit. We are not now able to pass judgement on that possibility. When dynamical instability occurs, common envelope evolution almost certainly follows, as the thermal time scale of the accreting star is invariably longer than that of the donor, which itself is generally much longer than its dynamical time scale. This ordering of time scales ensures that the envelope cannot cool efficiently on the transfer time scale, but remains extended and engulfs both stellar cores. It may happen that thermal time scale mass loss from the donor is still rapid enough to form a common envelope, but this case can only arise if the system is stable against dynamical mass transfer, while unstable to thermal time scale mass transfer. Since the stellar dynamical time scales of both donor and accretor are shorter than their thermal time scales, the prospect arises of a quasistatic common envelope, that is, of formation of a contact binary. Such large numbers of such objects are known (the W UMa systems, as examples) that they must be very long-lived, evolving in a very different fashion from CE evolution as we have used that term above. We believe that stellar adiabatic mass loss models provide the most useful approach to date toward defining the limits of dynamical stability in interacting binaries, essential input to the construction of binary population synthesis models." }, "1005/1005.3619_arXiv.txt": { "abstract": "We have derived Fe abundances of 16 solar-type Pleiades dwarfs by means of an equivalent width analysis of \\ion{Fe}{1} and \\ion{Fe}{2} lines in high-resolution spectra obtained with the Hobby - Eberly Telescope and High Resolution Spectrograph. Abundances derived from \\ion{Fe}{2} lines are larger than those derived from \\ion{Fe}{1} lines (herein referred to as over-ionization) for stars with \\teff\\ $< 5400$ K, and the discrepancy (\\dfe\\ = [\\ion{Fe}{2}/H] - [\\ion{Fe}{1}/H]) increases dramatically with decreasing \\teff, reaching over 0.8 dex for the coolest stars of our sample. The Pleiades joins the open clusters M\\,34, the Hyades, IC\\,2602, and IC\\,2391, and the Ursa Major moving group, demonstrating ostensible over-ionization trends. The Pleiades \\dfe\\ abundances are correlated with \\ion{Ca}{2} infrared triplet and H$\\alpha$ chromospheric emission indicators and relative differences therein. Oxygen abundances of our Pleiades sample derived from the high-excitation \\ion{O}{1} triplet have been previously shown to increase with decreasing \\teff, and a comparison with the \\dfe\\ abundances suggests that the over-excitation (larger abundances derived from high excitation lines relative to low excitation lines) and over-ionization effects that have been observed in cool open cluster and disk field main sequence (MS) dwarfs share a common origin. Curiously, a correlation between the Pleiades \\ion{O}{1} abundances and chromospheric emission indicators does not exist. Star-to-star \\ion{Fe}{1} abundances have low internal scatter ($< 0.11$ dex), but the abundances of stars with \\teff $< 5400$ K are systematically higher compared to the warmer stars. The cool star [\\ion{Fe}{1}/H] abundances cannot be connected directly to over-excitation effects, but similarities with the \\dfe\\ and \\ion{O}{1} triplet trends suggest the abundances are dubious. Using the [\\ion{Fe}{1}/H] abundances of five stars with \\teff\\ $> 5400$ K, we derive a mean Pleiades cluster metallicity of [Fe/H] $= +0.01 \\pm 0.02$. ", "introduction": "Studies of Galactic chemical evolution are dependent on accurately derived abundances of stars spanning all ages, populations, kinematics, masses, and metallicities. Stars with masses $M \\leq 1 \\mathrm{M}_{\\odot}$ are especially important given their dominance of the initial mass function \\citep[IMF; e.g.,][]{2002Sci...295...82K}. Abundance studies utilizing high-resolution spectroscopy and local thermodynamic equilibrium (LTE) analyses of near-solar metallicity G and K dwarfs in open clusters and in the disk field, however, have revealed that the observed abundances of at least some elements derived for these cool main sequence (MS) dwarfs may be spurious. In particular, studies have found evidence of over-ionization and over-excitation\\footnotemark[6], i.e., larger abundances are derived from lines of singly ionized species compared to neutral species and from high-excitation lines of neutral species compared to low-excitation lines, respectively. The first indication that some abundances derived for cool MS stars are problematic may have come from \\citet{1974ApJS...27..405O}, who found the over-ionization of Sc, Ti, Cr, and Fe for a sample of 10 K dwarfs ($4800 \\leq$ \\teff $\\leq 5600$ K) in the solar neighborhood. After a careful analysis of the procedures and stellar parameters used in the abundance derivations, the author was unable to account for the overabundances of the ionized species. \\citet{1998A&AS..129..237F} found similar over-ionization results for Sc, V, Cr, Fe, and Y in five field K dwarfs ($4510 \\leq$ \\teff\\ $\\leq 4833$ K). The authors could not exclude an inaccurate temperature scale that is several hundred K too low as a possible source of the anomalous abundances, but in the end, they suggest that non-LTE (NLTE) effects are the more likely cause. \\footnotetext[6]{In this paper, we use over-ionization and over-excitation to refer to the observed enhanced abundances derived from spectral lines of singly ionized species or from high excitation lines, as opposed to other common usages referring specifically to the non-LTE (NLTE) effects of over-ionization (the mean intensity, $J_{\\nu}$, is larger than the Planck function, $B_{\\nu}$, in lower atomic energy levels), resonance scattering, and photon pumping \\citep[e.g.,][]{1998A&AS..129..237F,2005ARA&A..43..481A}.} Open clusters have been important to the identification and continued study of the over-excitation/ionization effects, because the presumed internal chemical homogeneity of the clusters provides a baseline with which anomalous abundances can be compared. \\citet{2000ApJ...533..944K} derived O abundances from the high-excitation ($\\chi = 9.15$ eV) near-IR \\ion{O}{1} triplet of a K dwarf in each of the the Pleiades and NGC 2264 open clusters, and in both cases, the abundances were highly enhanced: [O/H] $= +0.85$ and $+0.43$, respectively. Such high O abundances are not expected for clusters with nominal metallicities of [Fe/H] $= 0.00$ \\citep{1988ApJ...327..389B} and -0.15 \\citep{2000ApJ...533..944K}, respectively. Following \\citet{2000ApJ...533..944K}, \\citet{2003AJ....125.2085S} derived the abundances of MS dwarfs in the M\\,34 cluster; over-ionization of Fe and over-excitation of Si (the abundances of which were derived from lines with excitation potentials in the range $5.61 \\leq \\chi \\leq 6.19$ eV) are seen in the coolest stars of the sample. Oxygen abundances of the cool M\\,34 dwarfs, as well as cool Pleiades dwarfs, derived from the high-excitation \\ion{O}{1} triplet are also highly enhanced \\citep{2004ApJ...602L.117S}, confirming the earlier results of \\citet{2000ApJ...533..944K}. Subsequent to these early open cluster studies, the over-ionization of Fe has been confirmed in the Hyades \\citep{2004ApJ...603..697Y,2006AJ....131.1057S} and Ursa Major (UMa) moving group \\citep{2005PASP..117..911K}, and of Ti in the young pre-MS clusters IC\\,2602 and IC\\,2391 \\citep{2009A&A...501..553D}. Overabundances of O derived from the \\ion{O}{1} triplet have been reported for the UMa moving group \\citep{2005PASP..117..911K}, the Hyades \\citep{2006ApJ...636..432S}, and IC\\,4665 \\citep{2007ApJ...660..712S}. Over-excitation effects have been reported for other elements, as well, including S in the Pleiades \\citep {2004ApJ...602L.117S}, Si, Ti, Ni, and Cr in IC\\,4665 \\citep{2005ApJ...635..608S}, Ni in the Hyades \\citep{2006AJ....131.1057S}, and Ca, Ti, and Na in IC\\,2602 and IC\\,2391 \\citep{2009A&A...501..553D}. Recent abundance analyses of cool field stars have also identified over-excitation/ionization effects \\citep{2004A&A...420..183A,2007A&A...465..271R,2008AJ....135..618C}, confirming the findings of earlier work. The over-excitation/ionization abundance anomalies are not thought to represent real photospheric overabundances; rather, we believe that they are a signal that our knowledge of cool dwarf atmospheres and/or spectral line formation therein is incomplete. As of yet, the source or cause of the effects has not been identified. Systematically erroneous stellar parameters, e.g., an inaccurate \\teff\\ scale, could lead to the observed abundance trends, but in general, unrealistically large parameter errors would have to be present \\citep[e.g.,][]{2005PASP..117..911K,2006ApJ...636..432S}. Furthermore, parameter changes made in response to the overabundances of one element often increase those of another \\citep[e.g.,][]{2003AJ....125.2085S}. NLTE effects have been suggested as the cause \\citep[e.g.,][]{1998A&AS..129..237F}, but in general, the over-excitation/ionization effects seen in cool dwarfs are in stark contrast to extant NLTE calculations. For instance, LTE analyses of the high-excitation \\ion{O}{1} triplet in the spectra of MS dwarfs are predicted to result in increasingly discrepant abundances with increasing \\teff\\ for stars with \\teff\\ $> 6000$ K, requiring negative NLTE corrections up to 0.4 -- 0.5 dex at 6500 K for solar metallicity dwarfs \\citep{2003A&A...402..343T, 2009A&A...500.1221F}. Below 6000 K, the NLTE corrections are predicted to be $< 0.1$ dex and essentially zero below 5500 K \\citep{2003A&A...402..343T}. Chromospheric emission and photospheric activity (spots, plages, and faculae) have also been suggested sources for the abundance anomalies \\citep[e.g.,][]{2006ApJ...636..432S}. These inhomogeneities could produce apparent over-excitation/ionization effects within a strict LTE framework. Continuing our efforts to delineate and understand the observed over-excitation/ionization effects in cool MS dwarfs, we have derived \\ion{Fe}{1} and \\ion{Fe}{2} abundances of 16 Pleiades dwarfs, 15 of which have had O abundances derived from the high-excitation \\ion{O}{1} triplet \\citep{2004ApJ...602L.117S}. The \\ion{O}{1} abundances evince a steep increase, reaching [O/H] $\\approx 1.0$ dex near 5000 K, and star-to-star dispersion below 5500 K. We use the newly derived Pleiades Fe abundances to investigate if the over-excitation and over-ionization effects observed in cool MS dwarfs are related and indeed manifestations of the same phenomenon. Future observational studies that could place stringent constraints on these effects and bring us closer to discovering the source of the anomalous abundances are also discussed. ", "conclusions": "We have derived Fe abundances via an EW analysis of \\ion{Fe}{1} and \\ion{Fe}{2} lines in high-resolution and moderate-S/N spectra of 16 MS dwarfs in the Pleiades open cluster. The \\feii\\ abundances increase dramatically relative to \\fei\\ at \\teff\\ below 5400 K, with the difference reaching over 0.8 dex in the coolest stars. This behavior is akin to what is seen in M\\,34, the Hyades, and the UMa moving group. Comparison of the \\dfe\\ abundance patterns in the Pleiades and Hyades, as well as the \\otrip\\ abundances in the Pleiades, Hyades, and UMa moving group, suggests that the trends may relax with age, though metallicity may yet prove to be a factor. Abundances of cool dwarfs in additional open clusters or other stellar associations, especially those older than the Hyades, are needed to determine if either age or metallicity are related to these anomalous abundances. The \\fei\\ abundances are also higher in Pleiads below 5400 K, but they show no evidence of an increase with decreasing \\teff. The inability to attribute the high \\fei\\ abundances of the cool stars to the over-excitation effects illustrates the difficulty of quantifying this phenomenon. With lines of exceptionally high excitation potential such as the \\ion{O}{1} triplet, the over-excitation effect is clearly seen \\citep[e.g.,][]{2006ApJ...636..432S,2007ApJ...660..712S}, but for lines with excitation potentials $\\lesssim 5$ eV, the effect is more difficult to pinpoint. Our \\ion{Fe}{1} linelist includes transitions ranging in excitation potential from 2.18 to 5.10 eV, but for the individual Pleiades stars, no increase in the line-by-line abundances as a function of excitation potential, like that seen in \\ion{Ni}{1} abundances of cool Hyades dwarfs \\citep[1.83 eV $\\leq \\chi \\leq$ 4.42 eV;][]{2006AJ....131.1057S}, is evident (Figure 4). Line-to-line sensitivities to the over-excitation/ionization effects have yet to be clearly delineated, and it needs to be determined if there is an excitation potential threshold above which the abundances derived from these lines become enhanced by these effects. Similarly, it needs to be determined if there is an excitation potential threshold {\\it below} which the opposite occurs, the abundances derived from the low-excitation lines are lower due to these effects. Such behavior would be expected if the overabundances of high-excitation (singly ionized) lines are due to the overpopulation of high-excitation (singly ionized) electronic states at the expense of depopulating low-excitation states. Whether or not the over-excitation effects impact the spectroscopic derivation of stellar parameters (\\teff, $\\log g$, and $\\xi$), an approach not adopted here, also needs to be determined. Future investigations of these effects will require high-quality high-resolution spectroscopy so that accurate line-by-line abundances can be derived, even from features of just a few m{\\AA} in strength. A strong correlation between the \\dfe\\ and \\otrip\\ abundances of the Pleiades dwarfs is evident in Figure 2, suggesting that the over-excitation/ionization effects share a common cause or origin. Chromospheric emission and photospheric spots have been shown to be promising culprits, but to this point, the data are inconclusive. Whereas strong correlations between the Pleiades \\dfe\\ and chromospheric emission indicators and their residuals exist, they do not exist between \\otrip\\ and chromospheric emission. These contradictory results complicate the interpretation of the observed over-excitation/ionization effects and will have to be addressed by future studies. Also, comparing abundances and chromospheric emission indicators measured using different spectra may not provide an accurate test of a true correlation because of potential temporal changes in chromospheric activity levels. Future investigations into the over-excitation/ionization effects in cool open cluster dwarfs should make every effort to derive chromospheric activity levels from the same spectra so that any possible relation between the two can be more definitively delineated. Determining the influence of photospheric spots on abundance derivations is more arduous. Multicomponent model atmospheres- simulating photospheres with different areal coverages of hot, cool, and quiescent spots- have been shown to able to reproduce the measured EWs of the \\ion{O}{1} triplet in a sample of Hyades stars, but, while such exercises are useful and demonstrate the plausibility of the photospheric spot hypothesis, the results are only suggestive. Observationally, a simultaneous photometric and spectroscopic monitoring program could be used to identify any correlated changes in spot coverage and spectral line strengths. Such observational constraints would be helpful to determine if spotted photospheres affect high-resolution abundance derivations. Despite the challenges, the possible connection between spots, over-excitation/ionization effects, and pre-MS Li depletion should provide sufficient motivation for future efforts. A final conclusion that can be drawn from this study is that those carrying out spectroscopic abundance analyses of open clusters should heed caution when their samples include cool dwarfs, particularly those with \\teff\\ $< 5400$ K. Including the abundances of these stars may skew cluster mean abundances. A similar caution may be needed for those studying cool MS dwarfs in the disk field, as well. Further investigations into the sensitivity of the over-excitation/ionization effects to excitation potential, first ionization potential, stellar age, and stellar metallicity are needed in order to identify the extent and ubiquity of these effects." }, "1005/1005.0040_arXiv.txt": { "abstract": "We discuss the properties of orbits within the influence sphere of a supermassive black hole (BH), in the case that the surrounding star cluster is nonaxisymmetric. There are four major orbit families; one of these, the pyramid orbits, have the interesting property that they can approach arbitrarily closely to the BH. We derive the orbit-averaged equations of motion and show that in the limit of weak triaxiality, the pyramid orbits are integrable: the motion consists of a two-dimensional libration of the major axis of the orbit about the short axis of the triaxial figure, with eccentricity varying as a function of the two orientation angles, and reaching unity at the corners. Because pyramid orbits occupy the lowest angular momentum regions of phase space, they compete with collisional loss cone repopulation and with resonant relaxation in supplying matter to BHs. General relativistic advance of the periapse dominates the precession for sufficiently eccentric orbits, and we show that relativity imposes an upper limit to the eccentricity: roughly the value at which the relativistic precession time is equal to the time for torques to change the angular momentum. We argue that this upper limit to the eccentricity should apply also to evolution driven by resonant relaxation, with potentially important consequences for the rate of extreme-mass-ratio inspirals in low-luminosity galaxies. In giant galaxies, we show that capture of stars on pyramid orbits can dominate the feeding of BHs, at least until such a time as the pyramid orbits are depleted; however this time can be of order a Hubble time. ", "introduction": "Following the demonstration that self-consistent equilibria could be constructed for triaxial galaxy models \\citep{Schwarzschild1979,Schwarzschild1982}, observational evidence gradually accumulated for non-axisymmetry on large (kiloparsec) scales in early-type galaxies \\citep{Franx1991,Statler2004,Cappellari2007}. On smaller scales, imaging of the centers of galaxies also revealed a wealth of features in the stellar distribution that are not consistent with axisymmetry, including bars, bars-within-bars, and nuclear spirals \\citep{Shaw1993,ErwinSparke2002,Seth2008}. In the nuclei of low-luminosity galaxies, the non-axisymmetric features may be recent or recurring, associated with ongoing star formation; in luminous elliptical galaxies, central relaxation times are so long that triaxiality, once present, could persist for the age of the universe. In a triaxial nucleus, torques from the stellar potential can induce gradual changes in the eccentricities of stellar orbits, allowing stars to find their way into the central BH. Gravitational two-body scattering also drives stars into the central BH, but only on a time scale of order the central relaxation time, which can be very long, particularly in the most luminous galaxies. Simple arguments suggest that the feeding of stars to the central BHs in many galaxies is likely to be dominated by large-scale torques rather than by two-body relaxation \\citep[e.g.][]{MerrittPoon2004}. This paper discusses the character of orbits near a supermassive BH in a triaxial nucleus. The emphasis is on low-angular-momentum, or ``centrophilic,'' orbits, the orbits that come closest to the BH. Self-consistent modelling \\citep{PoonMerritt2004} reveals that a large fraction of the orbits in triaxial BH nuclei can be centrophilic. Within the BH influence sphere, orbits are nearly Keplerian, and the force from the distributed mass can be treated as a small perturbation which causes the orbital elements (inclination, eccentricity) to change gradually with time. A standard way to deal with such motion is via orbit averaging \\cite[e.g.][]{SandersVerhulst1985}, i.e., averaging the equations of motion over the short time scale associated with the unperturbed Keplerian motion. The result is a set of equations describing the slow evolution of the remaining orbital elements due to the perturbing forces. This approach was followed by \\cite{SridharTouma1999} for motion in an axially-symmetric nucleus containing a massive BH, and by \\cite{SambhusSridhar2000} for motion in a constant-density triaxial nucleus. In their discussion of motion in triaxial nuclei, \\citet{SambhusSridhar2000} passed over one important class of orbit: the centrophilic orbits, i.e., orbits that pass arbitrarily close to the BH. Examples of centrophiliic orbits include the two-dimensional ``lens'' orbits \\citep{SridharTouma1997,SridharTouma1999} and the three-dimensional ``pyramids'' \\citep{MerrittValluri1999, PoonMerritt2001}. Centrophilic orbits are expected to dominate the supply of stars and stellar remnants to a supermassive BH \\citep[e.g.][]{MerrittPoon2004} and are the focus of the current paper. The paper is organized as follows. In \\S\\ref{sec_potential} we present a model for the gravitational potential of a triaxial nuclear star cluster, which is more general than that studied in \\cite{SambhusSridhar2000}, but which has many of the same dynamical features. Then in \\S\\ref{sec_orbitavg} we write down the orbit-averaged equations of motion, and in \\S\\ref{sec_analysis} present a detailed analytical study of their solutions, with emphasis on the case where the triaxiality is weak and the eccentricity large. In this limiting case, the averaged equations of motion turn out to be fully integrable. In \\S\\ref{sec_capture} we derive the equations that describe the rate of capture of stars on pyramid orbits by the BH. Comparison of orbit-averaged treatment with real-space motion is made in \\S\\ref{sec_realspace}, to test the applicability of the former. In \\S\\ref{sec_GR} we consider the effect of general relativity on the motion, which imposes an effective upper limit on the eccentricty. \\S\\ref{sec_RR} discusses the connection with resonant relaxation: we argue that a similar upper limit to the eccentricity should characterize orbital evolution in the case of resonant relaxation. Finally, in \\S\\ref{sec_estimates} we make some quantitative estimates of the importance of pyramid orbits for capture of stars in galactic nuclei. \\S\\ref{sec_summary} sums up. ", "conclusions": "\\label{sec_summary} We discussed the character of orbits within the radius of influence $\\rh$ of a supermassive BH at the center of a triaxial star cluster. The motion can be described as a perturbation of Keplerian motion; we derive the orbit-averaged equations and explore their solutions both analytically (when the triaxiality is small) and numerically. Orbits are found to be mainly regular in this region. There exist three families of tube orbits; a fourth orbital family, the pyramids, can be described as eccentric Keplerian ellipses that librate in two directions about the short axis of the triaxial figure. At the ``corners'' of the pyramid, the angular momentum reaches zero, which means that stars on these orbits can be captured by the BH. We derive expressions for the rate at which stars on pyramid orbits would be lost to the BH; there are many similarities with the more standard case of diffusional loss cone refilling, but also some important differences, due to the fact that the approach to the loss cone is deterministic for the pyramids, rather than statistical. The inclusion of general relativistic precession is shown to impose a lower bound on the angular momentum. We argue that a similar lower bound should apply to orbital evolution in the case that the torques are due to resonant relaxation. The rate of consumption of stars from pyramid orbits is likely to be substantially greater than the rate due to two-body relaxation in the most luminous galaxies, although in the absence of mechanisms for orbital repopulation, these high consumption rates would only be maintained until such a time as the pyramid orbits have been drained; however the latter time can be measured in billions of years." }, "1005/1005.2860_arXiv.txt": { "abstract": "We report the detections of two substellar companions orbiting around evolved intermediate-mass stars from precise Doppler measurements at Subaru Telescope and Okayama Astrophysical Observatory. HD 145457 is a K0 giant with a mass of 1.9 $M_{\\odot}$ and has a planet of minimum mass $m_2\\sin i=2.9 M_{\\rm J}$ orbiting with period of $P=176$ d and eccentricity of $e=0.11$. HD 180314 is also a K0 giant with 2.6 $M_{\\odot}$ and hosts a substellar companion of $m_2\\sin i=22 M_{\\rm J}$, which falls in brown-dwarf mass regime, in an orbit with $P=396$ d and $e=0.26$. HD 145457 b is one of the innermost planets and HD 180314 b is the seventh candidate of brown-dwarf-mass companion found around intermediate-mass evolved stars. ", "introduction": "Over 400 exoplanets have been discovered by various techniques during the past 15 years\\footnote {See, e.g., table at http://exoplanet.eu/}. Among the techniques, precise Doppler technique has been the most powerful method for planet detection around various types of stars including solar-type stars, evolved giants and subgiants, early-type stars, and so on (e.g. Udry \\& Santos 2007 and references therein). Recently, precise Doppler measurements in infrared wavelength region started to explore planets around very low-mass stars down to 0.1 $M_{\\odot}$ (Bean et al. 2010). These surveys will help us to understand properties of planets as a function of stellar mass, age, evolutionary stage, and so on, and thus provide us more general picture of planet formation and evolution. In particular, planets around giants and subgiants have been extensively surveyed over the past several years mainly from the viewpoint of planet searches around intermediate-mass (1.5--5$M_{\\odot}$) stars (e.g. Frink et al. 2002; Setiawan et al. 2005; Sato et al. 2008b; Hatzes et al. 2005, 2006; Johnson et al. 2010; Lovis \\& Mayor 2007; Niedzielski et al. 2009b; D$\\ddot{\\rm{o}}$llinger et al. 2009; de Medeiros et al. 2009; Liu et al. 2009; Omiya et al. 2009). Intermediate-mass dwarfs, namely BA-type dwarfs, are more difficult for Doppler planet searches because of paucity of spectral features and their rotational broadening. It is thus difficult to achieve high measurement precision in radial velocity (but see eg. Galland et al. 2005, 2006). On the other hand, those in evolved stages, namely GK-type giants and subgiants, have many sharp absorption lines in their spectra suitable for precise radial velocity measurements, which makes them promising targets for Doppler planet searches. Actually, more than 30 substellar companions to such evolved stars have been already found and they have shown remarkable properties: more than twice higher occurrence rate of giant planets for intermediate-mass subgiants than that for lower-mass stars (Johnson et al. 2007b; Bowler et al. 2010), larger typical mass of giant planets (Lovis et al. 2007; Omiya et al. 2009), lack of inner planets with semimajor axes $<$0.6 AU (Johnson et al. 2007a; Sato et al. 2008a; Niedzielski et al. 2009a), and lack of metal-rich tendency in host stars (Pasquini et al. 2007; Takeda et al. 2008). All these statistical properties should be confirmed by collecting a larger number of samples. Since 2001, we have been carrying out a Doppler planet search program targeting 300 GK giants at Okayama Astrophysical Observatory (OAO) and have discovered 9 planets and 1 brown dwarf so far from the program (Sato et al. 2003, 2007, 2008ab; Liu et al. 2008). In order to further extend the survey, we have established an international consortium between Chinese, Korean, and Japanese researchers using 2m class telescopes in the three countries (East-Asian Planet Search Network; Izumiura 2005), which recently announced discoveries of a planet (Liu et al. 2009) and a brown dwarf (Omiya et al. 2009) around GK giants. A total of about 600 GK giants are now under survey by the consortium. In this paper, we report the detections of two new substellar companions around intermediate-mass giants (HD 145457 and HD 180314) from our newly started planet search program using Subaru 8.2m telescope and the above 2m class telescopes. The substellar companions presented here were uncovered by the initial screening with Subaru and followed up with OAO 1.88m telescope. We describe the outline of the Subaru survey in Section 2 and the observations in Section 3. The stellar properties are presented in Section 4, and radial velocities, orbital solutions, and results of line shape analysis are provided in Section 5. Section 6 is devoted to summary and discussion. ", "conclusions": "We here reported the two new substellar companions to K0 III giants from Subaru and OAO planet search programs. The discoveries add to the recent growing population of substellar companions around evolved intermediate-mass stars. HD 145457 b ($m_2\\sin i=2.9M_{\\rm J}$, $a=0.76$ AU) is one of the innermost planets found around giants. All the planets currently known around evolved intermediate-mass stars orbit at a$\\ge$0.6 AU (Johnson et al. 2007a; Sato et al. 2008a; Niedzielski et al. 2009a). Two possible causes of the lack of inner planets has been proposed: they are originally deficient or engulfed by the central stars. In the case of intermediate-mass subgiants ($<2M_{\\odot}$), the former scenario is considered to be appropriate because they are obviously less evolved and have stellar radii of $\\lesssim 6R_{\\odot}$, which means that they could only have engulfed very short-period planets like hot-Jupiters (e.g. Johnson et al. 2007a). It is not easy, however, to discriminate between the two scenarios in the case of intermediate-mass giants (typically $\\ge2M_{\\odot}$) because it is difficult to know accurate evolutionary status of giants. The inner planets could be tidally engulfed by the central stars at the phase of the tip of red giant branch (RGB). Thus if we observe giants that have already passed through the tip of RGB, namely core-helium burning stars, we could not find any inner planets around the stars even if the planets originally existed (Sato et al. 2008a; Villaver \\& Livio 2009). However, it is apparently difficult to distinguish such giants from those ascending RGB for the first time because they locate at nearly the same region on the HR diagram (red clump region). Based on stellar evolutionary models (cf.evolutionary tracks by Girardi et al. 2000), core-helium burning giants stay at the clump region for $\\sim$100 times longer than first ascending RGB stars do. Therefore, if we can collect hundreds of planets around such giants, it will become possible to verify the causes of the lack of inner planets statistically. HD 180314 b has a minimum mass of 22 $M_{\\rm J}$ and orbits around the central star with 2.6$M_{\\odot}$. This is the 7 th brown-dwarf-mass (13--80$M_{\\rm J}$) companion to evolved intermediate-mass stars (Hatzes et al. 2005; Lovis \\& Mayor 2007; Liu et al. 2008; Omiya et al. 2009; Niedzielski et al. 2009b\\footnote{BD+20 2457 has two brown-dwarf-mass companions in the system.}). Actually all of their host stars are estimated to be more massive than 2.5 $M_{\\odot}$, suggesting that more massive stars tend to have more massive companions (Lovis \\& Mayor 2007; Omiya et al. 2009). Several scenarios have been proposed for the formation of brown-dwarf-mass companions including gravitational collapse in protostellar clouds like stellar binary systems (Bonnell \\& Bastien 1992; Bate 2000) and gravitational instability in protostellar disks (Boss 2000; Rice et al. 2003). Even by core-accretion scenario in protoplanetary disks, super-massive companions with $\\gtsim10M_{\\rm J}$ could be formed on a certain truncation condition for gas accretion (Ida \\& Lin 2004; Alibert et al. 2004; Mordasini et al. 2007). If the companions form like stellar binary systems, they are expected to have a wide variety of orbital eccentricity. The above 7 brown dwarf candidates, however, have relatively low eccentricies of 0--0.3, which may be favored by the scenarios that they formed in circumstellar disks and have not experienced significant gravitational interaction with other companions. One more thing to note here is that metallicity of the host stars of the above brown dwarf candidates are ranging from $\\rm {[Fe/H]}=-1$ to 0.2. Although the number of samples is still small, this suggests that the formation mechanism is independent of or less sensitive to metallicity in this range. \\\\ This research is based on data collected at Subaru Telescope and Okayama Astrophysical Observatory (OAO), which are operated by National Astronomical Observatory of Japan (NAOJ). We are grateful to all the staff members of Subaru and OAO for their support during the observations. We thank the National Institute of Information and Communications Technology for their support on high-speed network connection for data transfer and analysis. BS is supported by MEXT's program \"Promotion of Environmental Improvement for Independence of Young Researchers\" under the Special Coordination Funds for Promoting Science and Technology, and by Grant-in-Aid for Young Scientists (B) No.20740101 from the Japan Society for the Promotion of Science (JSPS). YJL is supported by the National Natural Science Foundation of China under grant number 10803010 and 10821061. BCL acknowledge the Astrophysical Research Center for the Structure and Evolution of the Cosmos (ARCSEC, Sejong University) of the Korea Science and Engineering Foundation (KOSEF) through the Science Research Center (SRC) program. This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France." }, "1005/1005.0923_arXiv.txt": { "abstract": "The circumgalactic medium (CGM) around galaxies is believed to record various forms of galaxy feedback and contain a significant portion of the ``missing baryons'' of individual dark matter halos. However, clear observational evidence for the existence of the hot CGM is still absent. We use intervening galaxies along 12 background AGNs as tracers to search for X-ray absorption lines produced in the corresponding CGM. Stacking \\chandra grating observations with respect to galaxy groups and different luminosities of these intervening galaxies, we obtain spectra with signal-to-noise ratios of 46-72 per 20-m\\AA\\ spectral bin at the expected \\ovii\\ K$\\alpha$ line. We find no detectable absorption lines of \\cvi, \\nvii, \\ovii, \\oviii, or \\neix. The high spectral quality allows us to tightly constrain upper limits to the corresponding ionic column densities (in particular $\\log[N{\\rm_{OVII}(cm^{-2})}]\\le$14.2--14.8). These nondetections are inconsistent with the Local Group hypothesis of the X-ray absorption lines at $z\\simeq0$ commonly observed in the spectra of AGNs. These results indicate that the putative CGM in the temperature range of $10^{5.5}-10^{6.3}$ K may not be able to account for the missing baryons unless the metallicity is less than 10\\% solar. ", "introduction": "\\label{sec:intro} Modern simulations are converging on the formation and evolution of the dark galactic halos as well as the large-scale intergalactic structure of the Universe (e.g., \\citealt{nav96, she01, cen99, dav99, spr05}). However, serious difficulties are present in reproducing the visible parts of galaxies, including the mass/luminosity functions at low and high masses, the angular momentum distribution of galactic disks, and the supermassive black hole (SMBH) and bulge-mass relationship (e.g., \\citealt{gil08, pri09, cat09}). For instance, the ratio of the visible baryonic mass to the gravitational mass of all galaxies is 3-10 times less than the cosmic value ($0.167\\pm0.006$) inferred from the {\\sl Wilkinson Microwave Anisotropy Probe} (\\citealt{kom09}; also see \\citealt{bre09} and references therein); the discrepancy seems to be more severe in less massive galaxies than in more massive ones (e.g., \\citealt{hoe05, mcg08}). This missing link between the intergalactic medium (IGM), dark halos, and visible galaxies is attributed largely to our poor understanding of the complex ``gastrophysics'', in particular the coupling between gas and galaxy feedback (e.g., \\citealt{dav07}). Galaxy feedback comes in many forms. Starbursts are known to generate galactic winds, which must be chiefly responsible for the chemical enrichment and non-gravitational heating of the IGM at high redshifts (e.g., \\citealt{mac99}). Nuclear starbursts are also believed to be intimately related to the formation and growth of the SMBHs, which are an additional source of feedback, especially in massive galaxies (e.g., \\citealt{str02, swa06}). While the radiation pressure in a high-accretion phase can be a key driving force of the winds, the mechanical energy injection in a radiatively inefficient accretion phase may balance the cooling of the surrounding gas. Feedback from stars, even from evolved ones alone (e.g., Type Ia supernovae), can play an essential role in maintaining hot gaseous halos around galaxies \\citep{tang09a, tang09b}. Cosmic-ray pressure can also play an important role in driving large-scale galactic outflows \\citep{eve08}. The relative importance of these various forms of the galaxy feedback, however, remains very unclear. Direct observational constraints on the physical process and history of the galaxy feedback are thus badly needed. A potentially effective approach is to observe the circumgalactic medium (CGM) around nearby galaxies. On scales from a few tens of kpc up to $\\sim 1$ Mpc, the CGM includes the gas that has been significantly affected by the galaxy feedback but is outside of the boundaries of the traditionally-known stellar and multi-phase interstellar components of such galaxies. It is quite possible that the CGM around the galaxies contains their ``missing'' baryons (e.g., \\citealt{deh98,som06}). Searches for the CGM have been conducted in ultraviolet (UV), but the association between galaxies and IGM absorbers still remains an unsettled issue. Within impact distances of $r_\\rho\\lsim350$ kpc from the sight lines of background QSOs, all galaxies with $L>0.1L^*$ seem to have ``associated'' Ly$\\alpha$ or other low ionized absorbers (e.g., \\citealt{chen08,wak09}). However, not all Ly$\\alpha$ absorbers have an associated galaxy found within 1$h^{-1}$ Mpc (e.g., \\citealt{mor93,sto06}). These absorbers could be either associated with faint galaxies that are beyond the detection limits of current galaxy surveys or not related to any galaxies at all (some absorbers are detected in voids; \\citealt{chen09, sto07}.) Also, it is possible that much of the injection into the IGM occurred at high redshifts, and in the time that has subsequently passed, substantial separations in space and velocity have developed between the ejected matter and the galaxies from which it originated \\citep{tri06}. The highly ionized absorbers, \\ovi\\ absorbers in particular, may trace the CGM at high temperatures. Covering fraction of these absorbers varies with respect to galaxy luminosities and the impact distances of these absorbers to their nearest galaxies \\citep{sto06, pro06, gan08, wak09}. X-ray observations could provide complementary or even key constraints on the existence and properties of the CGM. The bulk of the CGM, at least for galaxies with masses similar to and higher than our Galaxy, is believed to be in a gaseous phase at temperatures of $T\\gsim 10^6$~K and mainly emits and absorbs X-ray photons (e.g., \\citealt{bir03}). However, because of the low density and potentially low metallicity of the CGM, its X-ray emission, except for that in galaxy clusters where the missing baryon problem becomes less severe, is expected to be hard to detect (e.g., \\citealt{han06}), and the observational evidence is still absent (e.g., \\citealt{ras09, and10}). The high spectral resolution grating instruments aboard \\chandra\\ and \\xmm\\ now provide us with a new powerful tool --- X-ray absorption line spectroscopy --- to probe the low-surface brightness diffuse hot CGM. Unlike X-ray emission that is sensitive to the emission measure of the hot gas, absorption lines produced by helium- and hydrogen-like ions such as \\ovii, \\oviii, and \\neix\\ directly probe their column densities, which are proportional to the mass of the gas and are sensitive to its thermal, chemical, and kinematic properties. Indeed, both the \\chandra\\ and \\xmm\\ grating observations of bright AGNs and X-ray binaries have been used to firmly detect and characterize the global hot gas in and around the Milky Way (e.g., \\citealt{wang05, yao05, yao06, yao07, fang06, bre07}). However, our location in the midst of the hot gas makes it hard to determine the presence of the CGM around the Galaxy; our previous investigations only yield tentative upper limits to the column densities of the CGM beyond the hot gas-rich Galactic disk (e.g., \\citealt{yao08, yao09a}). In this work, we search for absorption lines of \\cvi, \\nvii, \\ovii, \\oviii, and \\neix\\ produced in the CGM of intervening galaxies along 12 AGN sight lines by extensively exploring \\chandra\\ grating observations. Because detection sensitivities of current high resolution X-ray instruments are still limited, to enhance the signal-to-noise (S/N) ratio, we stack spectra with respect to group membership and luminosity properties of intervening galaxies. This stacking technique, as implemented in our previous paper \\citep{yao09b}, enables us to probe the absorbing gas to unprecedented low column densities. The paper is organized as follows. In Section~\\ref{sec:data}, we discuss the background AGNs and survey of intervening galaxies, and we describe the data reduction process. We search for and measure the X-ray absorption lines produced in the CGM in Section~\\ref{sec:results}, and discuss the implications of our results % in Section~\\ref{sec:dis}. Throughout the paper, we adopt the Schechter luminosity function with characteristic luminosity $L_B^*\\simeq2\\times10^{10}L_\\odot$ (or absolute B magnitude $M_B^*\\simeq-20.5$) and use the cosmology with $H_0=73~{\\rm km~s^{-1}~Mpc^{-1}}$, $\\Omega_{M}=0.3$, and $\\Omega_{\\Lambda}=0.7$. ", "conclusions": "" }, "1005/1005.0112_arXiv.txt": { "abstract": "Many astrophysical phenomena are highly subsonic, requiring specialized numerical methods suitable for long-time integration. In a series of earlier papers we described the development of {\\tt MAESTRO}, a low Mach number stellar hydrodynamics code that can be used to simulate long-time, low-speed flows that would be prohibitively expensive to model using traditional compressible codes. {\\tt MAESTRO} is based on an equation set derived using low Mach number asymptotics; this equation set does not explicitly track acoustic waves and thus allows a significant increase in the time step. {\\tt MAESTRO} is suitable for two- and three-dimensional local atmospheric flows as well as three-dimensional full-star flows. Here, we continue the development of {\\tt MAESTRO} by incorporating adaptive mesh refinement (AMR). The primary difference between {\\tt MAESTRO} and other structured grid AMR approaches for incompressible and low Mach number flows is the presence of the time-dependent base state, whose evolution is coupled to the evolution of the full solution. We also describe how to incorporate the expansion of the base state for full-star flows, which involves a novel mapping technique between the one-dimensional base state and the Cartesian grid, as well as a number of overall improvements to the algorithm. We examine the efficiency and accuracy of our adaptive code, and demonstrate that it is suitable for further study of our initial scientific application, the convective phase of Type Ia supernovae. ", "introduction": "Many astrophysical phenomena of interest occur in the low Mach number regime, where the characteristic fluid velocity is small compared to the speed of sound. Some well-known examples are the convective phase of Type Ia supernovae (SN~Ia)~\\citep{hoflichstein:2002,kuhlen-ignition:2005,ZABNW:IV}, classical novae~\\citep{glasner:2007}, convection in stars~\\citep{meakin:2007}, and Type I X-ray bursts~\\citep{Lin:2006}. Such problems require a numerical approach capable of resolving phenomena over time scales much longer than the characteristic time required for an acoustic wave to propagate across the computational domain. In a series of papers (see \\citet{ABRZ:I}---henceforth Paper I, \\citet{ABRZ:II}---henceforth Paper II, \\citet{ABNZ:III}---henceforth Paper III, and \\citet{ZABNW:IV}---henceforth Paper IV), we have described the initial development of {\\tt MAESTRO}, a low Mach number hydrodynamics code for computing stellar flows using a time step constraint based on the fluid velocity rather than the sound speed. {\\tt MAESTRO} is suitable for two- and three-dimensional local atmospheric flows as well as three-dimensional full-star flows. All simulations are performed in a Cartesian grid framework, but rely on the presence of a one-dimensional radial base state that describes the average state of the star or atmosphere. Starting with the development of the low Mach number equation set (see Paper I), we demonstrated how to capture the expansion of the base state in a local atmospheric simulation in response to large-scale heating (Paper II) and incorporate reaction networks (Paper III). In Paper IV, we presented the initial application of {\\tt MAESTRO}, following the last two hours of convection inside a white dwarf leading up to the ignition of a SN Ia using a three-dimensional, full-star simulation with a base state that is constant in time. In general, astrophysical flows are highly turbulent. In the case of the convective period preceding a SN Ia explosion, the Reynolds number is $\\mathcal{O}(10^{14})$ \\citep{Woosley:2004}, far larger than can be modeled on today's supercomputers. Nevertheless, to understand the role of turbulence in these events, we must use increasingly more accurate simulations. In this paper we describe how to incorporate adaptive mesh refinement (AMR), in which we locally refine the Cartesian grid in regions of interest, to allow us to efficiently push to higher spatial resolutions and better capture the turbulent flow in critical regions of the simulation. The primary difference between {\\tt MAESTRO} and other structured grid AMR approaches for incompressible and low Mach number flows is the presence of the time-dependent base state, whose evolution is coupled to the evolution of the full solution. We also describe how to incorporate a time-dependent base state for full-star problems, which involves a novel mapping technique between the one-dimensional base state and the Cartesian grid. This allows us to properly capture the effects of an expanding base state in full-star simulations. We have also made a number of overall improvements to the algorithm, and all together, these enhancements will allow us to compute more efficient and accurate solutions for our target applications, including the convective phase of SNe Ia and Type I X-ray bursts. This paper is divided into several sections, along with three detailed appendices. In \\S \\ref{Sec:Governing Equations}, we present the governing equations. In \\S \\ref{Sec:Overview}, we give an overview of the methodology, referring the reader to the appendices for full details. In \\S \\ref{Sec:Mapping}, we describe the new mapping procedure between one-dimensional and three-dimensional data structures in full-star problems. In \\S \\ref{Sec:AMR}, we discuss the extension of the algorithm to include AMR. In \\S \\ref{Sec:Test Problems}, we describe the results of our test problems. We conclude with \\S \\ref{Sec:Conclusions}, which includes future plans for scientific investigation. ", "conclusions": "\\label{Sec:Conclusions} We have developed a low Mach number hydrodynamics algorithm suitable for full star flows and local atmospheric regions with a time-evolving base state within an AMR framework. In forthcoming papers, we will use {\\tt MAESTRO} to further our scientific investigation of the convective phase of SNe~Ia. Our previous simulations in Paper IV were at modest resolution and assumed a constant base state. We are now performing simulations at higher effective resolutions with the use of AMR along with a time-varying base state. As part of this study, we are examining the tagging conditions necessary to model a full-star up to the point of ignition. We are also studying Type I X-ray bursts \\citep{STRO_BILD06,Lin:2006}, which are believed to be caused by the thermonuclear explosive burning of hydrogen and/or helium gas accreted into a thin shell on the surface of neutron stars. We pose this problem in planar geometry, model a small patch of the star, and refine near the base of the accreted layer." }, "1005/1005.4507_arXiv.txt": { "abstract": "Swift opened up a new era in the study of gamma-ray burst sources (GRB). Among a variety of discoveries made possible by Swift, here we focus on GRB\\,090423, the event at z=8.2 which currently holds the record of the most distant celestial object ever caught by human instrumentation. This GRB allowed us to have a direct look at the early Universe. The central engine activity giving origin to the GRB emission is also discussed starting from the observational findings of an updated GRB X-ray flares catalog. ", "introduction": "Each time we build and use innovative state of the art instrumentation we expect to detect new phenomena and details that allows a better understanding of the physics of the objects we are observing. In the case of the Swift mission\\cite{Gehrels04} we were able to build on the experience obtained by Beppo-Sax and on the theoretical developments that helped the design allowing targeting critical points of the GRB physics. The result was a multi-wavelength mission with very fast pointing, accurate astrometry of the transients and real time analysis and communication of the detected targets.\\\\ Out of the variegated science and related high lights we have recently published, here we will discuss GRB\\,090423, a burst holding the high z record for any celestial object so far discovered, and illustrate some of the results we obtained with the new sample of GRB flares. In other words we will look into the early Universe and perhaps into the activity intimately related to the central engine.\\\\ Nowadays we have evidence that the early star formation, pop III, is followed by the re-ionization (of HI) phase that likely terminated at $z \\sim 6$. As it is well known even a small amount of neutral hydrogen would heavily absorbs at wavelengths smaller than Ly$_{\\alpha}$(Gunn - Peterson effect) so that this region in which some neutral hydrogen still exist in the IGM is easily recognizable and indeed it is a powerful tool to track down the various phases of the process as a function of redshift.\\\\ In the X-ray light curve of the afterglow (generally covering the temporal range of few hundreds second since trigger to a million second after that) we define as flare an event that manifests itself as a sudden change in the observed flux and lasting a time largely smaller than the light curve of the GRB itself. The cause of these later injections of energy is not yet known and while it may be caused either by late shocks among shells emitted during the prompt emission or, more likely, by renewed activity of the central engine, they are certainly not due to external shock (it can be shown that the pulse width divided by the time of occurrence of the pulse is too small to be explained as due to external shock). But no matter what, a clear understanding of the observed emission will lead to the understanding of the mechanism at work and eventually lead to the understanding of the central engine. Any correlation among parameters of flares occurring at different time in different bursts must relate to a behavior of the central engine and may lead to its partial understanding. ", "conclusions": "Recent observations challenge the fireball internal shock model. Optical observations do not show strong evidence of the reverse shock\\cite{Kumar07}, and in any case even if present does not seem to be one of the main component of the early optical emission\\cite{Oates09}. The reverse shock is weak or suppressed in the magneto hydro dynamical models of GRBs\\cite{Thompson94,Spruit01,Lyutikov03} so that the goal is to find way to estimate whether $\\sigma_{0}=\\frac{F_{p}}{F_{b}}=\\frac{B_{0}^{2}}{(4\\,\\pi\\,\\gamma_{0}\\,\\rho\\,c^{2})}$ is $\\gg 1$ (the jet is dominated by a Pointing flux) or $\\ll 1$ (the kinetic energy of the baryonic jet dominates).\\\\ \\begin{figure} \\begin{center} \\includegraphics[width=\\hsize]{variability.ps} \\caption{The relative variability of the flares as a function of the width to time ratio has been plotted. The limits shown by the various lines have been calculated according to the equations given in Ref.\\,\\protect\\refcite{Ioka05}. The 4 black dots refer to short GRBs.}\\label{fig:variability} \\end{center} \\end{figure} We should add incidentally, that the mechanism by which strong magnetic fields and the acceleration of particles to very high energy is achieved is not yet fully known. It needs to be demonstrated that the linear Fermi mechanism is capable of doing the job following the back and for transit of the particles in the shock region.\\\\ To summarize part of the results in a diagram we refer to plot of the ratio $\\Delta F/F$ versus $\\Delta t/t$, Figure\\,\\ref{fig:variability}. Here $\\Delta F$ is the variation of the flux above the underlying continuum over the flux observed in the underlying continuum and $\\Delta t/t$ the flare width over the time of the peak of the flare.\\\\ The complete absence of flares with $\\Delta t/t > 1$ confirms that flares cannot be due to patchy shells. Flares furthermore seem to have an high probability, as expected, to be observed off axis and a large number of flares would, according to this plot, agree with the refreshed shock model.\\\\ Flares are not necessarily the result of late central engine activity, but may be produced in the decelerating phase of the flow. High $\\sigma_{0}$ values may lead to MHD instabilities during the interaction with the interstellar medium. In this case the isotropic equivalent energy emitted in a single flare and produced by a single reconnection event, is related to the ratio width/$t_{peak}$ by the relation: \\begin{equation} E_{flare} \\leq 5\\,\\epsilon \\left(\\frac{width}{t_{peak}}\\right)^{3} \\frac{E_{forward\\,shock}}{\\alpha^{2}} \\end{equation} where $\\epsilon \\sim 0.1$ and $\\alpha$ in the range 2 - 4 depending on the density of the ISM. The flares we observe are not in contradiction within errors with this relation but, at the same time, they do not prove it.\\\\ In conclusion we find that flares seem to retain memory of the previous event so that, as time progresses, each flare is weaker and softer of the preceding one. And finally while we are making significant progress in characterizing these events observationally, at the moment there is no satisfactory model explaining their origin, evolution and energetics." }, "1005/1005.4677_arXiv.txt": { "abstract": "Based on our recent work on tidal tails of star clusters \\citep{Kuepper09} we investigate star clusters of a few $10^4 \\msun$ by means of velocity dispersion profiles and surface density profiles. We use a comprehensive set of $N$-body computations of star clusters on various orbits within a realistic tidal field to study the evolution of these profiles with time, and ongoing cluster dissolution.\\\\ From the velocity dispersion profiles we find that the population of potential escapers, i.e. energetically unbound stars inside the Jacobi radius, dominates clusters at radii above about 50\\% of the Jacobi radius. Beyond 70\\% of the Jacobi radius nearly all stars are energetically unbound. The velocity dispersion therefore significantly deviates from the predictions of simple equilibrium models in this regime. We furthermore argue that for this reason this part of a cluster cannot be used to detect a dark matter halo or deviations from Newtonian gravity.\\\\ By fitting templates to the about $10^4$ computed surface density profiles we estimate the accuracy which can be achieved in reconstructing the Jacobi radius of a cluster in this way. We find that the template of King (1962) works well for extended clusters on nearly circular orbits, but shows significant flaws in the case of eccentric cluster orbits. This we fix by extending this template with 3 more free parameters. Our template can reconstruct the tidal radius over all fitted ranges with an accuracy of about 10\\%, and is especially useful in the case of cluster data with a wide radial coverage and for clusters showing significant extra-tidal stellar populations. No other template that we have tried can yield comparable results over this range of cluster conditions. All templates fail to reconstruct tidal parameters of concentrated clusters, however.\\\\ Moreover, we find that the bulk of a cluster adjusts to the mean tidal field which it experiences and not to the tidal field at perigalacticon as has often been assumed in other investigations, i.e. a fitted tidal radius is a cluster's time average mean tidal radius and not its perigalactic one.\\\\ Furthermore, we study the tidal debris in the vicinity of the clusters and find it to be well represented by a power-law with a slope of -4 to -5. This steep slope we ascribe to the epicyclic motion of escaped stars in the tidal tails. Star clusters close to apogalacticon show a significantly shallower slope of up to -1, however. We suggest that clusters at apogalacticon can be identified by measuring this slope. ", "introduction": "Velocity dispersion profiles and surface density profiles are among the most basic tools for investigating the structure of star clusters. However, such investigations indicate that the region around the tidal radius, at which the internal acceleration of a star cluster is similar to the tidal acceleration due to the galactic tidal field, is particularly poorly understood.\\\\ Velocity dispersion profiles sometimes show peculiarities which have been discussed in the literature. \\citet{Drukier98}, for example, observed a flattening in the outer parts of the velocity dispersion profile of the Galactic globular cluster M15 which they interpreted as an effect of tidal heating by the general Galactic tide or by tidal shocks. \\citet{Scarpa03} also found a flattening of the velocity dispersion profile for $\\omega$ Cen and more recently for other Galactic globular clusters like NGC6171, NGC7099 and NGC288 \\citep{Scarpa07}. The deviation from an expected Keplerian fall-off in the velocity dispersion profile occurred in all clusters at radii where the internal gravitational acceleration is about $a_0 = 1.2 \\times 10^{-8}\\mbox{cm s}^{-2}$ and was therefore interpreted by Scarpa et al.~as a hint of Modified Newtonian Dynamics \\citep{Milgrom83} in globular clusters. Alternatively, they briefly discussed the possible effect of tidal heating or a dark matter halo on the cluster stars.\\\\\\ On the contrary, \\citet{McLaughlin03b} found that by fitting a \\citet{Wilson75} model, which has a less sharp cut-off at the tidal radius than the commonly used \\citet{King66} model, to the re-analyzed $\\omega$ Cen data, its velocity dispersion profile could be explained without modifying Newtonian gravity and without adding dark matter. Similar investigations of cluster profiles, e.g. \\citet{Lane09, Lane10} and \\citet{Baumgardt10}, were also not in favour of MOND.\\\\ But it is not only the velocity dispersion profiles of star clusters which behave strangely at the tidal boundary; their surface density profiles also show peculiarities and sometimes controversial behaviour.\\\\ \\citet{King62} showed that the surface density profiles of many globular clusters can be fitted by a simple analytical formula having a sharp cut-off radius, which could be interpreted as the tidal radius of the cluster. Later he derived a set of physically motivated models, with a similar cut-off radius corresponding to an energy cut-off in the energy distribution function of the cluster stars, which provided an even better fit to cluster profiles \\citep{King66}.\\\\ In contrast to that, \\citet{Elson87}, and more recently \\citet{Gouliermis10}, found that young massive clusters in the LMC show an exponential surface density profile without any tidal truncation at the expected tidal radius. They interpreted these findings as being due to tidal debris which was expelled at birth from the clusters and has not had time to disperse yet.\\\\ Furthermore, \\citet{Cote02}, \\citet{Carraro07} and \\citet{Carraro09} find the outer halo Milky-Way globular clusters Palomar 13, Whiting 1 and AM 4, respectively, to have a significant excess of stars at the tidal boundary, which makes any fit to the surface density data very inaccurate. For all clusters they find the radial surface density profile, $\\Sigma(R)$, to be well represented by a power-law $R^\\eta$ with slopes of about $\\eta \\simeq -1.8$. This excess of stars is interpreted by the authors as heavy mass loss in a final stage of dissolution. The same was found for Palomar 5, which is a well studied MW globular cluster close to the apogalacticon of its orbit \\citep{Odenkirchen03}.\\\\ Moreover, \\citet{McLaughlin05} showed that most globular clusters of the Milky Way, the LMC and SMC, as well as of the Fornax dwarf spheroidal, are more extended than can be explained by \\citet{King66} models, and therefore are better represented by \\citet{Wilson75} models. In this context they emphasised the lack of physical understanding of this phenomenon.\\\\ On the contrary, \\citet{Barmby09} found that of 23 young massive clusters in M31 most were better fitted by \\citet{King66} models as these clusters do not show extended haloes.\\\\ As a consequence of this lack of understanding, \\citet{Barmby09} asked in their investigation how robust the physical parameters are which were derived in such analyses. This they tried to estimate by analyzing artificial clusters in the same way as the real observations. A similar analysis has also been performed by \\citet{Bonatto08}, although they tested the analytic formula of \\citet{King62} with artificial observations under various limiting conditions. Both investigations came to the conclusion that physical parameters can in principle be well recovered from such idealized mock observations.\\\\ But probably the idealized nature of these investigations is misleading, since both tests were performed with dynamically unevolved clusters. However, a self-consistent test of deriving physical parameters from a set of numerical computations of star clusters with a range of initial parameters has not been performed yet. Just a few investigations have touched this topic so far by means of numerical computations (e.g. \\citealt{Capuzzo05}, \\citealt{Drukier07}, \\citealt{Trenti10}). This is due to the fact that fast codes for globular cluster integration like Fokker-Planck or Monte-Carlo codes are not able to address this problem properly, as they cannot follow the evolution of the tidal debris and are restricted to cut-off criteria in energy or angular momentum space. Moreover, limits in computational power prevented us from carrying out such investigations by means of collisional $N$-body codes. But the recent improvements in computational speed of $N$-body codes and the availability of accelerator hardware (GRAPEs, GPUs) now allows us to study the dynamical evolution of star clusters with masses of up to several times $10^4\\msun$.\\\\ Thus, by computing various star clusters in a range of tidal conditions over several Gyr we investigate the structure of star clusters, and in particular the evolution of the region around the tidal radius - the transition region between the star cluster and the tidal tails - in terms of velocity dispersion and surface density profiles. The paper is organized as follows: first, a brief introduction to the topic of potential escapers is given in Sec.~\\ref{Sec:Pesc} as these stars play a key role in this investigation. Then a few methodological remarks will be made in Sec.~\\ref{Sec:Method} before we come to the velocity dispersion profiles in Sec.~\\ref{Sec:VDP} and then to the surface density profiles in Sec.~\\ref{Sec:SDP}. In the last section we will give a summary with a brief discussion. ", "conclusions": "\\label{Sec:Conclusions} This systematic investigation of velocity dispersion profiles and surface density profiles of star clusters with masses of about $10^4 \\msun$ has shown how complex and versatile these profiles can be, depending on the circumstances of the clusters.\\\\ Furthermore we have shown the importance of potential escapers for these two quantities. Both profiles are influenced by this population of energetically unbound stars for radii larger than about 50\\% of the Jacobi radius and are completely dominated by it for radii larger than about 70\\% of the Jacobi radius (Fig.~\\ref{vdps} \\& \\ref{vdpse075}). \\citet{Baumgardt01} found that for clusters in a constant tidal field the fraction of potential escapers varies with the number of stars within a cluster approximately as $N^{-1/4}$. Thus, even if the cluster mass increases by a factor of 10, the fraction (and influence) of potential escapers will only be reduced by less than 50\\% compared to this investigation.\\\\ As a consequence of the potential escaper population, the velocity dispersion profiles in our sample do not show a clear separation between the bulk of the cluster and the tidal debris at the Jacobi radius, but rather show a smooth transition from cluster to debris. For mass estimates which are based on velocity dispersion measurements we therefore recommend to use stars within 50\\% of the Jacobi radius to minimise the effect of potential escapers.\\\\ Moreover, we argue that investigations on velocity dispersion profiles like \\citet{Drukier98}, \\citet{Scarpa03} or \\citep{Scarpa07} have to be interpreted with caution. Also, detecting a possible dark matter halo around globular clusters is less feasible due to the population of potential escapers. Based on our investigation we suggest that no deviation from Newtonian gravity in a star cluster has been detected so far, neither has there been evidence for a dark matter halo.\\\\ From our large set of surface density profiles we saw that in most cases the structure of extended star clusters can be split into core, bulk and tidal debris. \\begin{enumerate} \\item The core is the innermost part of the cluster which extends out to the core radius. In most cases it is flat, and only for clusters with strong mass segregation or with an IMBH is it cuspy. Very concentrated clusters with $R_{hp}/R_J < 0.1$ have such a small core that fitting such clusters leads to problems with templates as well as with physical models (see also \\citealt{Baumgardt10}).\\\\\\\\ \\item The bulk contains most stars of the cluster and extends from the core to the tidal debris. We found that the bulk can be well represented by a King-like template. We named the radius at which such a template approaches zero density the edge of the bulk and not the tidal radius of the cluster, since we found it not to be consistent with the Jacobi radius. In the case of concentrated clusters, the edge radius is not even determined by tidal forces at all (Tab.~\\ref{table5}).\\\\ Moreover, for clusters on eccentric orbits we found that the edge radius of a cluster adjusts to the mean Jacobi radius (Fig.~\\ref{rt8500ecc}) and not the perigalactic Jacobi radius, which has been assumed in earlier investigations (e.g. \\citealt{Innanen83}, \\citealt{Fall01}), but has never been checked by self-consistent calculations on a star-by-star basis. Since nearly all globular clusters of the Milky Way are more compact and massive than our test cluster, and thus should be less influenced by tidal variations, this finding should also hold for most of these clusters.\\\\ The edge radius of extended and moderately concentrated clusters, as fitted by the King template, lies within about 10\\% accuracy at ~90\\% of the mean Jacobi radius (Tab.~\\ref{table3}-\\ref{table4c}). But we found the King template to be a bad representation of the true profile in the case of non-circular cluster orbits because it is significantly influenced by the tidal debris and by the potential escaper population. Furthermore the accuracy of the results strongly depends on the radial coverage of the cluster. We therefore created an enhanced version of the King template which has more flexibility in the core and has an additional tidal debris term, given by a power-law. With this KKBH template we achieved a higher stability over all fit ranges. Furthermore, we were able to measure the edge of the bulk more accurately and found it to lie at 80\\% of the Jacobi radius. This discrepancy is easily understandable if we take into account the fact that the edge radius of a template fits the azimuthally averaged mean of a cluster's tidal surface, whereas the Jacobi radius is by definition the semi-major axis of the equipotential surface.\\\\ Furthermore we found that the edge radii from concentrated clusters with $R_{hp}/R_J < 0.1$ cannot be extracted properly with currently available templates. The higher the concentration of the cluster the smaller is the ratio of the fitted edge radius to the theoretically evaluated Jacobi radius (Tab.~\\ref{table5}).\\\\\\\\ \\item The tidal debris falls off like a power-law with a slope of about -4 to -5, rather than the expected dependence of $R^{-1}$. This is due to the epicyclic motion of the stars in the tidal tails, which is most pronounced in the vicinity of the cluster.\\\\ For clusters on eccentric orbits, at a time shortly before reaching apogalacticon, the slope was found to deviate for a short time to a value of about -2 as a consequence of orbital compression of the tails (Fig.~\\ref{eta8500ecc} \\& \\ref{sdp17000}). We suggest that this enhanced slope can be used to identify star clusters which are close to or at the apogalacticon of their orbit. The most prominent example of such a cluster is the Milky Way globular cluster Palomar 5 which is well known to be close to the apogalacticon of its orbit \\citep{Dehnen04}, and which shows a power-law slope of -1.5 outside its assumed Jacobi radius \\citep{Odenkirchen03}. Moreover, the MW globular clusters Palomar 13, AM 4 and Whiting 1 may be further candidates for being close to apogalacticon, since their surface density profiles show power-law slopes of about -1.8 outside the Jacobi radius, which has been attributed to heavy ongoing mass loss by the authors of the corresponding investigations \\citep{Cote02, Carraro07, Carraro09}.\\\\ Furthermore we showed that due to the population of potential escapers the power-law slope of the tidal debris begins even before the Jacobi radius is reached. Thus, the bulk and the tidal debris partially overlap. With our KKBH template we found the debris to dominate for radii larger than about 50\\% of the Jacobi radius (Tab.~\\ref{table3}-\\ref{table4c}). \\end{enumerate} With the help of our comprehensive set of clusters it is possible to tailor a star cluster template to the facts we have found. Our KKBH template seems to be a good first step but, like all other available templates, shows discrepancies for concentrated clusters. Of course, templates are less attractive than physically motivated models. None of the existing models, like \\citet{King66} and \\citet{Wilson75}, however, account for the above mentioned facts. We suggest that a general, physically motivated model for star clusters should include a term for mass segregation and should include a potential escaper population when being fitted to observations. Finally, the possible influence of (compressed) tidal tails on the surface density profiles of star clusters should be kept in mind when fitting surface density profiles.\\\\ Moreover, the fact that all clusters in the sample showed a constant edge of the bulk with time (Fig.~\\ref{rt8500ecc}) and that stars which get unbound escape from a cluster with a delay, supports our theoretical treatment of epicyclic overdensities in tidal tails for clusters in time-dependent tidal fields in \\citet{Kuepper09}." }, "1005/1005.3057_arXiv.txt": { "abstract": "{We present a resolved dust analysis of three of the largest angular size spiral galaxies, NGC 4501 and NGC 4567/8, in the \\textit{Herschel} Virgo Cluster Survey (HeViCS) Science Demonstration field. \\textit{Herschel} has unprecedented spatial resolution at far-infrared wavelengths and with the PACS and SPIRE instruments samples both sides of the peak in the far infrared spectral energy distribution (SED). We present maps of dust temperature, dust mass, and gas-to-dust ratio, produced by fitting modified black bodies to the SED for each pixel. We find that the distribution of dust temperature in both systems is in the range $\\sim$19 - 22~K and peaks away from the centres of the galaxies. The distribution of dust mass in both systems is symmetrical and exhibits a single peak coincident with the galaxy centres. This Letter provides a first insight into the future analysis possible with a large sample of resolved galaxies to be observed by \\textit{Herschel}.} ", "introduction": "Infrared data have been widely used to determine the composition and distribution of dust in galaxies since the launch of the Infrared Astronomical Satellite (IRAS) in the 1980s. However, dust masses for nearby galaxies calculated from IRAS 60 and 100\\micron\\ measurements were found to be a factor of ten lower than expected when compared to the Milky Way gas-to-dust ratio of 100-200 \\citep{devereux1990}. The Milky Way dust mass was calculated by measuring the depletion of metals from the gaseous phase of the interstellar medium (ISM) and by comparing gas column densities to dust extinction \\citep{whittet03}, its value implying that most of the dust-mass emits radiation at wavelengths longer than 100\\micron\\ \\citep[e.g.,][]{devereux1990}. Analyses of data from the {\\em Spitzer Space Telescope\\/} have determined gas-to-dust ratios of $\\sim$150 \\citep[e.g.,][]{draine07}. However, these analyses were generally limited by the number, and low signal-to-noise ratio of data at wavelengths greater than 160\\micron. These studies therefore had difficulty detecting emission from dust with temperatures lower than 15~K, and their results were biased towards warmer dust temperatures and lower masses. Owing to the difficulties at these wavelengths, high resolution studies of galaxies have been previously limited. These studies are important for understanding how dust interacts with the other phases of the ISM, the sources of dust heating and how the distribution and temperature of dust varies with morphology. The \\textit{Herschel Space Telescope} \\citep{pilbratt10}, with its two photometric instruments, constrains both sides of the peak in the far infrared spectral energy distribution (SED, see Fig. \\ref{fig:SEDs}). The Photodetector Array Camera \\& Spectrometer (PACS, \\citealp{Poglitsch10}) has three photometric bands at 70, 100, and 160\\micron\\ at superior angular resolutions to those provided by \\Spitzer. The Spectral and Photometric Imaging Receiver (SPIRE, \\citealp{Griffin10}) also has three photometric bands observing simultaneously at 250, 350, and 500\\micron\\ with high sensitivity and angular resolution. The Herschel Virgo Cluster Survey (HeViCS\\footnote{More details on HeViCS can be found at http://www.hevics.org}) is a \\textit{Herschel} Open Time Key Project that will observe $\\sim$64~deg$^2$ of the Virgo cluster. This will provide a large sample of resolved galaxies because about 48 late-type galaxies will be observed by HeViCS with optical diameters larger than 3\\arcmin. In this Letter, we present an insight into what will be possible with the full HeViCS survey by applying a resolved dust analysis to infer dust temperatures, surface densities, and gas-to-dust ratios for NGC 4501 and NGC 4567/8. These galaxies were chosen because they are among the largest angular size systems in the HeViCS Science Demonstration Phase (SDP) field. In Sect. 2, we present the observations and data reduction, and in Sects. 3 and 4, we present our analysis and results, respectively. ", "conclusions": "We have fitted SEDs on a pixel-by-pixel basis to investigate the dust mass, temperature, and gas-to-dust ratios of NGC 4501 and NGC 4567/8. We have measured dust surface densities that peak at the centre of the galaxy and decrease towards the outer regions. In contrast, the temperature distribution is asymmetric, higher temperatures peaking at some distance away from the centre of the galaxy and then decreasing with increasing radius towards the outskirts. Once the full HeViCS observations are complete, we will be able to extend the analysis to larger radii and lower dust surface densities. For $\\sim$48 late-type galaxies with optical radii greater than 3\\arcmin\\ to be observed with HeViCS, we will be able to begin to tackle questions about the dominant source of dust heating, how the dust interacts with the other phases of the ISM, and how morphology influences the dust distribution and temperature. By combining the results from HeViCS and the Herschel Reference Survey \\citep{boselli10}, we will be able to study how environment affects the dust distribution and other properties of a galaxy." }, "1005/1005.5098_arXiv.txt": { "abstract": "{The recently discovered exoplanet Gl581d is extremely close to the outer edge of its system's habitable zone, which has led to much speculation on its possible climate. We have performed a range of simulations to assess whether, given simple combinations of chemically stable greenhouse gases, the planet could sustain liquid water on its surface. For best estimates of the surface gravity, surface albedo and cloud coverage, we find that less than 10 bars of CO$_2$ is sufficient to maintain a global mean temperature above the melting point of water. Furthermore, even with the most conservative choices of these parameters, we calculate temperatures above the water melting point for CO$_2$ partial pressures greater than about 40 bar. However, we note that as Gl581d is probably in a tidally resonant orbit, further simulations in 3D are required to test whether such atmospheric conditions are stable against the collapse of CO$_2$ on the surface.} ", "introduction": "In 2007, radial velocity measurements were used to discover two new planets in the Gl581 system \\citep{udry2007}. These planets have captured much attention both in the community and among the general public, as their minimum masses were measured to be below 10 $M_{Earth}$, and they are close to the edges of their system's nominal `habitable zone', i.e., the loosely defined orbital region in which planets can sustain liquid water on their surfaces. The first planet, Gl581c, which is closer to its star and was the first discovered, was initially estimated to be potentially habitable based on its equilibrium temperature $T_{eq}=320$ K, using an Earth-like planetary albedo 0.29. In contrast, the second planet Gl581d has an equilibrium temperature $T_{eq}= 195$ K for an albedo of 0.2, which suggests it may be too cold to sustain surface liquid water. However, these analyses neglect any possible warming of the surface due to the planet's atmosphere. In the first detailed assessment of the potential habitability of these planets, \\citet{selsis2007} reviewed a variety of factors that could influence their climates. They concluded that based on standard assumptions of atmospheric warming by a mixture of CO$_2$ and H$_2$O (with possible regulation of CO$_2$ via the carbonate-silicate cycle), Gl581c was unlikely to be habitable, while for Gl581d, the situation was much less clear. According to \\citet{kasting1993}, the outer edge of the habitability zone is most likely the distance at which CO$_2$ condensation begins to occur on the surface of the planet. However, CO$_2$ condensation in the atmosphere leads to the formation of CO$_2$ clouds, which can cause a strong warming effect due to the scattering of infrared radiation \\citep{forget1997}. Hence it was concluded that further climate simulations were required. To investigate the possible climate of Gl581d under a range of conditions, we have performed one-dimensional radiative-convective calculations. In Section \\ref{sec:method} we discuss the model we used, while in Section \\ref{sec:results} we present our results for varying atmospheric compositions, surface albedos, gravity and cloud coverage. We also present some simple three-dimensional simulations that highlight the limitations of the one-dimensional globally averaged approach. In Section \\ref{sec:discuss}, we discuss the implications of our results and suggest directions for future research. ", "conclusions": "\\label{sec:discuss} Through one-dimensional radiative-convective climate modeling, we have found that as little as 5 bars of CO$_2$ ($\\sim$2.5 bar equivalent column amount under Earth gravity) may be sufficient to maintain a global mean temperature above the melting point of water. For CO$_2$ partial pressures greater than 40 bar ($\\sim$13 bar equivalent column amount under Earth gravity), we found that the mean surface temperature of Gliese 581d is high enough to allow surface liquid water even when the most conservative values of gravity, surface albedo and N$_2$ partial pressure are chosen, although strong dependence on the total H$_2$O cloud opacity was observed. We tested the dependence of surface temperature on a wide range of atmospheric and geophysical parameters. Because the M-class spectrum of Gliese 581 is red-shifted with respect to that of the Sun, heating due to near-IR absorption in the upper atmosphere increases in our simulations, while Rayleigh scattering is much less important. This is in agreement with the general habitability study of \\citet{kasting1993}, but we find the difference to be even stronger than was reported there, due to the low effective temperature of Gl581. We also estimate an modest warming effect due to CO$_2$ clouds (up to 30 K at $p_{CO_2}$ = 50 bar), dependent on details of the cloud microphysics, although the efficiency of this process is reduced compared to the case for G-class stars. We find that increasing gravity primarily decreases the CO$_2$ column amount at which the maximum surface temperature occurs. Unsurprisingly, increased surface albedo decreases the mean surface temperature, although if there is cloud coverage, this effect is relatively small for the high CO$_2$ pressures at which the surface starts to become habitable. Increasing the partial pressure of N$_2$ has little effect below partial pressures $p_{N_2}$ of tens of bars. Finally, water vapour increases the surface temperature by a moderate amount below $T_s = 273 $ K, after which its role becomes increasingly important. We briefly assessed the impact of H$_2$O clouds, and found that they warmed (cooled) the surface at low (high) pressures, due to the interplay between their effects on incoming stellar and outgoing thermal radiation. As discussed by \\citet{selsis2007}, N$_2$-CO$_2$-H$_2$O atmospheres likely only represent a fraction of the possible range for Earth-like / super-Earth extrasolar planets. While the presence of radiatively inactive gases such as argon would be of little importance to the calculations discussed here, small amounts of other greenhouse gases such as methane would result in increased absorption, if such gases could remain chemically stable in the atmosphere. Unless the near-IR absorption of these gases were large enough to cause a strong temperature inversion in the upper atmosphere, this would further increase the surface temperature of the planet. The presence of other clouds or haze layers could alter the radiative balance in less predictable ways, depending on their scattering / absorption properties and height in the atmosphere. Unfortunately, effects of this kind are difficult to constrain further until direct atmospheric observations become available. The results discussed here clearly have interesting implications, as they show that Gl581d could be the first discovered habitable exoplanet. They are also testable by future observations - dense CO$_2$ atmospheres have recognizable spectral signatures that could be detected, for example, using the proposed Darwin or TPF missions. However, there is an important limitation to this study: all simulations performed were one-dimensional. As we noted in Section \\ref{subsec:3Dnoatm}, calculations of the \\emph{globally averaged} surface temperature neglect variations due to local changes in stellar insolation (as well as topography and other effects). This has serious implications for dense, relatively cold CO$_2$ atmospheres, as if the dark side temperature of the planet is too low, CO$_2$ will condense on the surface. To test whether a dense atmosphere on Gl518d would be stable against the collapse of CO$_2$ in this way, we plan to repeat the calculations reported here in the future using a three-dimensional climate model." }, "1005/1005.2392_arXiv.txt": { "abstract": "Inertial waves governed by Coriolis forces may play an important role in several astrophysical settings, such as eg. tidal interactions, which may occur in extrasolar planetary systems and close binary systems, or in rotating compact objects emitting gravitational waves. Additionally, they are of interest in other research fields, eg. in geophysics. However, their analysis is complicated by the fact that in the inviscid case the normal mode spectrum is either everywhere dense or continuous in any frequency interval contained within the inertial range. Moreover, the equations governing the corresponding eigenproblem are, in general, non-separable. In this paper we develop a consistent WKBJ formalism, together with a formal first order perturbation theory for calculating the properties of the normal modes of a uniformly rotating coreless body (modelled as a polytrope and referred hereafter to as a planet) under the assumption of a spherically symmetric structure. The eigenfrequencies, spatial form of the associated eigenfunctions and other properties we obtained analytically using the WKBJ eigenfunctions are in good agreement with corresponding results obtained by numerical means for a variety of planet models even for global modes with a large scale distribution of perturbed quantities. This indicates that even though they are embedded in a dense spectrum, such modes can be identified and followed as model parameters changed and that first order perturbation theory can be applied. This is used to estimate corrections to the eigenfrequencies as a consequence of the anelastic approximation, which we argue here to be small when the rotation frequency is small. These are compared with simulation results in an accompanying paper with a good agreement between theoretical and numerical results. The results reported here may provide a basis of theoretical investigations of inertial waves in many astrophysical and other applications, where a rotating body can be modelled as a uniformly rotating barotropic object, for which the density has, close to its surface, an approximately power law dependence on distance from the surface. ", "introduction": "In astrophysical applications inertial waves that can exist in rotating bodies may be excited by several different physical mechanisms, most notably through tidal perturbation by a companion (eg. Papaloizou \\& Pringle 1981, hereafter PP) or in the case of compact objects through secular instability arising through gravitational wave losses (eg. Chandrasekhar 1970, Friedman \\& Schutz 1978, Andersson 1998, Friedman $\\&$ Morsink 1998). They also can play a role in other physical systems. For example, they can also be excited by several mechanisms in the Earth's fluid core with possible detection being announced (Aldridge $\\&$ Lumb 1987). For rotating planets and stars that have a barotropic equation of state these wave modes are governed by Coriolis forces and so have oscillation periods that are comparable to the rotation period. They are accordingly readily excited by tidal interaction with a perturbing body when the characteristic time associated with the orbit is comparable to the rotation period, which is expected naturally when the rotation period and orbit become tidally coupled. They may then play an important role in governing the secular orbital evolution of the system. Inertial modes excited in close binary systems in circular orbit were considered by PP and Savonije \\& Papaloizou (1997). Wu (2005)a,b considered the excitation of inertial modes in Jupiter as a result of tidal interaction with a satellite and excitation as a result of a parabolic encounter of a planet or star with a central star was studied by Papaloizou \\& Ivanov (2005), hereafter referred to as PI and Ivanov \\& Papaloizou (2007), hereafter referred to as IP. The latter work was applied to the problem of circularisation of extrasolar giant planets starting with high eccentricity. In that work the planet was assumed coreless. Ogilvie \\& Lin (2004) and Ogilvie (2009) have considered the case of a cored planet in circular orbit around a central star and found that inertial waves play an important role. The importance of the role played by inertial waves in the transfer of the rotational energy of a rotating neutron star to gravitational waves via the Chandrasekhar-Friedman-Schutz (CFS) instability was pointed out by Andersson (1998). Later studies mainly concentrated on physical mechanisms of dissipation of energy stored in these modes that limit amplitudes of the modes, and, consequently, the strength of the gravitational wave signal. In these studies either numerical methods or simple local estimates of properties of inertial modes were mainly used, see eg. Kokkotas (2008) for a recent review and references. An analytical treatment of problems related to inertial waves, such as eg. finding normal mode spectra and eigenfunctions, and coupling them to other physical fields, etc., is difficult due to a number of principal complicating technical issues. In particular, the dynamical equations governing the perturbations of a rotating body (called planet later on) are, in general, non-separable, for compressible fluids. When such fluids are considered and rotation is assumed to be small, a low frequency anelastic approximation that filters out the high frequency modes is often used (see eg. PP). This simplifies the problem to finding solutions to leading order in the small parameter $R_{*}^{3}\\Omega^{2}/(GM)$, where $\\Omega $ is the rotation frequency, $G$ is the constant of gravity and $M_{*}$, $R_{*}$ are the mass and radius of the planet. In this approximation eigenfrequencies of inertial modes are proportional to $\\Omega ,$ while the form of the spatial distribution of perturbed quantities does not depend on the rotation rate. However, even when this approximation is adopted, the problem is, in general, non-separable apart from models with a special form of density distribution, see Arras et al (2003), Wu (2005)a and below. Additionally, the problem of calculating the inertial mode spectrum and its response to tidal forcing is complicated by the fact that in the inviscid case the spectrum is either everywhere dense or continuous in any frequency interval it spans (Papaloizou \\& Pringle, 1982). This is in contrast to the situation of, for example, high frequency $p$ modes, which are discrete with well separated eigenvalues. When the anelastic approximation is adopted the singular ill posed nature of the inviscid eigenvalue problem is seen to come from the fact that the governing equation is hyperbolic and the nature of the spectrum is determined by the properties of the characteristics (eg. Wood 1977). A discrete spectrum is believed to occur when there are no such trajectories that define periodic attractors. Otherwise the inviscid spectrum is continuous. Then, when a small viscosity is introduced the spectrum becomes discrete but normal modes have energy focused onto wave attractors (see eg. Ogilvie \\& Lin 2004). Given these complexities it is desirable to work with and compare a variety of analytical and numerical approaches. Coreless inviscid rotating planets with an assumed spherical or ellipsoidal shape have a discrete but everywhere dense spectrum that makes difficulties for example with mode identification and application of standard perturbation theory. However, numerical work indicates that there are well defined global modes that can be identified and followed through a sequence of models (eg. Lockitch \\& Friedman, 1999, hereafter LF, and PI). In this paper we investigate the inertial mode spectrum of a uniformly rotating coreless barotropic planet or star and its tidal response by a WKBJ approach coupled with first order perturbation theory and compare its eigenvalue predictions with numerical results obtained by a variety of authors and find good agreement apart from some unidentified WKBJ modes that are near the limits of the spectrum and for which the perturbation theory appears not to work. For the identified modes we also find remarkably good agreement for the form of the eigenfunctions. This indicates that they can be represented at low resolution with small scale phenomena being unimportant, meaningful mode identification (in that the modes can be followed from one model to another) and at least first order perturbation theory works for these modes. This is also confirmed in a following paper (hereafter referred to as PIN) where we investigate the inertial mode spectrum and its tidal response by numerical solution of an initial value problem {\\it without the anelastic approximation}. We are able to confirm the validity of the anelastic approximation and the applicability of the first order perturbation theory developed here for demonstrating this as well as estimating eigenvalues. Thus a suggestion of Goodman \\& Lackner (2009) that tidal interaction might be seriously overestimated by use of the anelastic approximation is not confirmed. A WKBJ approach to the same problem was also considered by Arras et al (2003) and Wu (2005)a. However, in this work only terms of leading order in an expansion in inverse powers of a large WKBJ parameter $\\lambda $ (see the text below for its definition) were taken into account and treatment of perturbations near the surface and close to the rotational axis were oversimplified. As a consequence, although their results are correct in the formal limit $\\lambda \\rightarrow \\infty, $ they cannot be used to make a correspondence between WKBJ modes and those obtained numerically, or an approximate description of modes with a scale that is not very small. In this paper we treat the problem in a more extended way, considering terms of the next $O(\\lambda^{-1})$ order together with an accurate treatment of perturbations near the surface and close to the rotation axis. Additionally, we consider a frequency correction of the next order, $O(\\lambda^{-2}),$ for modes having non-zero azimuthal number, $m$. We checked results obtained with use of the WKBJ formalism against practically all numerical data existing in the literature finding good agreement in practically all cases. Therefore, we can assume that our formalism may be applied to provide an approximate analytic description of inertial modes, including those with large scale variations, where the WKBJ approach might be expected to be invalid. Also, different quantities associated with the modes may be described within the framework of our formalism or its natural extension, such as the tidal overlap integrals (see PI and IP), quantities determining the growth rate due to the CFS instability and decay of inertial waves due different processes, eg. by non-linear mode-mode interactions (see eg. Schenk et al 2002, Arras et al 2003). Thus, the formalism developed here may provide a basis for the analytic treatment of inertial waves in many different astrophysical applications. The plan of the paper is as follows. In section \\ref{sec2} we briefly review the basic equations and their linearised form for a uniformly rotating barotropic planet or star. In section \\ref{anelastic} we go on to consider these in the anelastic approximation which is appropriate when the rotation frequency of the star is very much less than the critical or break up rotation frequency. We give a simple physical argument why we expect this approximation to be valid in this limit even when the sound speed tends to a small value or possibly zero at the surface of the configuration. In section \\ref{sec2.5} we give a brief discussion about when discrete normal modes may be expected to occur such as in the case of a coreless slowly rotating planet with surface boundary assumed to be either spherical or ellipsoidal. We then present a formal first order perturbation theory that can be used to estimate corrections to eigenfrequencies occurring as either a consequence of terms neglected in the WKBJ approximation or the anelastic approximation. The latter application is tested by a direct comparison with the results of numerical simulations in PIN. Section \\ref{sec2.6} concludes with a brief account of the form of the anelastic equations in pseudo-spheroidal coordinates in which they become separable for density profiles of the form $\\rho \\propto (1-r^2/R_*^2)^{\\beta},$ where $r$ is the local radius, $R_*$ is the surface radius and $\\beta$ is a constant. (Arras et al 2003, Wu 2005a). In section \\ref{sec3} we develop a WKBJ approximation for calculating the normal modes which is based on the idea that in the short wavelength limit these modes coincide with those appropriate to separable cases which include the homogeneous incompressible sphere as a well known example. Solutions of a general WKBJ form appropriate to the interior of the sphere are matched to solutions appropriate to the surface regions where they become separable which is the case when the density vanishes as a power of the distance to the boundary as is expected for a polytropic equation of state. This matching results in an expression for the eigenfrequencies given in section \\ref{sec3.5}. In section \\ref{surface} we go on to develop expressions for the eigenfunctions appropriate to any location in the planet including the rotation axis and the critical latitude region where one of the inertial mode characteristics is tangential to the planet surface. These solutions are then used to obtain corrections to the eigenfrequencies resulting from density gradient terms neglected in the initial WKBJ approximation in section \\ref{sec3.9}. In section \\ref{sec4} we compare the corrected eigenfrequencies obtained from the WKBJ approximation with those obtained numerically by several different authors who used differing numerical approaches and find good agreement even for global modes. A similar comparison with the results of numerical simulations for a polytropic model with positive results is reported in PIN. We also compare the forms of the eigenfunctions with those obtained in Ivanov \\& Papaloizou (2007) and find a good agreement even for global modes. {Finally in section \\ref{sec5.1} we discuss our results in the context of the evaluation of the overlap integrals that occur in evaluating the response to tidal forcing. We show that these vanish smoothly in the limit that the polytropic index tends to zero and we indicate that they vanish at the lowest WKBJ order and are thus expected to vanish rapidly as the order of the mode increases. We go on to summarize our conclusions in section \\ref{Conclu}.} ", "conclusions": "\\label{sec5} \\subsection{Overlap integrals}\\label{sec5.1} { As we pointed out in the Introduction, integrals of the form \\begin{equation} \\hat Q_{k}={Q_{k}\\over \\sqrt N_{k}} \\quad {\\rm with}\\quad Q_{k}= ({\\rho \\over c_{s}^{2}}W_{k}|\\Phi), \\label{eqn84} \\end{equation} where $W_{k}$ corresponds to a particular eigenmode and $\\Phi $ is some smooth function, appear in astrophysical applications of the theory developed in this paper. In particular, as was discussed in PI and IP, integrals of this type enter in expressions for the transfer of energy and angular momentum transferred during the periastron passage of a massive perturber. These apply to the case when the spectrum of normal modes is discrete and they involve integrals of form (\\ref{eqn84}), where $\\Phi=P_{2}^{m}r^{2}$ with $P_{2}^{m}$ being the associated Legendre function. Assuming that $W_{k}$ varies on a small spatial scale while the function $\\Phi $ is smoothly varying such integrals may, in principal, be evaluated using our formalism with help of a theory of asymptotic evaluation of multidimensional integrals, see eg. Fedoryuk (1987), Wong (1989). However, some important integrals of form (\\ref{eqn84}) require an extension of our formalism, which can provide a smooth matching of the solution close to the surface to the WKBJ solution in the inner part of the planet that is valid at the next orders in inverse powers of $\\lambda $. This is due to cancellations of leading terms in corresponding asymptotic series. Since this problem appears to be a rather generic one we would like to discuss it here in more detail for the important case when $\\Phi =P^{2}_{2}r^2=3\\varpi^{2}$. The overlap integral of this type determines excitation of the $m=2$ modes which are the most important for the tidal problem (eg. PI, IP and see also PIN). Explicitly, we have in this case \\begin{equation} Q_{k}= 3\\int dV \\left(\\varpi^2 {\\rho \\over c_{s}^{2}}W_{k}\\right), \\label{eqn84a} \\end{equation} where $dV=dz\\varpi d\\varpi $. Note that this integral must converge to zero in the incompressible limit $n\\rightarrow 0$ as in this case it is well known that inertial modes are not excited in the anelastic approximation. This fact, however, is not obvious for the integral written in the form (\\ref{eqn84a}) because close to the surface we have \\begin{equation} {\\rho \\over c_{s}^{2}}\\approx nCx^{(n-1)}, \\label{eqn85} \\end{equation} with the constant $C$ converging to a nonzero value as $n \\rightarrow 0.$ Therefore, as the eigenfunctions are regular, the integral $Q_{k}/n$ has a logarithmic divergence at the surface of the planet as $n \\rightarrow 0.$ This raises the possibility that the overlap integral might converge to a nonzero value or behave pathologically as the incompressible limit is approached. In order to show that, in fact, this is not so and $Q_{k}(n\\rightarrow 0)\\rightarrow 0$ in a smooth manner, let us consider some fiducial models having the property that the quantity \\begin{equation} \\omega_0^2 = -\\frac{c^2_s}{r \\rho }\\frac{d\\rho }{dr}\\label{Good1} \\end{equation} is constant. For models in hydrostatic equilibrium under their own gravity, constancy of $\\omega_0^2$ implies that ratio $M(r)/r^{3}$, where $M(r)$ is the mass interior to the radius $r,$ is constant. The model must accordingly be incompressible. Goodman $\\&$ Lackner (2009) obtained a wider class of models in hydrostatic equilibrium under a fixed quadratic gravitational potential. Because the potential is fixed independently of the mass distribution and there are no constraints on the equation of state, such models may be constructed for an arbitrary density distribution. Now let use consider the integral \\begin{equation} Q_{fud}= 3\\int dV \\left (\\varpi^2 {\\rho \\over c_{s}^{2}}\\omega_0^2W_{k}\\right). \\label{eqn84b} \\end{equation} For the fiducial models described above this is identical to the overlap integral (\\ref{eqn84a}) where we note that we may adopt natural units such that the constant $\\omega_0^2, $ which should be identified with the surface value of $GM(r)/r^3$ is equal to unity in that case. More generally the integrand in (\\ref{eqn84b}) can be transformed using equation of hydrostatic equilibrium (\\ref{Good1}) such that \\begin{equation} Q_{fud}= -3\\int dV \\left ({\\varpi^2\\over r} {d\\rho \\over dr}W_{k}\\right)=-3\\int dV \\left (\\varpi {\\partial \\rho \\over \\partial \\varpi}W_{k}\\right). \\label{eqn84c} \\end{equation} Now let us consider equation (\\ref{eq 41}) for free normal modes in the anelastic approximation by setting $\\sigma =\\sigma_k$ and the right hand side of this equation to zero. Then we multiply it by $\\varpi^2$, set $m=2$, and integrate over $dV$. After removing derivatives of $W_k$ by integrating by parts, assuming that the density vanishes at the surface boundary, it is easy to see that it follows from (\\ref{eq 41}) that $Q_{fud}=0$ in the anelastic approximation provided $\\sigma_k \\ne 2\\Omega.$ This means, in particular, that $m=2$ inertial waves cannot be excited in the Goodman $\\&$ Lackner (2009) models in this approximation as was found by these authors when compressibility was fully taken into account (see PIN for additional discussion). Using the fact that $Q_{fud}=0$ we may rewrite (\\ref{eqn84a}) for models under their own self-gravity quite generally, adopting natural units, as \\begin{equation} Q_{k}=Q_{k}-Q_{fud}= 3\\int dV \\varpi^2 {\\rho \\over c_{s}^{2}}\\left(1-\\frac{M(r)}{r^3}\\right)W_{k}. \\label{eqn84d} \\end{equation} Taking into account that the factor in the brackets is proportional to $x$ for small $x$, and, accordingly, in this limit the integrand is proportional to $nx^n$ we see that now the logarithmic divergence of $Q_k/n$ disappears and, therefore, it is clear from the representation (\\ref{eqn84d}) that the overlap integral indeed smoothly tends to zero in the limit $n\\rightarrow 0$. The theory of asymptotic evaluation of integrals of the form (\\ref{eqn84d}) tells that the values of such integrals are determined either by inner stationary points, where gradient of the WKBJ phase vanishes or contributions close to the surface or other parts of the integration domain, where the WKBJ approximation is not valid. From the expression of the $W_{k}$ in the WKBJ regime (\\ref{eqn30}) it follows that there are no stationary points in the inner region of the planet. Considering the regions close to the surface it appears to be reasonable to assume that the leading contribution is determined by the region close to the critical latitude, where a 'hot spot' is observed in distributions of $W_k$, see the previous section. In this region the quantities $\\delta_1=x_1-\\mu$ and $\\delta_2=\\mu-x_2$ are small. We can use them as new integration variables in (\\ref{eqn84d}) with help of (\\ref{eqn66}), separate the contribution of this region to the integral by introduction of the functions $\\eta[\\delta_{1,2}]$ defined in section 3.7 in the integrand, and decompose the quantities in front of $W_k$ in powers of $\\delta_{1,2}$ taking into account that $x\\propto \\delta_{1}\\delta_2$ and $\\varpi^{2}\\approx (1-\\mu^{2})$ in the leading order. Assuming that $W_{k}\\approx \\bar W(\\delta_{1}) \\bar W(\\delta_{2})$, where $\\bar W(y)$ is given by equation (\\ref{eqn60}), it is easy to see that the leading contribution to (\\ref{eqn84}) is given by a symmetric combination of two integrals involving Bessel functions \\begin{equation} Q_{k}\\propto I_{1}(\\delta_{1})I_{2}(\\delta_{2})+ I_{2}(\\delta_{1})I_{1}(\\delta_{2}), \\label{eqn84g} \\end{equation} where \\begin{equation} I_{1}(y)=\\int_{y=0}dy \\eta[y]y^{n}J_{(n-1)/2}(\\kappa y), \\quad I_{2}(y)=\\int_{y=0}dy \\eta[y]y^{(n+1)}J_{(n-1)/2}(\\kappa y), \\label{eqn84j} \\end{equation} where it is assumed that $\\eta[y >y_{*}]=0$, $y_{*}$ lies within the range of integration and $1/\\kappa \\ll y_{*} \\ll 1$. As was shown by Larichev (1973) the integral $I_{1}(y)=0$ for any particular form of the function $\\eta[y]$. Thus, the leading order contribution to the overlap integral from the surface region close to the critical latitude is equal to zero. In principal, one can look for the next order terms. However, in this case our simple approach to the problem seems to be inadequate since eg. the assumption that $W_{k}$ can be represented as a product of two functions separately depending on the coordinates may be broken at this level, etc.. A more accurate approach is left for a future work. We note, however, that this cancellation means that the overlap integral should decay rapidly with increasing $\\lambda,$ possibly being inversely proportional to a large power of $\\lambda.$ This may qualitatively explain why a small number of relatively large scale modes are significantly excited by dynamic tides, see PI, IP and PIN.} \\subsection{Conclusions}\\label{Conclu} In this paper we have developed a WKBJ approximation, together with a formal first order perturbation approach for calculating the normal modes of a uniformly rotating coreless planet under the assumption of a spherically symmetric structure. Matching of the general WKBJ form valid in the interior to separable solutions valid near the surface resulted in expressions for eigenfunctions that were valid at any location within the planet together with an expression for the associated eigenfrequencies given in section \\ref{sec3.5}. Corrections as a result of density gradient terms neglected in the initial WKBJ approach were also obtained from formal first order perturbation theory. The corrected WKBJ eigenfrequencies obtained using the WKBJ eigenfunctions were compared with results obtained numerically by several different authors and found to be in good agreement, away from the limits of the inertial mode spectrum where identifications could be made, even for modes with a global structure. We also compared the spatial forms of the eigenfunctions with those obtained using the spectral method described in IP finding similar good agreement. This is consistent with the idea that these global modes can be identified and that first order perturbation theory works even though they are embedded in a dense spectrum. In further support of this, the formal first order perturbation theory developed here is subsequently used to estimate corrections to the eigenfrequencies as a consequence of the anelastic approximation and is then compared with simulation results for a polytropic model with $n=1$ in PIN. These different approaches are found to be in agreement for small enough rotation frequencies and also indicates that, as implied by the simplified discussion in section \\ref{anelastic} of this paper, that corrections as a result of the anelastic approximation are never very significant for the models adopted. Our results show that the problem of finding eigenfrequencies and eigenvalues of inertial modes allows for an approximate analytical treatment, even in the case of modes having a large scale distribution of perturbed quantities. Although we consider only the case of a polytropic planet, our formalism can be applied to a much wider context. Indeed, the approach developed here is mainly determined by the form of the density close to planet's surface, where we assume that it is proportional to a power of distance from the surface. Thus, we expect that our main results remain unchanged for any density distribution, which is approximately power-law close to the surface. In particular, according to our results, all models of type having approximately the same behaviour of the density distribution close to the surface should have approximately the same eigenspectrum. The formalism developed here can be extended for an approximate analytic evaluation of different quantities associated with inertial modes, such as overlap integrals characterising interaction of inertial waves with different physical fields, growth rates due to the CFS instability and decay rates due to various viscous interactions and non-linear mode-mode interaction. It may provide a basis for a perturbative analytic analysis of more complicated models, such as realistic models of star and planets flattened by rotation or models of relativistic stars. As we discussed above, for a given value of WKBJ order, $l,$ some modes are identified with modes obtained numerically while others remain unidentified. Eigenfrequencies of the unidentified modes are always either situated close to the boundaries of the frequency range allowed for inertial modes, $\\sigma =\\pm 2\\Omega $ or situated close to the origin $\\sigma=0$. We believe that our theory is not applicable to these modes, and they develop a small scale contribution controlled by the closeness of the positions of their eigenfrequencies to $\\sigma=0$ and $\\pm 2\\Omega, $ and thus effectively move to higher order than allowed for. Accordingly we de not consider these modes when comparing our results with results of direct numerical calculations of the excitation of inertial waves due to a tidal encounter reported in PIN." }, "1005/1005.5267_arXiv.txt": { "abstract": "{A crucial issue in star formation is to understand the physical mechanism by which mass is accreted onto and ejected by a young star. To derive key constraints on the launching point of the jets and on the geometry of the winds, the visible spectro-polarimeter VEGA installed on the CHARA optical array can be an efficient means of probing the structure and the kinematics of the hot circumstellar gas at sub-AU.} {For the first time, we observed the Herbig~Ae star AB Aur in the H$\\alpha$ emission line, using the VEGA low spectral resolution (R=1700) on two baselines of the array.} {% We computed and calibrated the spectral visibilities of AB~Aur between 610~nm and 700~nm in spectral bands of 20.4~nm. To simultaneously reproduce the line profile and the inferred visibility around H$\\alpha$, we used a 1-D radiative transfer code (RAMIDUS/PROFILER) that calculates level populations for hydrogen atoms in a spherical geometry and that produces synthetic spectro-interferometric observables.} {We clearly resolved AB Aur in the H$\\alpha$ line and in a part of the continuum, even at the smallest baseline of 34~m. The small P-Cygni absorption feature is indicative of an outflow but could not be explained by a spherical stellar wind model. Instead, it favors a magneto-centrifugal X-disk or disk-wind geometry. The fit of the spectral visibilities from 610 to 700~nm could not be accounted for by a wind alone, so another component inducing a visibility modulation around H$\\alpha$ needed to be considered. We thus considered a brightness asymmetry possibly caused by large-scale nebulosity or by the known spiral structures. } {Thanks to the unique capabilities of VEGA, we managed to simultaneously record for the first time a spectrum at a resolution of 1700 and spectral visibilities in the visible range on a target as faint as $m_{V}$ = 7.1. It was possible to rule out a spherical geometry for the wind of AB~Aur and provide realistic solutions to account for the H$\\alpha$ emission compatible with magneto-centrifugal acceleration. It was difficult, however, to determine the exact morphology of the wind because of the surrounding asymmetric ne\\-bulosity. The study illustrates the advantages of optical interferometry and motivates observations of other bright young stars in the same way to shed light on the accretion/ejection processes.} ", "introduction": "The class of Herbig Ae/Be (HAEBE) stars gathers the pre-main-sequence stars of intermediate mass (1.5 $\\leq M/M_{\\sun} \\leq$ 10), with spectral types of B to F8, strong emission lines, and the presence of infrared to submillimeter excess flux. Like their solar-type analogs, the T Tauri objects, HAeBe objects are known to be surrounded by protoplanetary disks of gas and dust, responsible for this excess emission. In the case of lower mass stars, the formation is commonly explained by the gravitational collapse of a dusty disk cloud and magnetically controlled accretion via a disk. The formation of more massive stars is still very uncertain. However, determining if similar processes are at work in these objects is extremely important, as it would help to esta\\-blish whether the accretion/ejection mechanism is universal irrespective of the mass of the central body. % The hot gas that can be involved in accretion and ejection flows close to the source and can be used to probe the corresponding physical conditions through its emission in lines, especially those in the optical range that have a high diagnostic potential. These phenomena, however, occur in a small region of a few astronomical units (AU) around the star, corresponding, for close star formation regions, to a few milliseconds of arc. Such a small scale can only be resolved with interferometry. Guided by these considerations, we have attempted for the first time to observe a bright visible Herbig Ae/Be star, AB~Aurigae, with a spectro-interferometer operating at optical wavelengths, the VEGA spectrograph (\\cite{vega}) installed at the CHARA Array (\\cite{chara}). This unique combination of spectral and spatially resolved information allowed us to study the physical conditions for the gas producing the H$\\alpha$ line on sub-AU scales. Interpreting the results, however, required appropriate modeling of the emitted radiation. Using a radiative transfer code, we obtained observational insights on the generation mechanism of winds around AB~Aurigae, giving experimental support to the widely accepted, but not yet tested, theory of magneto-centrifugal acce\\-leration. In Sect.~2, we recall the present knowledge about the target. In Sect.~3, we describe the observations and the data processing. In Sect.~4, we present the radiative transfer code and discuss the wind models adopted to interpret the observed H$\\alpha$ visibilities and spectrum. ", "conclusions": "\\begin{figure*}[t] \\centering \\includegraphics[width=9.1cm, angle=0]{14720fg3.pdf} \\includegraphics[width=9.1cm, angle=0]{14720fg4.pdf} \\caption{Best model predictions compared to the VEGA spectrum (left) and spectral visibilities (right), both in full circles with error bars. Dashed lines correspond to the X-wind model and full lines to the disk-wind one. For the squared visibilities, triangles denote the model prediction including both the wind and the contribution of the asymmetric reflection nebula.} \\label{fig:model-disc-binary} \\end{figure*} The spectroscopic P-Cygni feature in AB~Aur shows that the line is formed at least in part in a wind. Therefore, we focused on a wind model to constrain the properties of the emitting region. To simultaneously reproduce both the line profile and the spectral visibilities, we have adapted two numerical codes in spherical symmetry for the production and transport of hydrogen (H) lines (\\cite{Rajabi}). The first one, RAMIDUS, calculates the level populations of H in a moving fluid for assigned radial dependencies of the basic wind parameters (density, velocity, tempe\\-rature). The code uses the Sobolev approximation and the escape probability method (\\cite{Castor70}). In cascade, the PROFILER code calculates synthetic images, line profiles and visibilities of H lines for various geometrical configurations. By selecting limited portions of the spherical winds, we approximated 2-D axisymmetric models. We did not include accretion in the models. A full spherical wind providing matter along the line of sight can be readily excluded, because it would produce too much blue-shifted absorption with respect to the observed P-Cygni feature that is shallow but not very deep (normalized flux$\\geq$0.88). Even if the issue of the inclination angle of the AB Aur system is still alive (see Sect. 2), we chose a wind geometry in agreement with a small inclination (i$\\leq40^{\\circ}$), and thus excluded a double cone geometry, reminiscent of a bipolar jet, which would lead, for small inclinations, to a line profile with two distinct emission features instead of the observed single peak. Inspired by the magneto-centrifugal (MC) scenario for the acceleration of stellar jets, we assume that, on the observed scales (much less than 2-3 AU from the star/disk), the wind occupies a region delimited by a flattened torus defined by the pale gray region in Fig.~\\ref{fig:model}. This configuration can simulate the base of either a X-wind (\\cite{Camenzind90}, \\cite{Shu95}) or a disk-wind (hereafter D-wind; \\cite{Ferreira97}, \\cite{KoniglPudritz}) accelerated along the magnetic field lines anchored in the inner few AU of the accretion disk. In this region the wind has a wide angle aperture, as it has not yet reached the Alfv\\`en surface, after which magnetic collimation occurs and the wind collimates into a jet. In Fig.~\\ref{fig:model}, R$_i$=0.02~AU is the corotation radius, R$_d$=0.24~AU is the inner radius of the optically thick dusty disk, R$_f$=2.3~AU is the extension of the wind, and $\\alpha $ is the angle between the disk mid-plane and the surface of the wind, constrained to be $\\leq 60^{\\circ}$ for MC acceleration to occur (\\cite{Spruit96}). In our X-wind model, all the material in the flow comes from the region located very close to the corotation radius and has a constant mass outflow rate $\\dot{\\rm{M}}_w$ through the torus section, while for the D-wind $\\dot{\\rm{M}}_w$ progressively increases from R$_i$ to about 1~AU, to simulate that the wind originates in an extended region of the disk. The wind velocity is assumed to be $V(r)=V_{0}+V_{f}(1-(R_{\\star}/r)^2)$, where $V_{0} =$ 20 km.s$^{-1}$ is the initial velocity, $r$ the radial distance from the star, $R_{\\star}$ the stellar radius, and $V_{f}$ the terminal velocity. The density profile is dictated by $V(r)$ and $\\dot{\\rm{M}}_w (r)$ through the mass continuity equation, while the gas temperature profile T$(r)$, which reaches a maximum of 12~000 K, is chosen in accordance with the theory of wind heating illustrated by Shang et al. (1998) and Garcia et al. (2001) for the X-wind and D-wind, respectively. Modeling of the Stark absorption in the extended line wings has been included using a template spectrum from Kurucz~(1979). % Regarding the modeling of the line profile, both an X-wind and a D-wind could provide a good fit (Fig.~\\ref{fig:model-disc-binary}-left) for the model parameters collected in Table~\\ref{tab:model}. To reproduce at the same time the line width and the absence of a strong blue-shifted absorption, the best predictions for the X-wind have nearly a null inclination ($i$) and a large opening angle ($\\alpha$) of 60$^{\\circ}$. For the D-wind model, since the disk hides part of the high velocity material in the red-shifted wing, the best predictions are obtained with a small inclination (20$^{\\circ}$) and a smaller opening angle of 35$^{\\circ}$ to avoid too much blue-shifted absorption. Both models lead to a small inclination in agreement with the previous interferometric determinations. As pointed out by Pogodin (1992), a change in the opening angle of the wind ($\\alpha$) affects the depth of the blue-shifted absorption. Concerning the mass loss rate $\\dot{\\rm{M}}_w$, our best prediction is of the same order as previous determinations from line profile analysis (\\cite{CatalaKunasz}) for the X-wind but is significantly higher for the D-wind. In fact, our various simulations clearly show that the estimate of $\\dot{\\rm{M}}_{\\rm w}$ is very dependent on the velocity law and the temperature profile considered in the wind model, and on the scale on which the wind is accelerated. Since the geometry of our D-wind model strongly deviates from a spherical wind, the value of $\\dot{\\rm{M}}_{\\rm w}$ obtained for this model cannot be directly compared to the previous determinations of $\\dot{\\rm{M}}_w$. \\begin{table}[t] \\centering \\caption{Parameters of the models, with N$_{\\rm{H},ej}$ the average hydrogen density at the wind footpoints. } % \\label{tab:model} \\begin{tabular}{c c c c c c c} \\hline Model & $i$ & $\\alpha$ & V$_{\\rm{f}}$ [km/s] & $\\dot{\\rm{M}}_w$ [M$_{\\odot}$/yr]& N$_{\\rm{H},ej}$ [cm$^{-3}$]\\\\ \\hline X-wind & 0$^\\circ$ & 60$^\\circ$ & 410 & 3.7 10$^{-8}$ & 4.3~$10^{10}$ \\\\ D-wind & 20$^\\circ$ & 35$^\\circ$ & 450 & 2.2 10$^{-7}$ & 7.7~$10^{8}$\\\\ \\hline \\end{tabular} \\end{table} \\begin{figure}[t] \\centering \\includegraphics[width=0.35\\textwidth]{14720fg5.pdf} \\caption{Geometry adopted for the empirical X-wind and disk-wind mo\\-dels described in Table~\\ref{tab:model}. R$_i$, R$_d$, and R$_f$ are the corotation radius, dust sublimation radius, and extension of the wind, respectively.} \\label{fig:model} \\end{figure} Regarding the modeling of V$^2$, the X-wind provides only a negligible decrease in H$\\alpha$ ($V^2$ = 0.98) with respect to the continuum (Fig.~\\ref{fig:model-disc-binary}-right). In this case, the emission is concentrated at R$_i$, while it is much more extended in the D-wind model that produces a larger decrease in $V^2$ (down to 0.89). In both cases, however, the observed visibility variation could not be modeled with the wind alone so we considered an additional component producing an extended continuum emission. This component could come from the extended reflection nebulosity around the system, already reported by several authors (\\cite{Gradyb, Fukagawa, MG06}). In practice, we used this secondary source to reproduce the modulation of the visibility in the continuum. Our wind models combined with this additional emission provided a good fit of the visibilities. For instance, Fig.~\\ref{fig:model-disc-binary}-right corresponds to a secondary source separated from the star by 38.2 mas (i.e. 5.5 AU) contributing to 7\\% of the total flux. Because of the large number of free parameters and the fact that we only have measurements along one baseline direction, we found many different configurations for the secondary emission that could fit, for both wind models, the observed visibilities (e.g. a much larger separation of 591.2 mas). In conclusion, thanks to the unique capabilities of VEGA/CHARA, we managed for the first time to record simultaneously a spectrum at a resolution of 1700 and a visibi\\-lity spectrum in the visible range on a target as faint as $m_{V}$ = 7.1. Despite the limited signal-to-noise ratio of our data, they allowed us to rule out a spherical wind and to make realistic predictions on the nature of the H$\\alpha$ emission region at sub-AU, and our data are compatible with a magneto-centrifugal mechanism for the production of the wind. It was difficult, however, to determine the exact morphology of the wind and disentangle the X-wind and the D-wind because of the extended nebu\\-losity. Moreover, all models neglect the contribution of mass accretion to the H-line emission, so self-consistent models of winds including this process are needed. Complementary observations with larger CHARA baselines are required to place more stringent constraints on the complex environment and the photosphere contribution of AB~Aur that can be resolved in the visi\\-ble with VEGA. We also expect to take advantage of the new instrumental facilities of CHARA to study other bright HAeBe and provide promising insights on the accretion/ejection mechanisms in intermediate-mass young stars." }, "1005/1005.2501_arXiv.txt": { "abstract": "We follow the approach of induced--matter theory for a five--dimensional ($5D$) vacuum Brans--Dicke theory and introduce induced--matter and induced potential in four dimensional ($4D$) hypersurfaces, and then employ a generalized FRW type solution. We confine ourselves to the scalar field and scale factors be functions of the cosmic time. This makes the induced potential, by its definition, vanishes, but the model is capable to expose variety of states for the universe. In general situations, in which the scale factor of the fifth dimension and scalar field are~not constants, the $5D$ equations, for any kind of geometry, admit a power--law relation between the scalar field and scale factor of the fifth dimension. Hence, the procedure exhibits that $5D$ vacuum FRW--like equations are equivalent, in general, to the corresponding $4D$ vacuum ones with the same spatial scale factor but a new scalar field and a new coupling constant, $\\tilde{\\omega}$. We show that the $5D$ vacuum FRW--like equations, or its equivalent $4D$ vacuum ones, admit accelerated solutions. For a constant scalar field, the equations reduce to the usual FRW equations with a typical radiation dominated universe. For this situation, we obtain dynamics of scale factors of the ordinary and extra dimensions for any kind of geometry without any \\emph{priori} assumption among them. For non--constant scalar fields and spatially flat geometries, solutions are found to be in the form of power--law and exponential ones. We also employ the weak energy condition for the induced--matter, that gives two constraints with negative or positive pressures. All types of solutions fulfill the weak energy condition in different ranges. The power--law solutions with either negative or positive pressures admit both decelerating and accelerating ones. Some solutions accept a shrinking extra dimension. By considering non--ghost scalar fields and appealing the recent observational measurements, the solutions are more restricted. We illustrate that the accelerating power--law solutions, which satisfy the weak energy condition and have non--ghost scalar fields, are compatible with the recent observations in ranges $-4/3<\\omega\\leq-1.3151$ for the coupling constant and $1.5208\\leq n<1.9583$ for dependence of the fifth dimension scale factor with the usual scale factor. These ranges also fulfill the condition $\\tilde{\\omega}>-3/2$ which prevents ghost scalar fields in the equivalent $4D$ vacuum Brans--Dicke equations. The results are presented in a few tables and figures. ", "introduction": "\\indent Attempts to geometrical unification of gravity with other interactions, using higher dimensions other than our conventional $4D$ space--time, began shortly after invention of the special relativity (\\textbf{SR}). Nordstr{\\o}m was the first who built a unified theory on the base of extra dimensions~\\cite{1}. Tight connection between SR and electrodynamics, namely the Lorentz transformation, led Kaluza~\\cite{2} and Klein~\\cite{3} to establish $5D$ versions of general relativity (\\textbf{GR}) in which electrodynamics rises from the extra fifth dimension. Since then, considerable amount of works have been focused on this idea either using different mechanism for compactification of extra dimension or generalizing it to non--compact scenarios (see e.g. Ref.~\\cite{OverduinWesson1997}) such as Brane--World theories~\\cite{4}, space--time--matter or induced--matter (\\textbf{IM}) theories~\\cite{5} and references therein. The latter theories are based on the Campbell--Magaard theorem which asserts that any analytical $N$--dimensional Riemannian manifold can locally be embedded in an $(N+1)$--dimensional Ricci--flat Riemannian manifold~\\cite{6}. This theorem is of great importance for establishing $4D$ field equations with matter sources locally to be embedded in $5D$ field equations without \\emph{priori} introducing matter sources. Indeed, the matter sources of $4D$ space--times can be viewed as a manifestation of extra dimensions. This is actually the core of IM theory which employs GR as the underlying theory. On the other hand, Jordan~\\cite{9} attempted to embed a curved $4D$ space--time in a flat $5D$ space--time and introduced a new kind of gravitational theory, known as the scalar--tensor theory. Following his idea, Brans and Dicke~\\cite{10} invented an attractive version of the scalar--tensor theory, an alternative to GR, in which the weak equivalence principle is saved and a non--minimally scalar field couples to curvature. The advantage of this theory is that it is more Machian than GR, though mismatching with the solar system observations is claimed as its weakness~\\cite{Bertotti2003fujiBook}. However, the solar system constraint is a generic difficulty in the context of the scalar--tensor theories~\\cite{banerjee2001Sen2001}, and it does~not necessarily denote that the evolution of the universe, at all scales, should be close to GR, in which there are some debates on its tests on cosmic scales~\\cite{bean2009}. Although it is sometimes desirable to have a higher dimensional energy--momentum tensor or a scalar field, for example in compactification of extra curved dimensions~\\cite{11}, but the most preference of higher dimensional theories is to obtain macroscopic $4D$ matter from pure geometry. In this approach, some features of a $5D$ vacuum Brans--Dicke (\\textbf{BD}) theory based on the idea of IM theory have recently been demonstrated~\\cite{12}, in where the role of GR as fundamental underlying theory has been replaced by the BD theory of gravitation. Actually, it has been shown that $5D$ vacuum BD equations, when reduced to four dimensions, lead to a modified version of the $4D$ Brans--Dicke theory which includes an induced potential. Whereas in the literature, in order to obtain accelerating universes, inclusion of such potentials has been considered in \\emph{priori} by hand. A few applications and a $D$--dimensional version of this approach have been performed~\\cite{Ponce1and2,12.5}. Though, in Refs.~\\cite{Ponce1and2}, it has also been claimed that their procedure provides explicit definitions for the effective matter and induced potential. Besides, some misleading statements and equations have been asserted in Ref.~\\cite{12}, and hence we have re--derived the procedure in Section $2$. Actually, the reduction procedure of a $5D$ analogue of the BD theory, with matter content, on every hypersurface orthogonal to an extra cyclic dimension (recovering a modified BD theory described by a 4--metric coupled to two scalar fields) has previously been performed in the literature~\\cite{qiang20052009}. However, the key point of IM theories are based on not introducing matter sources in $5D$ space--times. In addition, recent measurements of anisotropies in the microwave background suggest that our ordinary $4D$ universe should be spatially flat~\\cite{13}, and the observations of Type Ia--supernovas indicate that the universe is in an accelerating expansion phase~\\cite{14}. Hence, the universe should mainly be filled with a dark energy or a quintessence which makes it to expand with acceleration~\\cite{15}. Then after an intensive amount of work has been performed in the literature to explain the acceleration of the universe. In this work, we explore the Friedmann--Robertson--Walker (\\textbf{FRW}) type cosmology of a $5D$ vacuum BD theory and obtain solutions and related conditions. This model has extra terms, such as a scalar field and scale factor of fifth dimension, which make it capable to present accelerated universes beside decelerated ones. In the next section, we give a brief review of the induced modified BD theory from a $5D$ vacuum space--time to rederive the induced energy--momentum tensor, as has been introduced in Ref.~\\cite{12}, for our purpose to employ the energy density and pressure. In Section~$3$, we consider a generalized FRW metric in the $5D$ space--time and specify FRW cosmological equations and employ the weak energy condition (\\textbf{WEC}) to obtain the energy density and pressure conditions. Then, we probe two special cases of a constant scale factor of the fifth dimension and a constant scalar field. In Section~$4$, we proceed to exhibit that $5D$ vacuum BD equations, employing the generalized FRW metric, are equivalent, in general, to the corresponding vacuum $4D$ ones. This equivalency can be viewed as the main point within this work which distinguishes it from Refs.~\\cite{12,Ponce1and2}. In Section~$5$, we find exact solutions for flat geometries and proceed to get solutions fulfilling the WEC while being compatible with the recent observational measurements. We also provide a few tables and figures for a better view of acceptable range of parameters. Finally, conclusions are presented in the last section. ", "conclusions": "\\indent Analogous to the approach of IM theories, one can consider the BD gravity as the underlying theory. Hence, extra geometrical terms, coming from the fifth dimension, are regarded as an induced--matter and induced potential. We have followed, with some corrections, the procedure of Ref.~\\cite{12} for introducing the induced potential and have employed a generalized FRW type solution for a $5D$ vacuum BD theory. Hence, the scalar field and scale factors of the $5D$ metric can, in general, be functions of the cosmic time and the extra dimension. However, for simplicity, we have assumed the scalar field and scale factors to be only functions of the cosmic time, where this makes the induced potential, by its definition, vanishes. We then have revealed that in general situations, in which the scale factor of the fifth dimension and scalar field are~not constants, the $5D$ equations, for any kind of geometry, admit a power--law relation between the scalar field and scale factor of the fifth dimension. Hence, the procedure exhibits that $5D$ vacuum FRW--like equations are equivalent, in general, to the corresponding $4D$ vacuum ones with the same spatial scale factor but a new (or modified) scalar field and a new coupling constant. This equivalency can be viewed as the distinguished point of this work from Refs.~\\cite{12,Ponce1and2}. Indeed, through investigating the $5D$ vacuum FRW--like equations, we have shown that its equivalent $4D$ vacuum equations admit accelerated scale factors, contrary to what one may have expected from a vacuum space--time. Conclusions of the complete investigation of the induced $4D$ equations are as follows. Following our investigations for cosmological implications, we have shown that for the special case of a constant scale factor of the fifth dimension, the $5D$ vacuum FRW--like equations reduce to the corresponding equations of the usual $4D$ vacuum BD theory, as expected. In the special case of a constant scalar field, the action reduces to a $5D$ Einstein gravitational theory and the equations reduce to the usual FRW equations with a typical radiation dominated universe. For this situation, we also have obtained dynamics of scale factors of the ordinary and extra dimensions for any kind of geometry without any \\emph{priori} assumption among them. Solutions predict a limited life time for closed geometries and unlimited one for flat and open geometries. A typical time evolutions of scale factors correspond to closed, flat and open geometries have been illustrated in Fig.~$0$. Then, we have focused on spatially flat geometries and have obtained exact solutions of scale factors and scalar field. Solutions are found to be in the form of power--law and exponential ones in the cosmic time. We also have employed the WEC for the induced--matter of the $4D$ modified BD gravity, that gives two conditions (\\ref{49.1}) and (\\ref{49.2}). We then have pursued properties of these solutions and have indicated mathematically and physically acceptable ranges of them, and the results have been presented in a few tables and figures. All types of solutions fulfill the WECs in different ranges, where the exponential solutions are more restricted. The solutions fulfilling the WEC (\\ref{49.1}) have negative pressures, but the figures illustrate that for the power--law results there are decelerating solutions beside accelerating ones. For this condition, both $\\rho_{_{\\rm BD}}$ and $|p_{_{\\rm BD}}|$ decrease with the cosmic time, but the extra dimension grows. On the other hand, the solutions satisfying the WEC (\\ref{49.2}) have positive pressures, where the power--law results accept accelerating solutions in addition to decelerating ones. For this condition, again decreasing energy density and pressure with the time can occur for some solutions, however all with shrinking extra dimension. The homogeneity between the extra dimension and the usual spatial dimensions, i.e. $b\\propto a$, can take place in the solutions, but for the power--law ones the WECs exclude it. By considering non--ghost scalar fields and appealing the recent observational measurements, the solutions have been more restricted. Actually, we have illustrated that the accelerating power--law solutions, which satisfy the WEC and have non--ghost scalar fields, are compatible with the recent observations in ranges $-4/3<\\omega\\leq-1.3151$ for the BD coupling constant and $1.5208\\leq n<1.9583$ for dependence of the fifth dimension scale factor with the usual scale factor. These ranges also fulfill the condition $\\tilde{\\omega}>-3/2$ which prevents ghost scalar fields in the equivalent $4D$ vacuum BD equations. Incidentally, this range is more restricted than the one obtained in Ref.~\\cite{Ponce1and2}, i.e. $-1.5<\\omega<-1$, where the difference may have been caused by the distinct definition of the induced potential in two approaches of Ref.~\\cite{12} and Ref.~\\cite{Ponce1and2}. However, we should remind that it has also been shown~\\cite{Freund1982} that the WEC, for $5D$ space--times, requires $-4/3\\leq\\omega$, in which no other experimental evidences have been considered." }, "1005/1005.1502_arXiv.txt": { "abstract": "{Herschel PACS and SPIRE images have been obtained of NGC 6720 (the Ring Nebula). This is an evolved planetary nebula with a central star that is currently on the cooling track, due to which the outer parts of the nebula are recombining. From the PACS and SPIRE images we conclude that there is a striking resemblance between the dust distribution and the H$_2$ emission, which appears to be observational evidence that H$_2$ forms on grain surfaces. We have developed a photoionization model of the nebula with the Cloudy code which we used to determine the physical conditions of the dust and investigate possible formation scenarios for the H$_2$. We conclude that the most plausible scenario is that the H$_2$ resides in high density knots which were formed after the recombination of the gas started when the central star entered the cooling track. Hydrodynamical instabilities due to the unusually low temperature of the recombining gas are proposed as a mechanism for forming the knots. H$_2$ formation in the knots is expected to be substantial after the central star underwent a strong drop in luminosity about one to two thousand years ago, and may still be ongoing at this moment, depending on the density of the knots and the properties of the grains in the knots.} ", "introduction": "Grains play an important role in many environments, including planetary nebulae (PNe), because of extinction, photoelectric heating, their influence on the charge and ionization balance of the gas, as catalysts for grain-surface chemical reactions (e.g.\\ H$_2$ formation), and as seeds for freeze-out of molecules. Previous satellite missions such as IRAS, ISO, Spitzer, and AKARI have allowed us to study the dust in PNe, but unfortunately the angular resolution of these instruments was too low to get detailed information on the spatial distribution of the dust. This has now changed with the launch of Herschel, which allows us to study the spatial structures in unprecedented detail. In this paper we will do this for NGC 6720 (M57, the Ring nebula) to study H$_2$ formation. NGC 6720 is an evolved, oxygen-rich bipolar nebula seen nearly pole-on. The nebula is optically thick to ionizing radiation in most directions, but optically thin in the polar regions \\citep[hereafter OD07]{OD07}. Molecules such as H$_2$ and CO have been detected \\citep{Be78, Hu86}. The central star has exhausted hydrogen shell burning and is now on the cooling track. As a result the outer halo is recombining and re-ionization of the innermost recombined material due to expansion of the nebula has just started (OD07). This object is very similar to the Helix nebula, which seems to be further advanced along the same evolutionary path (OD07). In Sect.~\\ref{obs} we will describe the observations, and in Sect.~\\ref{origin} we will discuss various scenarios for the formation of H$_2$. ", "conclusions": "" }, "1005/1005.4782_arXiv.txt": { "abstract": "In this paper we present results of an in depth analysis of the High Mass X-ray Binary \\gx~ using \\ig~ data. The source has a double-peaked orbital lightcurve, with flares close to orbital phases 0.5 and 0.9. In addition to archival data three dedicated observations with a total exposure of 200~ksec were performed to observe the pre-periastron flare of the source when the flux from the source is maximal. We obtained broad-band (2-100 keV) pulse phase resolved spectra. We discuss the spectral models suitable for spectrum description and the impact of the model choice on CRSF energy pulse phase dependence. The history of the pulse period since the launch of \\ig~ is determined and the spin down trend, established by RXTE \\citep{Pravdo:2001p360,Kreykenbohm:2004p155} and BeppoSax \\citep{LaBarbera:2005p156} observations is confirmed. We update orbital ephemeris of the system published previously by \\cite{Koh:1997p138}, based on time arrival analysis of archival data. The updated value of orbital period 41.482(2) is found to be consistent with the period obtained from the RXTE ASM lightcurve of the source. ", "introduction": "% \\label{sec:introduction} \\gx~ (also known as 4U 1223-62) is a High Mass X-ray Binary system, consisting of a Neutron star orbiting an early B-type optical companion \\Wr~. The neutron star is a $\\sim$680s X-ray pulsar \\citep{White:1976p993}, accreting from a dense wind of the optical companion. The wind mass loss rate of the optical component is one of the highest known in the Galaxy ($\\dot{M}\\sim 10^{-5} M_{\\odot}\\cdot yr^{-1}$). Taking into account very low ($v_{0}\\sim 300~km\\cdot s^{-1}$, \\cite{Kaper:2006p1357}) terminal speed of the wind , the accretion rate shall be high enough to explain observed luminosity of order $L_x\\sim10^{37} ~erg\\cdot s^{-1}$. Distance to the source is estimated between 1.8~kpc \\citep{Parkes:1980p1360} and 5.3~kpc \\citep{Kaper:1995p151} depending on spectral classification of \\Wr, with latest estimates close to 3~kpc \\citep{Kaper:2006p1357}. We will adopt the later value through out the paper. The source exhibits regular X-ray flares just 1-2d before periastron passage (orbital phase $\\sim0.9$). There's also a sign of other flare at orbital phase $\\sim0.5$ \\citep{Koh:1997p138}. A number of hypotheses were proposed to explain observed orbital lightcurve, including circumstellar disk \\citep{Koh:1997p138} as well as presence of quasi-stable accretion stream \\citep{Leahy:2008p358}. In this paper we focus mainly on studying properties of the source during pre-periastron flare, there brightness of the source is highest, so spectral properties can be studied with high statistics. The spectrum of the source is quite complicated and feature rich. Lower energy part of spectra requires use of partial covering models for absorption with optional reflection component. An iron line of complicated shape is observed at $\\sim6.5$ keV \\citep{Watanabe:2003p357}. High energy cutoff at $\\sim20keV$ together with deep and broad cyclotron resonance scattering feature at $\\sim30-45keV$ are main features at higher energies. ", "conclusions": "\\label{sec:conclusions} \\pagebreak \\footnotesize" }, "1005/1005.4057_arXiv.txt": { "abstract": "Observations of the frequency dependence of the global brightness temperature of the redshifted 21 cm line of neutral hydrogen may be possible with single dipole experiments. In this paper, we develop a Fisher matrix formalism for calculating the sensitivity of such instruments to the 21 cm signal from reionization and the dark ages. We show that rapid reionization histories with duration $\\Delta z\\lesssim 2$ can be constrained, provided that local foregrounds can be well modelled by low order polynomials. It is then shown that observations in the range $\\nu=50-100{\\,\\rm MHz}$ can feasibly constrain the \\lya and X-ray emissivity of the first stars forming at $z\\sim15-25$, provided that systematic temperature residuals can be controlled to less than 1 mK. Finally, we demonstrate the difficulty of detecting the 21 cm signal from the dark ages before star formation. ", "introduction": "\\label{sec:intro} The transition of the Universe from the dark ages following hydrogen recombination through to the epoch of reionization remains one of the least constrained frontiers of modern cosmology. Observing the sources responsible for heating and ionizing the intergalactic medium (IGM) at redshifts $z\\gtrsim6$ pushes current observational techniques to the limit. Plans are underway to construct low-frequency radio telescopes, such as LOFAR\\footnote{http://www.lofar.org/}, MWA\\footnote{http://www.MWAtelescope.org/}, PAPER\\footnote{\\citet{parsons2009}}, and SKA\\footnote{http://www.skatelescope.org/}, to observe the red-shifted 21 cm line of neutral hydrogen. These experiments aim to map the state of the intergalactic medium via tomographic observations of 3D fluctuations in the 21 cm brightness temperature. A simpler and significantly lower cost alternative to this would be measurements of the global 21 cm signal integrated over the sky \\cite{mmr1997,shaver1999,sethi2005}, which can be achieved by single dipole experiments like EDGES \\cite{bowman2007edges} or CoRE \\cite{chippendale2005}. Although such experiments are today in their infancy, their potential is large. In this paper, we explore the potential for these global sky experiments to measure the 21 cm signal and constrain the high redshift Universe. We may draw a historical analogy with the Cosmic Background Explorer (COBE), whose FIRAS instrument measured the blackbody spectrum of the cosmic microwave background (CMB) \\cite{mather1994} while the DMR instrument measured the level of temperature fluctuations \\cite{smoot1992}. The precise measurement of a $T_{\\rm CMB}=2.726{\\,\\rm K}$ blackbody spectrum placed tight constraints on early energy injection, since no Compton-$y$ or $\\mu$-distortion were seen, and provided important evidence confirming the big bang paradigm. The detection of angular fluctuations paved the way for more sensitive experiments such as BOOMERANG \\cite{lange2001} and WMAP \\cite{spergel2003}, which provided precision measurements of the CMB acoustic peaks. While, at the moment, attention is focussed on experiments designed to measure 21 cm fluctuations, it is important not to neglect the possibility of measuring the global signal. The evolution of the 21 cm signal is driven primarily by the amount of neutral hydrogen and the coupling between the 21 cm spin temperature and the gas temperature. It is able to act as a sensitive thermometer when the IGM gas temperature is less than the CMB temperature placing constraints on energy injection that leads to heating. For example, the first black holes to form generate X-rays, which heat the gas. More exotic processes such as annihilating dark matter might have also been important. Additionally, energy injection in the form of \\lya production modifies the strength of the coupling. This provides a way of tracking star formation, which will be the dominant source of \\lya photons. As we show, the spectral structure of the 21 cm signal is much richer than that of a blackbody so that many things can be learnt about the early Universe. Given the uncertainties, we develop a model approach based upon those physical features most likely to be present. The single most important factor determining the sensitivity of dipoles to astrophysics will be their ability to remove galactic foregrounds \\cite[e.g.][]{oh2003,dimatteo2004}. Exploitation of spectral smoothness to remove foregrounds by fitting low order polynomials is key to avoiding throwing the signal away with the foreground. To quantify this, we develop a simple Fisher matrix formalism and validate it against more detailed numerical parameter fitting. This provides us with a way of quantitatively addressing the ability of global 21 cm experiments to constrain reionization and the astrophysics of the first galaxies \\cite{loeb2010book}. Similar work on the subject \\cite{sethi2005} ignored the influence of foregrounds limiting its utility considerably. Much of the power of this technique stems from the limitations of other observational probes. While next generation telescopes such as JWST\\footnote{http://www.jwst.nasa.gov/}, GMT\\footnote{http://www.gmto.org/}, EELT\\footnote{http://www.eso.org/sci/facilities/eelt/} or TMT\\footnote{http://www.tmt.org/} may provide a glimpse of the Universe at $z\\gtrsim12$ they peer through a narrow field of view and are unlikely to touch upon redshifts $z\\gtrsim20$. As we will show, 21 cm global experiments could potentially provide crude constraints on even higher redshifts at a much lower cost. The structure of this paper is as follows. In \\S\\ref{sec:physics}, we begin by describing the basic physics that drives the evolution of the 21 cm global signature and drawing attention to the key observable features. We follow this in \\S\\ref{sec:foreground} with a discussion of the foregrounds, which leads into our presenting a Fisher matrix formalism for predicting observational constraints in \\S\\ref{sec:fisher}. In \\S\\ref{sec:reion} and \\S\\ref{sec:astro} we apply this formalism to the signal from reionization and the first stars, respectively. After a brief discussion in \\S\\ref{sec:darkages} of the prospects for detecting the signal from the dark ages before star formation, we conclude in \\S\\ref{sec:conclude}. Throughout this paper where cosmological parameters are required we use the standard set of values $\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, $\\Omega_b=0.046$, $H=100h\\,\\rm{km\\,s^{-1}\\,Mpc^{-1}}$ (with $h=0.7$), $n_S=0.95$, and $\\sigma_8=0.8$, consistent with the latest measurements \\citep{komatsu2009}. ", "conclusions": "\\label{sec:conclude} Observations of the redshifted 21 cm line potentially provide a new window into the high redshift Universe. Detecting this signal in the presence of large foregrounds is challenging and it is important to explore all avenues for exploiting the signal. In this paper, we have focussed upon the possibility of using single dipole experiments to observe the all-sky 21 cm signal, in contrast to the 21 cm fluctuations targeted by MWA, LOFAR, PAPER, and SKA. Experiments targeting this global signal are in their infancy. We emphasise that instruments built from a few dipoles targeting the global 21 cm signal can be several orders of magnitude cheaper to build than interferometers targeting the fluctuations. Their scientific return will be similarly less, but at this stage where we know so little about the first sources, even that little is extremely valuable. As we have outlined in this paper, the 21 cm signal generated by astrophysical processes has a well defined form, although the input parameters are only poorly understood. We have demonstrated that, at the level of our current knowledge, describing the Galactic foregrounds requires at least a 3rd order polynomial. At this level, we are able to remove the foregrounds to the sub-mK level, although in practice this procedure may be more complicated. In order to characterise the sensitivity of these experiments to the signal, we developed a Fisher matrix formalism and validated it against more numerical fitting of the model parameters. This Fisher matrix approach allows rapid calculations of the experimental sensitivity and appears to reproduce more detailed calculations very well. Having developed this formalism we applied it to the signal from reionization and the epoch of the first stars. Using a toy model of reionization, we demonstrated that EDGES-like experiments should be capable of constraining rapid reionization histories with $\\Delta z\\lesssim2$. More promisingly, these experiments can rule out a wide variety of astrophysical models for the signal from the first stars where the evolution of the spin temperature is important. We used a straightforward fitting form for the signal based upon the positions of the turning points and showed that these features could be constrained, with the deepest absorption trough providing the best observational target. Finally, we briefly explored the possibility of detecting the absorption feature present before star formation began. The increased foreground brightness at low frequencies make it very difficult to constrain this feature and will require long integration times and more sophisticated methods of foreground removal. This paper represents a first serious look at the prospects for using global measurements of the 21 cm signal to constrain astrophysics. As a result, there are a number of places where future work might improve upon our calculations. These include investigating the effects of finite sky coverage, incorporating an arbitrary instrumental frequency response, and allowing for the removal of frequency channels corrupted by terrestrial radio interference." }, "1005/1005.0629.txt": { "abstract": "The Advanced Camera for Surveys (ACS) and the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) have been used to obtain new {\\em Hubble Space Telescope} images of NGC~4038/4039 (``The Antennae''). These new observations allow us to better differentiate compact star clusters from individual stars, based on both size and color. We use this ability to extend the cluster luminosity function by approximately two magnitudes over our previous WFPC2 results, and find that it continues as a single power law, $dN/dL \\propto L^{\\alpha}$ with $\\alpha = -2.13 \\pm 0.07$, down to the observational limit of $\\Mv\\approx -7$. Similarly, the mass function is a single power law $dN/dM \\propto M^{\\beta}$ with $\\beta = -2.10 \\pm 0.20$ for clusters with ages $<\\!3\\times10^8$~yr, corresponding to lower mass limits that range from $10^4$ to $10^5\\,\\msun$, depending on the age range of the subsample. Hence the power law indices for the luminosity and mass functions are essentially the same. The luminosity function for intermediate-age clusters (i.e., $\\sim$100--300 Myr old objects found in the loops, tails, and outer areas) shows no bend or turnover down to $\\Mv\\approx -6$, consistent with relaxation-driven cluster disruption models which predict the turnover should not be observed until $\\Mv \\approx -4$. An analysis of individual $\\sim$0.5-kpc sized areas over diverse environments shows good agreement between values of $\\alpha$ and $\\beta$, similar to the results for the total population of clusters in the system. There is tentative evidence that the values of both $\\alpha$ and $\\beta$ are flatter for the youngest clusters in some areas, but it is possible that this is caused by observational biases. Several of the areas studied show evidence for age gradients, with somewhat older clusters appearing to have triggered the formation of younger clusters. The area around Knot~B is a particularly interesting example, with an $\\sim$10--50~Myr old cluster of estimated mass $\\sim\\!10^6\\,\\Msun$ having apparently triggered the formation of several younger, more massive (up to $5\\times10^6\\,\\Msun$) clusters along a dust lane. A comparison with new NICMOS observations reveals that only $16 \\pm 6$\\% of the IR-bright clusters in the Antennae are still heavily obscured, with values of $\\Av > 3$ mag. ", "introduction": "The Antennae (NGC 4038/39) are the youngest and nearest example of a pair of merging disk galaxies in the Toomre (1977) sequence. As such, they have been observed in virtually every wavelength regime: radio (e.g., Wilson \\etal\\ 2000, 2003; Neff \\& Ulvestad 2000; Hibbard \\etal\\ 2001; Zhu, Seaquist \\& Kuno 2003; Schulz \\etal\\ 2007), infrared (e.g., Mirabel \\etal\\ 1998; Kassin \\etal\\ 2003; Wang \\etal\\ 2004; Brandl \\etal\\ 2005, 2009; Mengel \\etal\\ 2005; Clark \\etal\\ 2007; Rossa \\etal\\ 2007; Gilbert \\& Graham 2007; Mengel \\etal\\ 2008), optical (Rubin, Ford, \\& D'Odorico 1970; Whitmore \\& Schweizer 1995; Whitmore \\etal\\ 1999; Kassin \\etal\\ 2003; Bastian \\etal\\ 2006), ultraviolet (e.g., Hibbard \\etal\\ 2005), and X-ray (e.g., Fabbiano \\etal\\ 2003, 2004; Metz \\etal\\ 2004; Zezas \\etal\\ 2006, 2007). One important result has been the discovery that much of the star formation occurs in the form of massive compact star clusters (Whitmore \\& Schweizer 1995; Whitmore \\etal\\ 1999), often referred to as ``super star clusters.'' Some of these clusters have properties expected of young globular clusters (i.e., mass, effective radius, etc.). Hence, it may be possible to study the formation (and disruption, see Fall, Chandar, \\& Whitmore 2005, hereafter FCW05; Whitmore, Chandar, \\& Fall 2007, hereafter WCF07; and Fall, Chandar, \\& Whitmore 2009, hereafter FCW09) of globular clusters in the local universe rather than trying to determine how they formed some 13 Gyr ago. One of the primary challenges for previous observations of the Antennae was the difficulty of distinguishing faint young clusters from individual young stars, which can be as luminous as $\\Mv \\approx -9$. The Advanced Camera for Surveys, with its superior spatial resolution over a wider field of view than the Wide Field and Planetary Camera 2 (WFPC2), provides an opportunity to improve this situation. The ability to better distinguish faint clusters from stars provides several new science opportunities. The first is an extension of the cluster luminosity and mass functions to lower values. A large number of studies have found that the young cluster luminosity function (LF) in a variety of star-forming galaxies is a power law of the form $\\phi(L)dL \\propto L^{\\alpha} dL$ with a value of $\\alpha\\approx -2$ (e.g., see compilations in Whitmore 2003; Larsen 2005, 2006; de Grijs \\& Parmentier 2007). However, recently there have been claims (e.g., Fritze-von Alvensleben 2004; de Grijs \\& Parmentier 2007; Anders \\etal\\ 2007) that the initial cluster luminosity function has a turnover at faint magnitudes, and is closer to the Gaussian form found for old globular clusters. The ability to push the LF about 2--3 mag fainter provides a more stringent test of the power-law nature of the initial cluster LF. Areas in the outskirts of the galaxy provide an especially sensitive test, since their low background levels allow cluster detections to fainter magnitudes. Most of the clusters in these tails, loops, and outer areas have intermediate ages (approximately 100--300 Myr) and have, therefore, survived the initial stages of cluster disruption or ``infant mortality'' associated with the removal of interstellar material (see Hills 1980; Lada, Margulis \\& Dearborn 1984; Boily \\& Kroupa 2003; Bastian \\& Goodwin 2006; and Fall \\etal\\ 2010 for theoretical discussions and Lada \\& Lada 2003, FCW05, and WCF07 for observational evidence). Several studies have also found that the young cluster mass function (MF) in a variety of galaxies is a power law of the form $\\phi(M)dM \\propto M^{\\beta} dM$ with a value of $\\beta\\approx -2$ (e.g., Zhang \\& Fall 1999 in the Antennae; Hunter \\etal\\ 2003 in the Magellanic Clouds; Bik \\etal\\ 2003 in M51; and Chandar \\etal\\ 2010 in M83). While the LF and MF are sometimes considered to be interchangeable, this is not true for populations of clusters that include a wide range of ages, and hence mass-to-light ratios. In fact, the observed similarity in the values of $\\alpha$ and $\\beta$ can be used as indirect evidence that the age and mass distributions are independent, at least up to ages of 300 Myr (e.g., Fall 2006). This independence has been established more directly in FCW05, and can be expressed explicitly by the fact that the bivariate mass-age distribution $g(M,\\tau)$ can be approximated by the product of the univariate mass and age distribution: $g(M,\\tau)\\approx M^{\\beta} \\tau^{\\gamma}$ (FCW05, WCF07, FCW09). There have also been claims for several galaxies, including the Antennae, that there is a cutoff to the upper mass with which clusters can form (e.g., Gieles \\etal\\ 2006; Bastian 2008; Larsen 2009), and that there are bends in the age distributions of some galaxies (e.g., Boutloukos \\& Lamers 2003, and Lamers \\etal\\ 2005). However, based on our WFPC2 data, we found that both the mass function and the age distribution are simple power laws (see FCW09 for discussion). We suspect that these apparent discrepancies are due to selection effects in other studies, in particular to the exclusion of crowded regions which contain most of the young clusters. In this paper we examine whether or not there is evidence for such bends or cutoffs from our new ACS data set. A second new science opportunity enabled by the high-resolution ACS images is to dissect the contents, both clusters and individual stars, in some of the spectacular star forming knots found in the Antennae. We will be able to assess whether some of the universal trends found for star-forming galaxies, for example the correlation between the absolute visual magnitude of the brightest cluster and the total number of clusters (e.g., Whitmore 2003), holds or not in individual sub-kpc regions of the Antennae. Age and mass estimates for the clusters in these regions will allow us to assess how star formation proceeds in these extreme environments, and whether there is evidence for sequential star formation. In this paper, we adopt a distance modulus of 31.71 mag for the Antennae, as found by Schweizer \\etal\\ (2008) based on the Type Ia supernova 2007sr. This modulus corresponds to a distance of 22~Mpc for a Hubble constant of $H_0 = 72$ km s$^{-1}$ Mpc$^{-1}$. Note that this distance is slightly further than that adopted in our earlier papers (19.2 Mpc), and quite different from the shorter distance of 13.3 Mpc found by Saviane \\etal\\ (2008), based on an apparent misidentification of the tip of the red giant branch (see Schweizer \\etal\\ 2008 for details). At a distance of 22 Mpc, 1\\arcsec\\ is equivalent to 107 pc, and one $0\\farcs05$ ACS/WFC pixel covers 5.33 pc. The remainder of this paper is organized as follows: \\S2 describes the observations and photometric reductions; \\S3 gives details of our procedure to estimate the ages and masses of the clusters; and \\S4 summarizes our method for separating stars and clusters (with more details included in Appendix~A). We compare the 50 most luminous, massive, and IR-bright clusters in \\S5; examine the LF of star clusters in the Antennae in \\S6, and then the mass and age distributions in \\S7. In \\S8 we perform a detailed analysis of the cluster populations in star-forming knots and other regions, and \\S9 discusses patterns of star formation, including triggered star formation. A summary is provided in \\S10. Future papers will include a more detailed study of the age, mass, and size distributions of the clusters (Chandar \\etal, in prep.) and will incorporate our NICMOS observations of the Antennae more fully (Rothberg \\etal, in prep). ", "conclusions": "The Advanced Camera for Surveys on {\\em HST\\/} was used to obtain high-resolution images of NGC~4038/4039, which allow us to better differentiate between individual stars and compact star clusters. We have used this improved ability to extend the cluster LF by approximately 2--3 mag at its faint end, when compared to the previous WFPC2 photometry. In addition, the Near Infrared Camera and Multi-Object Spectrometer (NICMOS) was used to help study the clusters still largely embedded in dust. Following are our main results: 1. The cluster luminosity function continues as a single power law, $dN/dL \\propto L^{\\alpha}$, with $\\alpha=-2.13 \\pm 0.07$ down to the current observational limit of $\\Mv \\approx -7$. This value is uncorrected for extinction and based on variable-binning fits. Similarly, the cluster mass function is a single power law $dN/dM \\propto M^{\\beta}$ with $\\beta = -2.10 \\pm 0.20 $ for clusters with ages $<\\!3\\times10^8$~yr, corresponding to lower mass limits that range from $10^4$ to $10^5\\,\\msun$, depending on the age range of the subsample. Hence the power-law indices for the luminosity and mass functions are essentially the same. 2. The LF for intermediate-age clusters (e.g., $\\sim$200 Myr old clusters found in the loops, tails, and outer areas) does not show any hint of a turnover down to $\\Mv \\approx -6$. This is consistent with relaxation-driven cluster disruption models, which predict the turnover should not be observed until $\\Mv \\approx -4$ for this age range. 3. The brightest individual stars in the Antennae reach absolute visual magnitudes of $\\Mv \\approx -9.5$, a limit compatible with the so-called Humphreys--Davidson (1979) limit observed in the Milky Way and nearby galaxies (see Appendix A for details). 4. Our photometrically determined cluster ages are in good agreement with spectroscopically determined ages from Bastian \\etal\\ (2009) for 13 of the 15 clusters in common. In the other two cases, spatial resolution effects are likely to be responsible for the age differences, since the ground-based spectroscopic observations contain light from a much larger area than our \\textit{HST} measurements. 5. An area-by-area analysis shows a strong correlation between the absolute magnitude of an area's brightest cluster, \\Mv(brightest), and the (logarithmic) number of clusters brighter than $\\Mv = -9$ in that area, $\\log N$. The \\Mv(brightest) vs.\\ $\\log N$ diagram for all areas of the Antennae studied is essentially identical to the corresponding diagram for entire galaxies, supporting the universality of this observed trend on a scale of a few hundred pc and larger. In addition, the reddening vs.\\ age correlation found for individual clusters by Whitmore \\& Zhang (2002), Mengel \\etal\\ (2005), and Bastian \\etal\\ (2005) has been confirmed for the area-by-area analysis as well. 6. The same area-by-area analysis also reveals a tentative discovery of a new relationship between the power-law exponent $\\alpha$ of the LF and the median age of clusters in each area, in the sense that younger clusters have flatter LFs (i.e., values of $\\alpha\\/ \\approx -1.8$ for the youngest star- and cluster-forming knots and $-$2.1 for the outer areas). However, it is possible that this is caused by observational biases. The power-law indices for the LFs and MFs are in good agreement, and are similar to the results for the total population of clusters in the galaxy. 7. Using our age and mass estimates of clusters in the Antennae, we have found evidence of sequential cluster formation in several regions. Often there is a massive, somewhat older cluster which appears to have triggered the formation of younger, lower-mass satellite clusters. The area around Knot~B is a particularly dramatic example of triggered star cluster formation, with evidence for one massive ($\\sim$10$^6\\,\\Msun$) cluster, with an age in the range 10--50~Myr, having triggered the formation of several younger massive clusters (up to $5\\times 10^6$\\,\\Msun) in the direction of a strong dust lane, with an amplification factor of $\\sim$8$\\times$ in the total mass of triggered clusters. 8. NICMOS observations show that $16 \\pm 6$\\% of the clusters in the Antennae are still largely embedded in their dust cocoons, with values of $3 \\la \\Av \\la 7$ mag. However, the top 50 IR-bright clusters {\\it all\\/} have optical counterparts in the much deeper ACS observations, hence the effect on our optically-based catalog derived from deep ACS images is minimal. In conclusion, the results presented here confirm and extend those based on previous WFPC2 data (e.g., Whitmore \\etal\\ 1999, Zhang \\& Fall 1999, Whitmore \\& Zhang 2002, FCW05, WCF07, FCW09). In addition, similarities between our results for the Antennae and those for a growing number of other star-forming galaxies (e.g., Mora \\etal\\ 2009) support the universality model for the formation and disruption of star clusters." }, "1005/1005.0611.txt": { "abstract": "{Free fall has signed the greatest markings in the history of physics through the leaning Pisa tower, the Cambridge apple tree and the Einstein lift. The perspectives offered by the capture of stars by supermassive black holes are to be cherished, because the study of the motion of falling stars will constitute a giant step forward in the understanding of gravitation in the regime of strong field. After an account on the perception of free fall in ancient times and on the behaviour of a gravitating mass in Newtonian physics, this chapter deals with last century debate on the repulsion for a Schwarzschild black hole and mentions the issue of an infalling particle velocity at the horizon. Further, black hole perturbations and numerical methods are presented, paving the way to the introduction of the self-force and other back-action related methods. The impact of the perturbations on the motion of the falling particle is computed via the tail, the back-scattered part of the perturbations, or via a radiative Green function. In the former approach, the self-force acts upon the background geodesic; in the latter, the geodesic is conceived in the total (background plus perturbations) field. Regularisation techniques (mode-sum and Riemann-Hurwitz $z$ function) intervene to cancel divergencies coming from the infinitesimal size of the particle. An account is given on the state of the art, including the last results obtained in this most classical problem, together with a perspective encompassing future space gravitational wave interferometry and head-on particle physics experiments. As free fall is patently non-adiabatic, it requires the most sophisticated techniques for studying the evolution of the motion. In this scenario, the potential of the self-consistent approach, by means of which the background geodesic is continuously corrected by the self-force contribution, is examined. } ", "introduction": "The two-body problem in general relativity remains one of the most interesting problems, being still partially unsolved. Specifically, the free fall, one of the eldest and classical problems in physics, has characterised the thinking of the most genial developments and it is taken as reference to measure our progress in the knowledge of gravitation. Free fall contains some of the most fundamental questions on relativistic motion. The mathematical simplification, given by the reduction to a 2-dimensional case, and the non-likelihood of an astrophysical head-on collision should not throw a shadow on the merits of this problem. Instead, it may be seen as an arena where to explore part of the relevant features that occur to general orbits, e.g. the coupling between radial and time coordinates. Although it is easily argued that radiation reaction has a modest impact on radial fall due to the feebleness of cumulative effects (anyhow, in case of high or even relativistic - a fraction of $c$ - initial velocity of the falling particle, it is reasonable to suppose a non-modest impact on the waveform and possibly the existence of a signature), it would be presumptuous to consider free fall simpler than circular orbits, or even elliptic orbits if in the latter adiabaticity may be evoked. Adiabaticity has been variously defined in the literature, but on the common ground of the secular effects of radiation reaction occurring on a longer time scale than the orbital period. One definition refers to the particle moving anyhow, although radiating, on the background geodesic (local small deviations approximation), of obviously no-interest herein; another, currently debated for bound orbits, to the secular changes in the orbital motion being stemmed solely by the dissipative effects (radiative approximation); the third to the radiation reaction time scale being much longer than the orbital period (secular approximation), which is a rephrasing of the basic assumption. But in radial fall such an orbital period doesn't exist. And as the particle falls in, the problem becomes more and more complex. In curved spacetime, at any time the emitted radiation may backscatter off the spacetime curvature, and interact back with the particle later on. Therefore, the instantaneous conservation of energy is not applicable and the momentary self-force acting on the particle depends on the particle's entire history. There is an escape route, though, for periodic motion. But energy-momentum balance can't be evoked in radial fall, lacking the opportunity of any adiabatic averaging. The particle reaction to its radiation has thus to be computed and implemented immediately to determine the effects on the subsequent motion. It is a no-compromise analysis, without shortcuts. Thus, the computation and the application of the back-action all along the trajectory and the continuous correction of the background geodesic, it is the only semi-analytic way to determine motion in non-adiabatic cases. And once this self-consistent approach shall be mastered for radial infall, where simplification occurs for the two-dimensional nature of the problem, it shall be applicable to generic orbits. It is worth reminding that the non-adiabatic gravitational waveforms are one of the original aims of the self-force community, since they express i) the physics closer to the black hole horizon; ii) the most complex trajectories; iii) the most tantalising theoretical questions. The head-on collisions of black holes and the associated radiation reaction were evoked recently in the context of particle accelerators and thereby showing the richness of the applicability of the radial trajectory also beyond the astrophysical realm. As gravity is claimed by some authors to be the dominant force in the transplanckian region, the use of general relativity is adopted for their analysis. This chapter reviews the problem of free fall of a small mass into a large one, from the beginning of science, whatever this may mean, to the application of the self-force and of a concurring approach, in the last fourteen years. There is no pretension of exhaustiveness and, furthermore, justifiably or not, from this review some topics have been disregarded, namely: any orbit different from radial fall; radiation reaction in electromagnetism; but also the head-on of comparable masses and Kerr geometry, post-Newtonian (pN) and effective one-body (EOB) methods; quantum corrections to motion. Herein, the terms of self-force and radiation reaction are used rather loosely, though the latter does not include non-radiative modes. Thus, the self-force describes any of the effects upon an object's motion which are proportional to its own mass. Nevertheless, to the term self-force is often associated a specific method and it is preferable to adopt the term back-action whenever such association is not meant. Geometric units ($G = c = 1$) and the convention (-,+,+,+) are adopted, unless otherwise stated. The full metric is given by ${\\bar g}_{\\alpha\\beta} (t,r) = g_{\\alpha\\beta}(r) + h_{\\alpha\\beta} (t,r)$ where $g_{\\alpha\\beta}$ is the background metric of a black hole of mass $M$ and $h_{\\alpha\\beta}$ is the perturbation caused by a test particle of mass $m$. ", "conclusions": "It has been shown that free fall is still the arena for a deeper comprehension of gravitation. Furthermore, it still generates acute observations like relativistic gliding for which the asymmetric oscillations of a quasirigid body slow down or accelerate its fall in a gravitational background \\cite{gumo07}. But is the problem of radial fall solved? Do we know the laws of motion of a star falling into a black hole, the relativistic modern version of the falling stone? A fair and objective answer leads to a moderate optimism. The general relativistic problem has had undeniable progress from 1997, but a careful analysis of the literature shows that some issues are either still partly open or simply not fully at hand, in terms of clear procedures, by means of which clearly cut answers are obtained. The remaining steps to be fulfilled for a satisfactory level of comprehension for radial fall of a small particle into a large mass (represented by a SD black hole) within the first perturbative order in $m/M$ are divided in three groups. Investigations to be pursued before recurring to the self-consistent prescription: \\begin{enumerate}[I] \\item{ Compute in proper time the first order perturbation equation with deviation terms, i.e. all terms of eq. (\\ref{eq:p13}).} \\item{ Identify and compare expressions for renormalised, semi-renormalised and unrenormalised accelerations in the SD perturbed geometry; evaluate repulsive effects. The last century debate on repulsion has not addressed the effects of mass finitude ($l=0,1$) and radiation reaction on the motion ($l\\geq 2$). The inclusion of perturbations blurs the issue further and renders it more complex. } \\item{Compare the waveform corrections obtained with energy-balance by Aoudia \\cite{ao08} with those to be obtained by the integration in the equation of motion of the approximate analytic expression of the self-force of Barack and Lousto \\cite{balo02}. The comparison is of interest only in case of concurring outcomes, as both procedures have a large degree of approximation. An other comparison \\cite{sp09} could be made with the numerical results for a 10 : 1 mass ratio, as larger mass ratios imply the solution of the numerical two-scale problem, not yet at hand.} \\item{Solve the initial conditions for particles with non-null or even relativistic initial velocities with and without large eccentricities. If the outcome of the previous item is successful, coarsely identify possible signatures of radiation reaction\\footnote{For Mino and Brink \\cite{mibr08}, the energy and momentum radiated are computed on the assumption that the small body falls in a dynamical time scale, with respect to proper time, well short of the radiation reaction time scale and therefore the gravitational radiation back-action on the orbit is considered negligible. The particle plunges on a geodesic trajectory, incidentally starting from a circular orbit, thus at zero initial radial velocity.}. It is likely that a non-null initial velocity would act as an amplifier of the waveform shifts due to radiation reaction. In \\cite{lopr97a}, Lousto and Price show that the energy radiated by a non-null initial velocity may rise up to almost two orders of magnitude (see also fig. 3 in \\cite{lo97}). Incidentally, this line of work is confluent with the interests shown recently in the particle physics community \\cite{ardodsto08, spcaprbego08, shokya08, ve10, yosh09} for head-on collisions and the associated radiation reaction \\cite{gakospto10a, gakospto10b}. Hopefully, progress in numerical relativity may lead to analysis of larger mass ratios and thereby test future results of perturbation theory\\footnote{For head-on collisions, Price and Pullin have surprisingly shown the applicability of perturbation theory for the computation of the radiated energy and of the waveform for two equal black holes, starting at very small separation distances \\cite{prpu94}, the so called close-limit approximation.} in the range beyond $10^{-3}$.} \\vskip 5pt Investigations to be performed solely by the implementation of the self-consistent prescription: \\vskip 5pt \\item{Evaluate the trajectory by means of the Gralla and Wald self-consistent method \\cite{grwa08}. At each integration step, for a given number of modes: evaluation of the perturbation functions at the position of the particle; regularisation by mode-sum or $\\zeta$ methods; correction of the geodesic and identification of the cell crossed by the particle; computation of the source term; reiteration of the above. Perturbations and self-force analysis in de Donder's gauge are on-going \\cite{balo05, basa10}.} \\item{Repeat steps 1-4 on the basis of the acquired self-consistent trajectory. An iterative scheme may be envisaged also in coordinate time.} \\vskip 5pt Investigations to be performed independently from the self-consistent prescription: \\vskip 5pt \\item{Identify whether in a thought radial fall experiment, there is a physical observable independent of the gauge choice or else manifesting a recognisable effects in a given gauge, following Detweiler \\cite{de08}.} \\item{Identify the domain of applicability of the self-force. Although self-consistency represents a closer description to perturbed motion, it is limited to cases where local deviations are small, since after all, it remains a perturbative approach. Thus, a quantitative identification of the domain of applicability of such prescription would be of interest to the community developing LISA templates. The adiabatic approximation can't be evoked to establish the limits of any self-force based analysis. Indeed, the hypothesis on the feeble magnitude of local deviations may be less constraining than the adiabatic hypothesis, for the absence of requirements on averaging. An explicit condition, referring to the geometry of the orbit and to the mass ratio and defining the domain of applicability of the self-force versus fully numerical approaches, is not yet available.} \\item{Investigate other gauges than Regge-Wheeler and other methods than the self-force for radial fall. A confirmation of the results by Barack, Lousto in \\cite{lo01, balo02} by an independent group is missing, though in \\cite{ao08} the necessary basic tools are developed, e.g. a numerical programme confirming the waveforms of \\cite{lopr97b, mapo02}. It would be also beneficial to carry out an analysis by the EOB method \\cite{da09b} \\footnote{A non-recent analysis by Simone, Poisson and Will \\cite{sipowi95}, between pN and perturbation methods in head-on collisions, was limited to the computation of the gravitational wave energy flux.}. } \\end{enumerate} The self-force community and even more the EMRI community are animated by different interests ranging from fundamental physics and theory, to numerical applications, data analysis and astrophysics. To these variegated communities and generally to physicists and astrophysicists, the capture of stars by supermassive black holes will bring a development comparable to the ancient markings made by the leaning Pisa tower, the Cambridge apple tree and the Einstein lift. \\begin{acknowledgement} Discussions throughout the years with S. Aoudia, L. Barack, S. Chandrasekhar, S. Detweiler, C. Lousto, J. Martin-Garcia, S. Gralla, E. Poisson, R. Price, R. Wald, B. Whiting are acknowledged. I would like to thank E. Lallier Verg\\`es for the support to the CNRS School on the Mass and the $11^{th}$ Capra conference, held in June 2008 in Orl\\'eans. It was my sincere hope that both events could put together different communities working on mass and motion: part of the contributors to this book and their colleagues (Blanchet, Detweiler, Le Tiec and Whiting) have already made concrete steps towards such cooperation \\cite{bldeltwh10a, bldeltwh10b}. \\end{acknowledgement}" }, "1005/1005.5491_arXiv.txt": { "abstract": "This short review is addressed to cosmologists.% \\footnote{Plenary talk at the \\emph{Invisible Universe} conference held in Paris in July 2009. To appear in the Proceedings, edited by J. M. Alimi et al (AIP Publications)} General relativity predicts that space-time comes to an end and physics comes to a halt at the big-bang. Recent developments in loop quantum cosmology have shown that these predictions cannot be trusted. Quantum geometry effects can resolve singularities, thereby opening new vistas. Examples are: The big bang is replaced by a quantum bounce; the `horizon problem' disappears; immediately after the big bounce, there is a super-inflationary phase with its own phenomenological ramifications; and, in presence of a standard inflaton potential, initial conditions are naturally set for a long, slow roll inflation independently of what happens in the pre-big bang branch. % ", "introduction": "\\label{s1} At this conference we heard of the spectacular progress that has occurred in observational cosmology in recent years. We also learned about the upcoming missions that are poised to provide new data to further constrain or even rule out leading theoretical models. These advances have been and continue to be the engines that drive contemporary cosmology. They have brought to forefront the astonishing success of the Friedmann, Lema\\^{\\i}tre, Robertson, Walker (FLRW) models, and perturbations thereof. Indeed, it appears that the rich data that we now have, and are likely to accumulate in the near future, would be adequately described by these simple applications of general relativity and quantum field theory on the resulting cosmological backgrounds. However these theories are conceptually incomplete. They assume that the universe begins with a big bang at which matter densities and space-time curvature become infinite. With inflationary scenarios there were initial hopes that perhaps the big-bang singularity could be avoided because the inflaton fails to satisfy the strong energy condition often used in the singularity theorems of general relativity. However, Borde, Guth and Vilenkin \\cite{bgv} soon established that this hope was misplaced. Inflation can be eternal in the future but, if we go back in time using Einstein's equations, one again finds that the space-time ends and physics simply comes to a halt at the big bang. But all our experience with fundamental physics suggests that this cannot be the situation in the real world. This must be a prediction of a theory that has been pushed well beyond the domain of its validity. To know what \\emph{really} happened near the putative big bang, we must work with a genuine unification of general relativity and quantum physics, an unification which does not pre-suppose that space-time is a smooth continuum, and which can encompass the rich non-linear structure of strong gravity that lies beyond the scope of perturbation theory. But the burden on this desired unification is heavy. In the context of cosmology there is a long list of fundamental questions that must be satisfactorily addressed by such a theory. Here is an illustrative list encompassing some of the contemporary issues. \\medskip \\begin{quote} \\noindent$\\bullet$ If general relativity is transcended, how close to the putative big bang does a smooth space-time of Einstein's make sense? In particular, can one show from first principles that this approximation is valid at the onset of inflation?\\\\ $\\bullet$ Is the big-bang singularity naturally resolved by the quantum version of Einstein's equations? Or, is some external input such as a new principle or a boundary condition at the big bang essential? An outstanding example of such an external input is the Hartle-Hawking proposal \\cite{hh}.\\\\ $\\bullet$ Is the quantum evolution across the `singularity' deterministic? One needs a fully non-perturbative framework to answer this question in the affirmative. In the pre-big-bang \\cite{pbb} and ekpyrotic/cyclic \\cite{ekp1,ekp2} scenarios, for example, so far the answer is in the negative because these theories pre-suppose the space-time continuum of general relativity and this approximation fails at the big-bang.\\\\ $\\bullet$ If the singularity is resolved, what is on the `other side'? Is there just a `quantum foam', far removed from any classical space-time, or, is there another large, classical universe? \\end{quote} Such questions are fundamental and must be faced squarely because the resolution of classical singularities can profoundly shift the paradigm underlying contemporary cosmology. This in itself makes it imperative that we understand the quantum nature of the big bang. But there could also be another rich pay-off: the new paradigm may enable us to address open issues that are observationally significant. For example, if there is a classical pre-big-bang branch to the universe, the horizon problem would disappear and the observed large scale homogeneity could be simply a consequence of the fact that even the most distant parts of the universe would have been in causal contact in the past. If there is a pre-big-bang branch, we would not have to specify the initial conditions for perturbations on the singularity, where the applicability of current theories is least reliable. There would be more natural ways of specifying these conditions which, in turn, may well lead to small, potentially observable deviations from current predictions. Finally, our experience with general relativity itself suggests that, once the physics of the Planck regime is well understood, we may be handed with novel predictions of central importance to the next generation of astrophysicists and cosmologists. Loop quantum gravity (LQG) \\cite{alrev,crbook,ttbook} is well suited to embark on this mission because it does not pre-suppose a classical space-time --it is background independent. At a fundamental level, everything, \\emph{including geometry,} is described in the paradigm of quantum physics \\cite{almmt,rs,al5,alvol}. Classical space-times emerge only on coarse graining of semi-classical quantum states. Finally, since the approach is fully non-perturbative, it is well suited for the strong field regime near the putative big bang. In this chapter I will discuss loop quantum cosmology (LQC), the application of the principles of LQG to cosmology \\cite{mb-rev,aa-badhonef}. Initial ideas appeared in \\cite{mb1,abl} and were developed in detail for a variety of cosmological models in \\cite{aps1,aps2,aps3,ps,apsv,kv,acs,cs1,bp,ap,aps4,hybrid,awe2,awe3,cs2}. \\emph{In all cases, the big-bang and big-crunch singularities are resolved in a direct physical sense.} The resulting Planck scale physics has been explored using analytical and numerical solutions to the quantum Einstein equations as well as effective equations which capture the leading quantum corrections. Singularity theorems are avoided \\emph{not} because one uses matter violating energy conditions. Indeed, in the models that have been studied in most detail, \\emph{all energy conditions hold}. The theorems are inapplicable because quantum geometry effects modify Einstein equations themselves. Physically, quantum geometry gives rise to a new repulsive force. This force is \\emph{utterly} negligible under normal circumstances. It is only when the matter density becomes about 1\\% of the Planck density or curvature approaches $1/\\lp^2$ that the repulsive force ---and hence the deviation from classical general relativity--- becomes significant. But then the repulsive force rises \\emph{very} quickly, overwhelms classical attraction and causes a quantum bounce. The density and curvature start falling and once they are below the scales just mentioned, the force again becomes negligible and classical general relativity again becomes an excellent approximation. Immediately to the future of the bounce there is a robust phase of super-inflation which is not encountered in general relativity \\cite{mb2,ps2}. But it is short lived and in absence of a suitable inflaton potential it does not yield a sufficient number of e-foldings. However, in presence of suitable potentials ---such as $m^2\\phi^2$--- super-inflation \\emph{funnels the phase space trajectories to initial conditions which virtually guarantee a slow roll inflation with 60 or more e-foldings} \\cite{as}. This is in striking contrast to what happens in general relativity where it has been argued \\cite{gt} that the probability of $N$ e-folding decreases as $e^{-3N}$. The super-inflationary phase is also likely to have other phenomenological consequences ---such as production of gravitational waves--- that are being analyzed. The article is organized as follows. Section \\ref{s2} lays out the conceptual setting and section \\ref{s3} provides a bird's eye view of LQC through illustrative results. We will see that not only has LQC answered many of the long standing questions but it has also opened new vistas. Section \\ref{s4} summarizes the origin of the novel predictions and places them in a broader perspective. ", "conclusions": "" }, "1005/1005.2278_arXiv.txt": { "abstract": "{} {We test predictions of evolution models on mixing of CNO-cycled products in massive stars from a fundamental perspective. Relative changes within the theoretical C:N:O abundance ratios and the buildup of helium are compared with observational results.} {A sample of well-studied Galactic massive stars is presented. High-quality optical spectra are carefully analysed using improved NLTE line-formation and comprehensive analysis strategies. The results are put in the context of the existing literature data.} {A tight trend in the observed $N/C$ vs. $N/O$ ratios and the buildup of helium is found from the self-consistent analysis of main-sequence to supergiant stars for the first time. The catalytic nature of the CNO-cycles is confirmed quantitatively, though further investigations are required to derive a fully consistent picture. Our observational results support the case of strong mixing, as predicted e.g. by evolution models that consider magnetic fields or by models that have gone through the first dredge-up in the case of many supergiants.} {} ", "introduction": "Energy production in massive stars is governed by the CNO-cycles throughout most of their lifetimes. The general correctness of our understanding of the CNO cycles and of the relevant nuclear data \\citep[e.g.][]{maeder83} is confirmed impressively by observation: when massive stars enter the Wolf-Rayet phase as WN subtypes, equilibrium CNO-processed material becomes exposed on their surface \\citep[see e.g.][]{crowther07}. However, traces of mixing of CNO-cycled products from the stellar core to the stellar surface can already be found much earlier in the lives of massive stars. Observational indications of superficial abundance anomalies for carbon, nitrogen, and oxygen (and the burning product helium) in OB-type stars even on the main sequence (MS) and, more prominently, in the blue supergiants was found early in classification spectrograms \\citep[e.g.][]{walborn76}. Subsequent analyses provided evidence for a characteristic enrichment of nitrogen -- which is the easiest to be detected -- and helium in many early-type stars, both near the MS and in blue supergiants \\citep[e.g.][]{schoenberner88,gila92,herrero92,kilian92,venn95,lyubimkov96,mcerlean99}. A theoretical understanding of {\\em early} mixing of CNO-pro\\-cessed material to the stellar surface could not be achieved within the framework of evolution calculations for non-rotating stars with mass-loss, which were state-of-the-art at that time \\citep[e.g.][]{chma86}. The pollution of the surface layers with CNO-cycled material only occurs when the star reaches the red supergiant phase in such models, via convective dredge--up. Considered a secondary effect for a long time, stellar rotation has come lately into focus. It turned out that rotationally-induced mixing through meridional circulation and turbulent diffusion in rotating stars provides the means to change all model outputs substantially and to bring theory and observation into much better agreement \\citep{mame00,hela00}. The latest step taken in the modelling was to consider the effects from an interplay of rotation and magnetic fields, which -- depending on the detailed input physics and approximations made -- only provide minor modifications to the surface abundances \\citep{heger05} or substantial changes \\citep[henceforth abbreviated as MM05]{mame05}. These differences result mainly from the two groups using different sets of equations for computing the effects of magnetic fields; see in particular, the changes brought to the system of equations given by \\citet{spruit02} by MM05 in their Sect.~2. The only means to verify the models is via systematic comparison with observations covering the relevant parts of the Hertzsprung-Russell diagram. Homogeneous analyses of a larger star sample from the main sequence to the supergiant stage have only recently become available. The results in particular for N abundances apparently challenge the concept of rotational mixing in massive stars \\citep{hunter09} and thus the present-day evolution models \\citep[see, however,][]{maederetal09}. \\begin{figure*} \\centering \\hfill \\includegraphics[width=.485\\textwidth]{14164fg1a.eps}\\hfill \\includegraphics[width=.485\\textwidth]{14164fg1b.eps} \\hfill \\caption{Status of observational constraints on the (magneto-)hydrodynamic mixing of CNO-burning products in massive stars from previous NLTE analyses. Mass ratios $N/C$ over $N/O$ are displayed. Left panel: main-sequence stars. Circles: \\citet{kilian92}; triangles: \\citet{gila92}; diamonds: \\citet{cula94}, \\citet{daflon99,daflon01a,daflon01b}; squares: \\citet{morel08}; crosses: \\citet{hunter09}. Right panel: BA supergiants. Triangles: \\citet{venn95}, \\citet{vepr03}; circles: \\citet{takeda00}; squares: \\citet{crowther06}; diamonds: \\citet{searle08}. Error bars are omitted for clarity: uncertainties in the abundances of the individual elements are typically about a factor 2, such that the error bars can be larger than the plotting range. The lines represent predictions from evolution calculations, for a rotating 15\\,$M_\\odot$ star \\citep[$v_\\mathrm{rot}^\\mathrm{ini}$\\,=\\,300\\,km\\,s$^{-1}$,][MM03; until the end of the MS: solid red line, until the end of He burning: dashed blue line]{mema03} and for a star of the same mass and $v_\\mathrm{rot}^\\mathrm{ini}$ that in addition takes the interaction of rotation and a magnetic dynamo into account \\citep[MM05; until the end of the MS: dotted line]{mame05}, respectively. The predicted trends are similar for the entire mass range under investigation, see Fig.\\ref{cnomix920}.} \\label{litsummary} \\end{figure*} Here, we address the topic of early mixing of CNO-cycled products in massive stars from a fundamental perspective. We compare the model predictions for the relative changes within the C:N:O abundance ratios and the buildup of helium (Sect.~\\ref{theory}) with observations. For this we briefly review the status of the literature on CNO abundances and introduce a well-studied Galactic star sample for which high-accuracy analyses of high-quality spectra were performed using improved NLTE modelling and comprehensive analysis techniques (Sect.~\\ref{obs}). Conclusions from this comparison are drawn in Sect.~\\ref{conclusions}. ", "conclusions": "} Evolution models for massive stars predict very tight relations for the change of surface $N/O$ and $N/C$ abundance ratios and for the buildup of helium, as a consequence of mixing with CNO-cycled matter (for given initial chemical composition). Massive stars in the solar neighbourhood, which supposedly share a nearly uniform initial composition (PNB08), are therefore expected to follow the predicted relations, which are governed by nuclear reactions and the dilution effects produced by mixing. A comparison with NLTE spectral analyses from the literature on CNO abundances in massive stars of $\\sim$8--25\\,$M_\\odot$ leaves room for broad interpretation because of large uncertainties. Observations may even be interpreted as posing a challenge to theory, in particular when the full set of information is not accounted for and conclusions are drawn from one indicator alone, e.g. nitrogen. On the other hand, our high-precision analyses of a sample of Galactic massive stars from the main sequence to the supergiant stage find these tight correlations. Even though further investigations are required to refine the observational constraints on helium abundances before a fully coherent picture can be obtained, the case for strong mixing is clearly supported. It is predicted, e.g., by the models of MM00, by models with magnetic field (MM05), or in cases where the models have gone through the first dredge-up. The separation of the different possibilities may come from studing the evolution of the rotation velocities with time. On the one hand, models with magnetic field predict higher rotational velocities at the end of the MS phase, because the strong internal coupling transmits some of the fast core rotation to the surface. On the other hand, stars on a blue loop should be slower rotators than objects on their first passage from blue to red on average because of the additional angular-momentum loss experienced through strong mass loss during the red supergiant phase. Indeed, some of the slowest rotators show very strong mixing signatures (Table~\\ref{table1}). Finally, we have to point out that some of our results have the potential of challenging the currently available evolution models. The star $\\tau$\\,Sco (HD149438) stands out in the sample as it shows characteristics that may be explained by a homogeneous evolution, but it requires a highly-efficient spin-down mechanism. One may speculate on magnetic breaking due to angular-momentum losses by a magnetically confined line-driven stellar wind or magnetic coupling to the accretion disc during the star-forming process in the case of a fossil field. Even though the topic is not understood theoretically in a comprehensive way, spin-down times of the order of 1\\,Myr \\citep{ud-Doula09} or even less \\citep{mikulasek08} are reported for some magnetic massive stars, possibly leading to the required slow rotation already on the zero-age MS in this case as well. Then, if the present helium abundances are confirmed and the supergiants are shown to have evolved directly from the MS, a different kind of mixing may also be required. We conclude that the tight observational constraints that are required for a thorough testing of the stellar evolution models are within reach. Applications of the improved modelling and analysis techniques to high-quality observations of larger star samples in the Milky Way and other galaxies may finally provide the empirical basis to benchmark the models. It may thus become feasible to disentangle the effects of metallicity, rotation, magnetic fields, and binarity on massive star evolution." }, "1005/1005.3754_arXiv.txt": { "abstract": "{Oxygen is the third most common element in the Universe. The measurement of oxygen lines in metal-poor unevolved stars, in particular near-UV OH lines, can provide invaluable information about the properties of the Early Galaxy.} {Near-UV OH lines constitute an important tool to derive oxygen abundances in metal-poor dwarf stars. Therefore, it is important to correctly model the line formation of OH lines, especially in metal-poor stars, where 3D hydrodynamical models commonly predict cooler temperatures than plane-parallel hydrostatic models in the upper photosphere.} {We have made use of a grid of 52 3D hydrodynamical model atmospheres for dwarf stars computed with the code \\cobold, extracted from the more extended CIFIST grid. The 52 models cover the effective temperature range 5000--6500\\,K, the surface gravity range 3.5--4.5 and the metallicity range $-3 < [{\\rm Fe}/{\\rm H}] < 0$.} {We determine 3D-LTE abundance corrections in all 52 3D models for several OH lines and \\ion{Fe}{i} lines of different excitation potentials. These 3D-LTE corrections are generally negative and reach values of roughly -1~dex (for the OH 3167 with excitation potential of approximately 1~eV) for the higher temperatures and surface gravities.} {We apply these 3D-LTE corrections to the individual O abundances derived from OH lines of a sample the metal-poor dwarf stars reported in Israelian et al.(1998, 2001) and Boesgaard et al.(1999) by interpolating the stellar parameters of the dwarfs in the grid of 3D-LTE corrections. The new 3D-LTE [O/Fe] ratio still keeps a similar trend as the 1D-LTE, i.e, increasing towards lower [Fe/H] values. We applied 1D-NLTE corrections to 3D \\ion{Fe}{i} abundances and still see an increasing [O/Fe] ratio towards lower metallicites. However, the Galactic [O/Fe] ratio must be revisited once 3D-NLTE corrections become available for OH and Fe lines for a grid of 3D hydrodynamical model atmospheres. } ", "introduction": "\\label{intro} The metal-poor stars of the Galactic halo provide the fossil record of the early Galaxy's composition. Dwarf halo stars are particularly relevant, because their atmospheres are not significantly altered by internal mixing and provide a unique tracer to constrain Galactic chemical evolutionary models. Oxygen is a key element in this scenario, because it is the most abundant element in stars after H and He. It is produced in the interiors of massive stars by hydrostatic burning and its content is modified during the explosive nucleosynthesis in type~II supernovae (SNe) and hypernovae (HNe) and returned into the interstellar medium. On the other hand, iron is created by both type~II and type~I SN explosions. However, type~I SNe progenitors have longer lifetimes, which is why the abundance ratio of [O/\\ion{Fe}{i}] can be used to constrain the chemical evolution of the Galaxy. There have been numerous studies of the oxygen abundance in halo stars. Despite considerable observational and theoretical efforts, the trend of [O/Fe] ratio\\footnote{$[{\\rm O}/{\\rm Fe}]=\\log [N({\\rm O})/N({\\rm Fe})]_\\star-\\log [N({\\rm O})/N({\\rm Fe})]_\\odot$} versus [Fe/H] is still unclear. The analysis of the forbidden line \\ion{O}{i} 6300~{\\AA} in giants shows a plateau with [O/Fe$] \\sim 0.4-0.5$ in the metallicity range $-2.5 < [{\\rm Fe}/{\\rm H}] < -1$ \\citep{bar88} and [O/Fe$] \\sim 0.7$ for $-4. < [{\\rm Fe}/{\\rm H}] < -2.5$ \\citep{cay04}. A similar behaviour is seen for metal-poor subgiant stars in the range $-3. < [{\\rm Fe}/{\\rm H}] < -1.5$ \\citep[{[O/Fe]} $\\sim 0.4-0.5$,][]{gar06}. However, \\citet{gar06} already noted that by plotting all measurements from the [OI] line for dwarfs \\citep{nis02}, subgiants \\citep{gar06} and giants \\citep{cay04}, the picture changes and an increasing trend [O/Fe] towards lower metallicities clearly appears. The near infrared (IR) triplet \\ion{O}{i} 7771--5~{\\AA} in metal-poor dwarfs and subgiants \\citep{abi89,isr01,gar06} points towards increasing [O/Fe] values with decreasing [Fe/H], although the O abundances derived from the near-IR triplet are typically $\\sim 0.4-0.7$ higher than those derived from the forbidden \\ion{O}{i} line \\citep{ful03}. The OH A-X electronic lines in the near ultraviolet (UV) provide also higher [O/Fe] ratios towards lower [Fe/H] values in dwarf stars \\citep{isr98,boe99,isr01,gon08}. However, \\citet{gar06} found a quasi-plateau of [O/Fe] for subgiant stars in the range $-3. < [{\\rm Fe}/{\\rm H}] < -1.5$, using Fe abundances determined from \\ion{Fe}{ii} lines. The [O/Fe] ratio shows a negative slope if one instead uses the Fe abundances estimated from \\ion{Fe}{i} lines, although with lower [O/Fe] values than those determined for metal-poor dwarfs. It is advisable to use \\ion{Fe}{i} lines instead of \\ion{Fe}{ii} lines to derive the [O/Fe] ratio because of its similar sensitivity to the surface gravity. These O-abundance indicators present different complications. The near-IR \\ion{O}{i} triplet is susceptible to non-local thermodynamical equilibrium (NLTE) effects \\citep[and references therein]{kis01}, with abundance corrections below 0.2 dex, and is quite sensitive to the adopted \\teffo. The [OI] is not sensitive to departures from LTE, but it is essentially undetectable in dwarfs with $[{\\rm Fe}/{\\rm H}]\\lesssim -2$. The near-UV OH lines are strongly sensitive to the temperature structure and inhomoginities \\citep{asp01,gon08}. In addition, \\ion{Fe}{i} lines suffer from severe NLTE effects in metal-poor stars \\citep[see e.g.][]{the99}. We note that \\citet{shc05} have performed NLTE computations for the metal-poor subgiant HD140283 ($[{\\rm Fe}/{\\rm H}]\\sim -2.5$) with a single snapshot of a 3D hydrodynamical simulation \\citep{asp99}, and found NLTE-LTE corrections of $+0.9$ and $+0.4$ for \\ion{Fe}{i} and \\ion{Fe}{ii} lines, respectively. Finally, the oxygen abundance in the Sun is still a matter of debate. We will adopt throughout this work the value of $\\log [N({\\rm O})/N({\\rm H}]_\\odot =8.76$, which was determined through 3-dimensional (3D) hydrodynamical models \\citep{caf08}. We have used a subset of the CIFIST grid of 3D hydrodynamical model atmospheres \\citep{lud09} to investigate the 3D-LTE and 3D-NLTE [O/Fe] and [O/H] trends in metal-poor dwarf stars. ", "conclusions": "The large grid of 3D hydrodynamical model atmospheres of dwarf stars quite captivated our attention during the last three years \\citep{lud09}. It took thousands of hours in computing time to build such a grid. We have used 52 3D models extracted from this grid with the following stellar parameters and metallicities: \\teff$=5000$, 5500, 5900, 6300 and 6500~K, \\logg$=3.5$, 4 and 4.5, and [Fe/H$]=0$, --1, --2 and --3. The main difference with the ``classical'' 1D models is that 3D models show temperature inhomogeneities and a cooler average temperature profile in models with [Fe/H$]<-1$. This allowed us for the first time to compute 3D abundance corrections of several near-UV OH and \\ion{Fe}{i} lines. These 3D corrections are generally larger for higher effective temperatures, larger surface gravities, and lower metallicities. In addition, lines with lower excitation potentials show stronger 3D corrections. We applied this grid of 3D corrections to a sample of metal-poor dwarf stars from \\citet{isr98,isr01}, \\citet{boe99} and the most metal-poor dwarf stars of the binary \\mbox{CS 22876--032} from \\citet{gon08}. We interpolated within this grid, using the stellar parameters and metallicities of these stars. Finally, we are able to display a new trend with the 3D-NLTE [O/\\ion{Fe}{i}] ratio versus metallicity and this trend still increases towards lower metallicities, as the 1D-NLTE, but with a smaller slope in absolute value. However, we caution that some assumptions (as e.g. the restriction to a 6-bin scheme in most 3D models and NLTE corrections only in 1D), which we believe to be reasonable, have been adopted to achieve this result and that therefore, this result should be taken with that in mind. A full 3D-NLTE study of all 3D model atmospheres presented in this paper must be performed for both near-UV OH lines and \\ion{Fe}{i} lines to see if this trend changes." }, "1005/1005.4668_arXiv.txt": { "abstract": "The PAMELA satellite has observed an excess of positrons over electrons in the energy range $1-100~GeV$ that increases with energy. We propose that the excess is not due to a change in the local interstellar spectrum, but is due to heliospheric modulation. We motivate this from the known form of the heliospheric magnetic field and predict that the excess will disappear when we enter a period of solar maximum activity. ", "introduction": "The PAMELA satellite \\cite{Adriani:2009p431} has observed a rise in the positron to electron fraction that increases with energy above $10~GeV$. This has been interpreted as proof of a rise in the positron fraction in the local interstellar spectrum of cosmic rays. Such a result requires that there be some hardening of the positron spectrum at these energies and has caused a great deal of discussion of possible sources from decaying or annihilating dark matter to acceleration from a population of pulsars. We propose instead that this rising fraction is due to the heliospheric magnetic field. The heliospheric magnetic field is ordered over large distances during solar minimum \\cite{Zhou:2009p1228,UlyssesBook} and is well described up to medium latitudes by a Parker spiral field \\cite{Parker:1958p1139}. During solar minimum this means that for a particle to propagate directly to earth it must traverse a large scale ordered magnetic field that is primarily oriented perpendicularly to the radial. Direct transport is impossible for low energy particles ($E<1~GeV$) as their Larmor radius is small even in the outer heliosphere. They instead reach the central solar system primarily by following the magnetic field lines along with a complex process involving convection, diffusion, drifts and energy loss. At high energies ($E>1~TeV$), particles can traverse the ordered magnetic field with little deviation. At these energies we see the interstellar spectrum directly. If the magnetic field were the only effect on the particle transport, and if the magnetic field were time invariant, then the switch from one set of particles paths to the other would have no effect on the flux by Liouville's theorem. However the magnetic field is time variant. The magnetic field lines are carried by the supersonic solar wind and travel out towards the edge of the heliosphere at 400~km/s. This causes large scale convection of cosmic rays that results in a deficit in the cosmic ray density in the central solar system when compared to the interstellar spectrum. This has been well measured for cosmic rays with rigidities up to 2.5GV \\cite{UlyssesBook,Ndiitwani:2005p2956}. At these rigidities, the particles generally have a very long path length to the center of the solar system as they follow the magnetic field lines. Therefore they have a low radial velocity, and are greatly effected by the convection of the solar wind. The solar magnetic field also varies on longer timescales, affected by the 27 day rotation period of the sun, and by the reversal of the solar magnetic field every 11 years. At higher rigidities particles can travel directly to the central solar system at the speed of light. Their path length within the heliosphere is short and they are almost unaffected by convection from the solar wind. Therefore at high energies we expect there to be no radial gradient of cosmic rays as they should free stream through the solar system. From these two observations - the two separate transport paths for low and high energy particles, and radial gradient of low rigidity cosmic rays from the convection of the solar wind - we can already say that there must be a hardening of the cosmic ray spectrum between rigidities of $1~GV$ and $1000~GV$. This will be due to the change over from a locally suppressed cosmic ray flux to the interstellar flux. To explain the PAMELA signal we need one further element. We require that the change over from the local flux to the interstellar flux be charge asymmetric. We already know that the orientation of the magnetic field results in asymmetric cosmic ray transport. Indeed this is the cause of the low energy heliospheric modulation below a few $GeV$\\cite{Potgieter:2001p1465}. We argue that this is also true for high energy particles and that the asymmetry should become more pronounced with increasing energy. We use a simple model of the heliospheric magnetic field to show this in section \\ref{lense1}. This exhibits the primary features of the asymmetry that we are interested in. We then extend the study to consider the asymmetry in a more complex realisation of the field in section \\ref{lense2}. By postulating a heliospheric cause of the positron excess we can immediately make a clear prediction. We predict that the excess should disappear when the solar system enters a period of solar maximum activity which it is due to do shortly. In addition, in the opposite solar cycle when the sun's magnetic field is oppositely aligned, we would expect to see the opposite modulation - giving a rising excess of electrons over positrons in the energy range $1-100~GeV$. We will also show that the tilt angle of the heliospheric current sheet (HCS) is an important feature of the model, and predict that any temporal variation of the latitudinal extent of the current sheet should be matched by a variation in the observed positron excess, with a time delay due to the time taken for magnetic field features at the sun to propagate out to the heliopause in the solar wind. These predictions are robust features of our model and testable in the near future. ", "conclusions": "} We have proposed that the PAMELA positron excess is not down to annihilating dark matter or new astrophysical sources, but instead results from the configuration of the heliospheric magnetic field. The ordered nature of the magnetic field on large scales creates a lense that allows particles of one sign to free stream into the center of the solar system whilst particles off the opposite sign travel out. This effect rises with energy. It naturally occurs in the correct energy range for the correct sign of particle to account for the observed positron excess. Though our model of the solar system magnetic field is rough, we can already make two clear predictions. Firstly, the PAMELA result will disappear when we go through solar maximum (with a suitable delay to account for the time it takes the conditions at the sun to propagate out through the solar system). Secondly, we predict that there will be fluctuations in the observed excess that correlate (again with a delay) to the tilt angle of the heliospheric current sheet." }, "1005/1005.4497_arXiv.txt": { "abstract": "We study mass transfer by Roche lobe overflow in close-in exoplanetary systems. The planet's atmospheric gas passes through the inner Lagrangian point and flows along a narrow stream, accelerating to 100-$200~\\kms$ velocity before forming an accretion disk. We show that the cylinder-shaped accretion stream can have an area (projected in the plane of the sky) comparable to that of the planet and a significant optical depth to spectral line absorption. Such a ``transiting cylinder'' may produce an earlier ingress of the planet transit, as suggested by recent HST observations of the WASP-12 system. The asymmetric disk produced by the accretion stream may also lead to time-dependent obscuration of the starlight and apparent earlier ingress. We also consider the interaction of the stellar wind with the planetary magnetosphere. Since the wind speed is subsonic/sub-Alfvenic and comparable to the orbital velocity of the planet, the head of the magnetopause lies eastward relative to the substellar line (the line joining the planet and the star). The gas around the magnetopause may, if sufficiently compressed, give rise to asymmetric ingress/egress during the planet transit, although more works are needed to evaluate this possibility. ", "introduction": "The close-in exoplanets (with period less than a few days) discovered in radial velocity and transit surveys are of great interest, not only because they constrain theories of planet formation and evolution, but also because they provide a probe of various physical processes that are otherwise unimportant in ``normal'' planets. WASP-12b is a transiting exoplanet orbiting extremely close to a late-F/early-G star ($M_\\star=1.35M_\\sun$, $R_\\star=1.57R_\\sun$, $T_{\\rm eff}=6300$~K), with the orbital period $P=1.09$~days and orbital semi-major axis $a=0.023~{\\rm AU}=4.94R_\\sun =3.15R_\\star$. The planet mass $M_p=1.41M_J$ and radius $R_p=1.79R_J$, as determined by transit observation in optical continuum (Hebb et al.~2009; Campo et al.~2010). The planet is one of the most irradiated exoplanets (with equilibrium temperature $T_{\\rm eq}=2500$-3000~K) and is highly inflated. The fact that the Roche radius (Hill sphere radius) of the planet, $R_L=a(M_p/3M_\\star)^{1/3}=1.85R_p$, is only slightly larger than the $R_p$ derived from optical transit measurements, suggests that mass loss from the planet is likely (Li et al.~2010). The small orbital separation also suggests that stellar wind may influence the atmosphere the planet. Recently, Fossati et al.~(2010) obtained near-UV transmission spectroscopy of WASP-12b with the Cosmic Origins Spectrograph on the {\\it Hubble Space Telescope}. The data revealed enhanced transit depths (by about a factor of 2) in two wavelength bands, NUVA (2539-2580~$\\AA$) and NUVC (2770-2811~$\\AA$), which were attributed to flux attenuation by absorption lines of metals in the vicinity of the planet. Most interestingly, the NUVA data exhibits an earlier ingress compared to the transit in optical J,B and Z bands, while the egress of the transit occurs at about the same time as the optical transit. The asymmetric behavior of the ingress/egress in the NUVA band relative to the continuum is difficult to understand if the absorbing gas surrounding the planet arises entirely from an irradiation-driven wind -- such a wind has been studied extensively in the context of hot Jupiter HD 209458b (e.g., Yelle 2004; Tian et al.~2005; Garcia Munoz 2007; Murray-Clay et al.~2009): The wind is most strongly generated at the planet's dayside, and would be distributed on both sides (terminations) of the substellar line (the line joining the planet and the star). If anything, it would be preferentially on the west side because of the planetary rotation, which is almost likely synchronized with the orbit (see Schneiter et al.~2007 for a simulation). In this paper, we consider two possible explanations for the asymmetric excess absorption during WASP-12b transit. First, we study the Roche lobe overflow from the planet to the parent star (see Li et al.~2010) and the associated accretion stream. The stream is asymmetric with respect to the line joining the planet and the star. We show that the stream has a sky-projected area comparable to the projected planet area, and a sufficiently large column density to cause blockage of the star light prior to optical ingress. Second, we qualitatively discuss the magnetopause produced by the interaction between the stellar wind and the planet's magnetosphere. Because the planet's orbital velocity, $v_{\\rm orb}=(GM_\\star/a)^{1/2} =228~\\kms$, is comparable to the stellar wind velocity ($v_w\\simeq 100~{\\rm km~s}^{-1}$ at $a=0.023$~AU), the head of the magnetopause lies eastward relative to the substellar point. Again, ingress/egress asymmetry may be produced if the gas density in around the magnetopause is sufficiently high. This paper is organized as follows. In Sect.~2 we discuss line absorption by a moving medium with a velocity gradient and derive the observational constraints on the absorbing gas in WASP-12b. In Sect.~3 we study the property of the accretion stream and the obscuration of the star light by the stream. Section 4 examines the possibility that the gas around the magnetopause may absorb the star light. We conclude in Sect.~5. ", "conclusions": "Motivated by the recent near-UV spectral observations of the WASP-12 planetary system (Fossati et al.~2010), which showed an earlier ingress of the planet's transit compared to the optical continuum, we have studied two possible explanations for the ingress/egress asymmetry. The first involves mass transfer through Roche lobe overflow, which results in an elongated accretion stream in which gas flows from the L1 Lagrangian point toward the star. Our analysis of the geometric and velocity structure of the stream shows that it may indeed provide asymmetric obscuration of the star light during the planet transit. In addition, it is also possible that an asymmetry in the accretion disk, caused by the impact of the accretion stream, produces an apparent earlier ingress. Another possibility involves the magnetopause separating the stellar wind and the planetary magnetosphere. Because of the planet's orbital motion, the head magnetopause lies eastward relative to the substellar point. We suggest that line absorption by the gas around the magnetopause may also explain the asymmetric ingress/egress behavior, although more theoretical works are needed to understand the property of the absorbing gas around the magnetopause. The two possibilities studied in this paper may be distinguished by the fact that in the accretion stream, the gas falls away from the observers, producing redshifted absorption, while the flow around the magnetosphere tends to move toward the observer, thus producing blueshifted absorption. Finally, we note the observed behaviors reported by Fossati et al are only marginally significant (at most $3\\sigma$ effects). Thus, more observations will be useful. Nevertheless, the physical processes discussed in this paper are quite general and therefore may be applicable in other close-in exoplanetary systems. \\medskip We thank the participants of the morning coffee of the KITP Exoplanets program (2010.1-2010.5) for lively discussions. D.L. especially thanks Phil Arras, Doug Lin and Steve Lubow for useful discussion during the final phase of our work, and P. Arras for his comments on an earlier draft of our paper. C.H. thanks Luca Fossati for discussion of early-ingress observations. Part of this work was performed while the authors were in residence at KITP, funded by the NSF Grant PHY05-51164." }, "1005/1005.3048_arXiv.txt": { "abstract": "We present molecular line observations, made with angular resolutions of $\\sim$20\\arcsec, toward the filamentary infrared dark cloud G34.43+0.24 using the APEX [\\dco , \\tco, \\cdo\\ and \\cssiete\\ transitions], Nobeyama 45 m [\\csdos, \\sio, \\cstyc, \\hco , \\hcotrece\\ and \\metanol\\ transitions] , and SEST [\\csdos\\ and \\cdodos\\ transitions] telescopes. We find that the spatial distribution of the molecular emission is similar to that of the dust continuum emission observed with 11\\arcsec\\ resolution (Rathborne et al. 2005) showing a filamentary structure and four cores. The cores have local thermodynamic equilibrium masses ranging from $3.3\\times10^2$ -- $1.5\\times10^3$ \\Msun\\ and virial masses from $1.1\\times10^3$ -- $1.5\\times10^3$ \\Msun, molecular hydrogen densities between $1.8\\times10^4$ and $3.9\\times10^5$ cm$^{-3}$, and column densities $>2.0\\times10^{22}$ cm$^{-2}$; values characteristics of massive star forming cores. The \\tco\\ profile observed toward the most massive core reveals a blue profile indicating that the core is undergoing large--scale inward motion with an average infall velocity of 1.3 \\kms\\ and a mass infall rate of $1.8\\times10^{-3}$ \\Msun\\ yr$^{-1}$. We report the discovery of a molecular outflow toward the northernmost core thought to be in a very early stage of evolution. We also detect the presence of high velocity gas toward each of the other three cores, giving support to the hypothesis that the excess 4.5 $\\mu$m emission (``green fuzzies'') detected toward these cores is due to shocked gas. The molecular outflows are massive and energetic, with masses ranging from 25 -- 80 \\Msun, momentum 2.3 -- $6.9\\times10^2$ \\Msun\\ \\kms, and kinetic energies 1.1 -- $3.6\\times 10^3$ \\Msun\\ km$^2$ s$^{-2}$; indicating that they are driven by luminous, high-mass young stellar objects. ", "introduction": "Infrared dark clouds (IRDCs) are cold ($<$25 K), massive ($\\sim$$10^2$ -- $10^4$ \\Msun), and dense ($>$$10^5$ cm$^{-3}$) molecular clouds with high column densities ($\\sim$$10^{23}$--$10^{25}$ cm$^{-2}$) seen as a dark silhouette against the bright mid--infrared background emission (Perault et al. 1996; Egan et al. 1998; Carey et al. 1998, 2000; Hennebelle et al. 2001; Simon et al. 2006a, 2006b). A wealth of recent observations show that IRDCs are the sites of high-mass stars and star cluster formation (Rathborne et al. 2006, 2007; Pillai et al. 2006; Jackson et al. 2008; Chambers et al. 2009). Dust continuum observations reveal that IRDCs harbor compact cold cores with typical masses of $\\sim$120 \\Msun\\ and sizes $<$0.5 pc (Rathborne et al. 2006). The masses and sizes of the IRDC cores are similar to those of the massive hot cores (e.g., Garay \\& Lizano 1999), but IRDC cores are much colder. Recently, Chambers et al. (2009) investigated a sample of 190 cores found toward 38 IRDCs by Rathborne et al. (2006), and classified them as ``active'' if they are associated with enhanced 4.5 $\\mu$m emission, the so-called green fuzzies, and an embedded 24 $\\mu$m source; and ``quiescent'' if they contain neither of these indicators. They found 37 active cores and 69 quiescent cores. Further, active cores are usually associated with bright ($>$1 Jy) methanol and water masers (Pillai et al 2006; Wang et al. 2006; Chambers et al. 2009), and have high bolometric luminosities, indicating that they are forming high--mass stars (M $>$ 8 \\Msun) stars. The quiescent cores are the most likely candidates for high--mass starless cores. G34.43+0.24 is an IRDC with a filamentary morphology, extending by $\\sim$9\\arcmin\\ from north to south in equatorial projection (9.7 pc at the source distance of 3.7 kpc; Fa\\'undez et al. 2004; Simon et al. 2006b). This IRDC is in the list of Rathborne et al. (2006) and Chambers et al. (2009). It is located roughly 11\\arcmin\\ north of the ultra-compact (UC) H II complex G34.26+0.15 (Molinari et al. 1996). Molecular line observations toward this cloud were first reported by Miralles et al. (1994), who from ammonia observations with 1.5\\arcmin\\ resolution found an elongated structure in the N-S direction with a total mass of 1000 \\Msun. Near the center of the filament lies the IRAS point source 18507+0121 which has a luminosity of $3.4\\times10^4$ \\Lsun. Bronfman et al. (1996) detected toward IRAS 18507+0121 strong \\csdos\\ emission with broad line wings, indicating it is associated with a dense massive star forming region. Ramesh et al. (1997) observed the IRAS source in the \\hco, \\hcotrece, \\csdos\\ and \\cstyc\\ molecular lines, with angular resolutions of $\\sim$16\\arcsec, modeling the observed line profiles as due to emission from a collapsing hot core that is hidden behind a cold and dense envelope. From observations of 3 mm continuum and H$^{13}$CO$^+$ line emissions, with $\\sim$5\\arcsec\\ angular resolution, Shepherd et al. (2004) detected two compact molecular cores, separated by 40\\arcsec, toward the center of the filament. Millimeter continuum observations revealed the presence of four dust cores (Fa\\'undez et al. 2004; Garay et al. 2004; Rathborne et al. 2005), labeled by the latter authors as MM1, MM2, MM3, and MM4. The MM1 dust core, which corresponds to the northern compact molecular core of Shepherd et al. (2004), contains a deeply embedded luminous object surrounded by several hundred solar masses of warm gas and dust. Based on the weak 6 cm continuum emission and the lack of detection at NIR wavelenghts, Shepherd et al. (2004) suggested that the embedded object appears to be a massive B2 protostar in an early stage of evolution. The MM2 dust core, which corresponds to the southern compact molecular core of Shepherd et al. (2004), is associated with IRAS 18507+0121; with a NIR cluster of young stars with a central B0.5 star (Shepherd et al. 2004); with an UC H II region (Miralles et al. 1994; Molinari et al. 1998); with a variable H$_2$O maser (Miralles et al. 1994); and with CH$_3$OH maser emission (Szymczak et al. 2000). The MM3 dust core is located close to the northern edge of the filament, $\\sim$3.5\\arcminspace north of MM2. Garay et al. (2004) suggested that MM3 corresponds to a massive and dense cold core that will eventually collapse to form a high-mass star. The MM4 core is located $\\sim$30\\arcsecspace south of MM2 (Rathborne et al. 2005). Table~\\ref{tbl-1.2mm} lists the measured 1.2 mm dust continuum emission properties as measured by Rathborne et al. (2006). Rathborne et al. (2005) and Chambers et al (2009) reported an excess 4.5 $\\mu$m emission toward MM1, MM3, and MM4 cores, which could be produced by either ionized gas and/or shocked gas and acts therefore as a signature of current star formation (MM2 is classified by the last authors as a ``red'' core, namely with bright 8 $\\mu$m emission). They also showed that the four MM cores are associated with 24 $\\mu$m point sources (emission from warm dust), which is an indicator of accretion onto an embedded protostar. Water maser emission, a well--known tracer of low--mass and high--mass star formation, has been detected toward all four MM cores (Wang et al. 2006; Chambers et al. 2009). Chambers et al. (2009) detected Class I CH$_3$OH masers, another well--know signpost of star formation toward MM1, MM2, and MM3 cores. Rathborne et al. (2006) identified nine 1.2 mm continuum cores with masses ranging from 80 to 1300 \\Msun\\ toward G34.43+0.24. Shepherd et al. (2007) observed the central region of the filament, encompassing the MM1 and MM2 cores, at millimeter wavelengths, in CO$(1\\rightarrow0)$, $^{13}$CO$(1\\rightarrow0)$, and C$^{18}$O$(1\\rightarrow0)$ at $\\sim4$\\arcsec\\ angular resolution and at near infrared wavelengths with the {\\it Spitzer} Space Telescope. They discovered five massive outflows, three of them associated with the MM2 core and the remaining two associated with the MM1 core. From the {\\it Spitzer} data they identified 31 young stellar objects in the filament, with a combined mass of $\\sim$127 \\Msun, and additional 22 sources that they believe could be cluster members based on the presence of strong 24 $\\mu$m emission. Cortes et al. (2008) observed MM1 and MM2 in polarized thermal dust emission at 3 mm and CO$(1\\rightarrow0)$ line emission with BIMA array. Their results suggest that there is a magnetic field othogonal to the axis formed by the MM1 and MM2 cores. Recently, Rathborne et al. (2008) carried out high angular submillimeter observations toward the MM1 core using the Submillimeter Array. They determined, using continuum emission images, that MM1 remains unresolved with a size of $\\sim$0.03 pc and a mass of 29 \\Msun. Their molecular line spectrum shows the presence of several complex molecules, suggesting that MM1 is a hot core. They also found an extended \\tco\\ structure which may be evidence of a rotating envelope surrounding the central high--mass star, perpendicular to one of the outflows that Shepherd et al. (2007) detected toward MM1. In this work we report observations of 11 molecular transitions using the APEX 12-m, SEST 15-m, and Nobeyama 45-m telescopes toward the G34.43+0.24 cloud. Since different molecular species trace different density and kinematics regimes, the capability of observing multiple spectral lines is particularly advantageous to determine the morphology and kinematics of star forming cores. Due to the high abundance of CO, the \\dco\\ line is an ideal tracer of the low intensity, high--velocity wing emission. The \\tco\\ line, being more optically thin than \\dco, traces, in addition to the ambient gas, the low--velocity gas of the molecular outflow. The \\sio\\ line is usually associated with the presence of outflows since its abundance is highly enhanced by shocks. Methanol transitions are usually used as tracers of dense gas, temperature probes and tracers of shocks, since this molecule is evaporated from grain mantles. The \\hco\\ and \\csdos\\ lines are high--density tracers and thus are good probes of the dense envelopes surrounding protostars. The \\cdo, \\cdodos, \\hcotrece, \\cssiete\\ and \\cstyc\\ molecular transitions probe the inner regions of the cores. The last two species are weakly influenced by infall and outflow motions, and are used as reference to estimate the central ambient cloud velocity. The main goals of these multi--line observations are to determine the morphology and kinematics of the gas within the molecular cores in the G34.43+0.24 filamentary IRDC, in particular to investigate the presence of outflowing gas, in order to increase our knowledge of the evolutionary state of the cores and to achieve a better understanding of the overall process of massive star formation along the filament. This is the first survey of several molecular lines mapping portions and the whole cloud. It gives us, for the first time, a comprehensive view of the molecular gas distribution at the scale of parsecs. ", "conclusions": "We carried out a multi--line study of cores within the filamentary IRDC G34.43+0.24. The emission from the filament was fully mapped in the \\dco\\ and \\tco\\ lines. In addition, emission from the central region of the filament was observed in eight different molecular transitions [\\cdo, \\cssiete, \\csdos, \\cstyc, \\hco, \\hcotrece, \\sio\\ and \\metanol], and from the north region in the \\csdos\\ and \\cdo\\ lines. The results of this study are summarized as follows. The spatial distribution of the molecular emission in all, except methanol, lines is similar to that of the dust continuum emission derived from observations with similar angular resolution, showing a filamentary structure and four cores. The mass of the molecular cores derived assuming LTE conditions are 300 and 1460 \\Msun\\ for the MM1 and MM2 cores, respectively. The masses derived assuming virial equilibrium are 1100, 1500 and 1400 \\Msun\\ for the MM1, MM2 and MM3 cores, respectively. The average molecular hydrogen densities of the cores are $1.8\\times 10^5$ cm$^{-3}$ in MM1, $1.7\\times 10^5$ cm$^{-3}$ in MM2 and $1.9\\times10^4$ cm$^{-3}$ in MM3. We also find that the molecular hydrogen column densities at the peak positions of MM1 and MM2 are, respectively, $2\\times 10^{22}$ cm$^{-2}$ and $8\\times 10^{22}$ cm$^{-2}$. The derived parameters of the MM1 and MM2 cores are typical of massive and dense cores harboring high-mass YSO's (Plume et al. 1997; Garay et al. 2007). Those of the MM3 core suggest that it harbors young intermediate mass stars, in accord with the result obtained by Shepherd et al. (2007) from the modeling of the SED of the NIR sources detected using {\\it Spitzer}. We modeled the molecular emission in the \\tco, \\cdo, and \\cssiete\\ lines toward the MM2 core, concluding that this core is undergoing infalling motions, with an infall velocity of $\\sim$1.3 \\kms\\ and a mass infall rate of $\\sim$$1.8\\times 10^{-3}$ \\Msun\\ yr$^{-1}$. This value is large enough to allow the formation of massive stars by accretion (see Osorio, Lizano \\& D'Alessio 1999). We report the discovery of a molecular outflow associated with the massive, dense MM3 core located in the northern region of G34.43+0.24, and possibly of a second one associated with the MM4 core. The molecular outflow within MM3 has a total mass of 41 \\Msun, a momentum of 270 \\Msun\\ \\kms, and a kinetic energy of 1120 \\Msun\\ km$^2$ s$^{-2}$; indicating that it is driven by high--intermadiate mass stars. We also detected high velocity molecular gas toward the MM1 and MM2 cores with total masses of 25 and 80 \\Msun, momenta of 230 and 690 \\Msun\\ \\kms, and kinetic energies of 1290 and 3630 \\Msun\\ km$^2$ s$^{-2}$, respectively." }, "1005/1005.3562_arXiv.txt": { "abstract": "We performed stellar population synthesis on the nuclear and extended regions of NGC~1068 by means of near-infrared spectroscopy to disentangle their spectral energy distribution components. This is the first time that such a technique is applied to the whole 0.8$-$2.4~$\\mu$m wavelength interval in this galaxy. NGC~1068 is one of the nearest and probably the most studied Seyfert 2 galaxy, becoming an excellent laboratory to study the interaction between black holes, the jets that they can produce and the medium in which they propagate. Our main result is that traces of young stellar population are found at $\\sim$100~pc south of the nucleus. The contribution of a power-law continuum in the centre is about 25$\\%$, which is expected if the light is scattered from a Seyfert 1 nucleus. We find peaks in the contribution of the featureless continuum about 100 - 150 pc from the nucleus on both sides. They might be associated with regions where the jet encounters dense clouds. Further support to this scenario is given by the peaks of hot dust distribution found around these same regions and the H$_2$ emission line profile, leading us to propose that the peaks might be associate to regions where stars are being formed. Hot dust also has an important contribution to the nuclear region, reinforcing the idea of the presence of a dense, circumnuclear torus in this galaxy. Cold dust appears mostly in the south direction, which supports the view that the southwest emission is behind the plane of the galaxy and is extinguished very likely by dust in the plane. Intermediate age stellar population contributes significantly to the continuum, specially in the inner 200~pc. ", "introduction": "The coexistence of black holes and starburst clusters is known to exist in many galaxies, and there are many evidences that suggest a connection between these phenomena. Many studies point out that both the active nucleus and starbursts might be related to gas inflow, probably triggered by an axis-asymmetry perturbation like bars, mergers or tidal interactions (Shlosman, Frank \\& Begelman 1989, Shlosman, Begelman \\& Frank 1990, Maiolino et al. 1997, Knapen, Shlosman \\& Peletier 2000, Fathi et al. 2006, Riffel et al. 2008). In addition, one of the most intriguing research areas in contemporary extragalactic astrophysics involves the study of the interplay between nuclear black holes, the jets which they can produce and the interstellar/intergalactic medium (ISM) in which they propagate. These jets can have a considerable impact on this medium. One aspect of jet-ISM interaction is that it can trigger star formation. Such jet-induced star formation is considered a possible mechanism to explain the UV continuum emission observed in the host galaxies of distant radio sources and the ``alignment effect'' between the radio emission and this continuum (Rees 1989). Although this effect might play a very important role in high-z radio galaxies, detecting and studying the jet-ISM interaction in them is very challenging. Because of the observational problems, it is important to find nearby examples of this kind of interaction where a detailed study can be carried out. NGC 1068 is an ideal object in this case. It is one of the nearest and probably the most intensely studied Seyfert 2 galaxy. Observations in all wavelength bands from radio to hard X-rays have formed a uniquely detailed picture of this source. NGC 1068 hosts a prominent narrow-line region (NLR) that is approximately co-spatial with a linear radio source with two lobes (Wilson \\& Ulvestad 1983). Star formation activity coexistent with the active galactic nucleus (AGN) was detected on both larger (e.g., Telesco \\& Decher 1988) and smaller scales (Macchetto et al. 1994; Thatte et al. 1997). However, the link between all the processes is still under debate. The Near Infrared Region (NIR) is particularly interesting to help unveiling this link because it is accessible to ground-based telescopes and, at the same time, able to probe highly obscured sources. However, tracking the star formation in the NIR is not simple (Origlia \\& Oliva 2000) although recent studies exploring this region have already shown its strong potential at detecting intermediate-age stellar population not easily tracked in the optical without ambiguity (e.g. Riffel et al. 2009, Davies et al. 2007). At near-IR wavelengths stellar photospheres usually remain the dominant sources of light, and galaxy spectra are shaped by red supergiants (RSG) shortly after starbursts, and then by giants of the first and of the asymptotic giant branches (AGB). AGB stars are rare members of stellar populations. However, they are among the most luminous cool stars and can therefore be detected sometimes even individually in galaxies. The TP-AGB stars leave a unique fingerprint on the integrated spectra, like the 1.1 $\\micron$ CN band (Maraston 2005, Riffel et al. 2007, 2009). Hence when detected, they can help to determine the age of the stellar population through the integrated light. The contribution of this stellar phase in stellar population models has been recently included in both the energetics and the spectral features (Maraston 2005). In particular these models employ empirical spectra of oxygen-rich and carbon stars (Lan\\c con \\& Wood 2000), which are able to foresee characteristic NIR absorption features. With this in mind, we present here for the first time in the literature a detailed fitting of the continuum emission components in the 0.8$-$2.4 $\\micron$ interval of NGC~1068 across the central 15\" ($\\sim$ 1100 pc) of this source. The main purpose is to determine the fraction with which the different components contributes to the observed integrated light and how they are related to each other. The paper is structured as follows: in \\S 2 we describe the observations. In \\S 3 we describe the fitting method and in \\S 4 the results are presented and discussed. Final remarks are given in \\S 6. ", "conclusions": "We investigated the NIR spectra of NGC~1068 extended across 15\" ($\\sim$ 1100 pc) in the North-South direction. The spectra was taken with the IRTF SpeX instrument, obtained in the short cross-dispersed mode. The objective was to disentangle the galaxy's spectral energy distribution components along the full wavelength coverage (0.8 $\\micron$ - 2.4 $\\micron$) using stellar population synthesis. This was done here for the first time for this galaxy. We used the STARLIGHT code, which considers the whole observed spectrum, continuum and absorption features. We find that intermediate age stellar population contributes significantly to the continuum, specially in the inner 200 pc. We don't have the spatial resolution to determine the size of this nuclear stellar cluster, but results agree with previous determinations of the stellar population of NGC~1068 from Davies et al. (2007) and Riffel et al. (2009). We find a small contribution of a young stellar population component about 100 pc from the nucleus. Although this result has to be considered carefully, since the detection might be an artefact from the method used, taken together with our other findings this suggests that this is the region where the jet interacts with the ISM and might be forming stars. The contribution of a power-law continuum in the centre is about 25$\\%$, which is expected from a scattered Seyfert~1 nucleus (Cid Fernandes et al. 1995). We also find peaks in the contribution of the featureless continuum about 100 - 150 pc from the nucleus, both in the north and in the south direction. These might be associated with the region where the jet encounters dense clouds. This result is further supported by peaks of hot dust found around the same regions. We compared these results with the H$_2$ emission line profile obtained from Paper I. This line also shows extended emission around the same region where the hot dust is found. Taking this together with the young stellar population detection, we believe there is strong evidence that the interaction of the jet with the dense ISM in NGC~1068 is forming stars. Further investigation with more spatial resolution is needed to confirm this hypothesis. Hot dust also has an important contribution to the nuclear region, which reinforces the idea of the presence of a dense, circumnuclear torus in this galaxy. Cold dust appears mostly in the south direction, which supports the idea that the southwest emission is behind the plane of the galaxy, and is extinguished by the dust in the plane. We determine the dust mass in this galaxy to be about 3.4 $\\times$ 10$^{-2}$ M$_{\\sun}$, which puts NGC 1068 as one of the Seyferts with the highest dust content measured." }, "1005/1005.5422_arXiv.txt": { "abstract": "Much of the progress in our understanding of dynamo mechanisms has been made within the theoretical framework of magnetohydrodynamics (MHD). However, for sufficiently diffuse media, the Hall effect eventually becomes non-negligible. We present results from three dimensional simulations of the Hall-MHD equations subjected to random non-helical forcing. We study the role of the Hall effect in the dynamo efficiency for different values of the Hall parameter, using a pseudospectral code to achieve exponentially fast convergence. We also study energy transfer rates among spatial scales to determine the relative importance of the various nonlinear effects in the dynamo process and in the energy cascade. The Hall effect produces a reduction of the direct energy cascade at scales larger than the Hall scale, and therefore leads to smaller energy dissipation rates. Finally, we present results stemming from simulations at large magnetic Prandtl numbers, which is the relevant regime in hot and diffuse media such a the interstellar medium. ", "introduction": "\\label{sec:intro} The generation of magnetic fields by dynamo activity plays an important role in a wide range of astrophysical objects, ranging from stars to clusters of galaxies. The gas in these objects is characterized by turbulent flows, as shown for instance by scintillation observations of the interstellar medium \\cite{Span01, Mint96}, or from pressure maps in galaxy clusters \\cite{Schue04}. Mechanisms able to generate magnetic fields by dynamo action are often classified as large- and small-scale dynamos, depending on the correlation length of the induced magnetic field. In this context, large and small are referred to the energy containing scale of the turbulent hydrodynamic flow. This classification is not rigid, as in many astrophysical objects both dynamos may be at work, but it gives a useful framework considering the limitations in the scale separation that can be achieved in numerical simulations. Also, the physical properties of the flows that can give rise to one or the other are somewhat different. Helical flows have proved efficient in generating large-scale dynamos, i.e., on scales larger than the energy-containing eddies of the flow \\cite{Pouq76,Meneg81,Brand01,Gomez05}. It is now known that large-scale dynamo action can also be produced by anisotropic and inhomogeneous flows (e.g., flows with a large scale shear). On the other hand, non-helical flows can be instrumental in generating small-scale dynamos \\cite{Kazantsev68}, i.e., on sizes smaller than those of the energy-containing eddies \\cite{Scheko01,Scheko04a,Haugen04}. In recent years, the study of small-scale dynamos with magnetic Prandtl number $Pm = \\nu /\\eta $ (the ratio between the viscosity and the magnetic diffusivity of the plasma) different from unity has received special attention \\cite{Scheko04b,Mininnietal05}, both for $Pm \\gg 1$ \\cite{Scheko02} and for $Pm \\ll 1$ \\cite{Ponty05,Iska07}. Motivations to study these regimes include recent experiments of dynamo action using liquid sodium \\cite{Monchaux07}, as well as the fact that many astrophysical plasmas are characterized by magnetic Prandtl numbers different from unity. For instance, the magnetic Prandtl number is much smaller than one in the solar convective region, and it is typically much larger than one in the interplanetary medium and also in the interstellar medium (ISM). For sufficiently low-density media such as the one that pervades the ISM, kinetic effects such as the Hall effect or ambipolar diffusion might also become relevant \\cite{Sano02}. The potential relevance of ambipolar diffusion in astrophysical dynamos was studied in Refs.~\\cite{Branden00,Zweibel02}. The relevance of the Hall effect has been recognized in various astrophysical applications \\cite{Balbus01,Sano02,Mininni02}, space plasmas \\cite{Deng01,Oieroset01,Mozer02}, and also laboratory plasmas \\cite{Yamada97,Mirnov03,Ji08}. The role of the Hall effect on large-scale dynamos subjected to helical forcing has also been addressed in the literature \\cite{Mininni03a,Mininni03b}. Less attention has received the impact of kinetic effects on the small-scale dynamo. A theoretical model of the kinematic small-scale dynamo with Hall effect was presented in \\cite{Kleeorin94}, but to the best of our knowledge no numerical studies of the non-linear and saturated regime were considered in the literature. In this paper, we present results from three dimensional simulations of the Hall-MHD equations subjected to random non-helical forcing. The main aim is to study the role of the Hall effect in the small-scale dynamo efficiency for different values of the Hall parameter. As a result of the study, we also discuss the impact of the Hall effect on the dynamo saturation values, and on magnetic and total dissipation rates. The structure of the paper is as follows. A brief introduction to the theoretical framework known as Hall-MHD is presented in Sect.~\\ref{sec:hmhd}. The role of the Hall effect in the efficiency of the dynamo is shown in Sect.~\\ref{sec:lin}. In Sect.~\\ref{sec:spec} we characterize the stationary regime that is attained when the dynamo process saturates, showing the corresponding energy power spectra. The energy transfer rates participating in the nonlinear energy cascade are displayed in Sect.~\\ref{sec:trans}. In Sect.~\\ref{sec:Pm}, we explore the regime of large magnetic Prandtl number (i.e., when the viscous dissipation scale is larger than the resistive dissipation scale) which, as mentioned, is particularly relevant in diffuse media such as the ISM. Finally, the conclusions are summarized in Sect.~\\ref{sec:conclu}. ", "conclusions": "\\label{sec:conclu} We present results from three dimensional simulations of small-scale dynamo action for magnetic Prandtl numbers $Pm=1$ and $10$ in conducting flows with the Hall effect. This effect is believed to be non-negligible in sufficiently diffuse media, and its relevance has been recognized in various astrophysical, space, and laboratory plasmas. As a first step toward a better description of dynamo action in such media, only the incompressible Hall-MHD equations were solved, and the inclusion of compressible effects as well as other kinetic effects such as ambipolar difussion is left for future studies. However, the inclusion of only the Hall effect acting at the smallest relevant dynamical scales of the flow gives rise to measurable differences with previous studies of dynamo action. A magnetic non-linear regime is identified when the magnetic field (and the current density) becomes large enough to differentiate the electron velocity from the bulk flow velocity. After saturation, differences in the stationary level of magnetic energy and in the total and magnetic energy dissipation rates are obtained, depending on the amplitude of the Hall effect. Finally, the peak of the current density spectrum is found to be dependent on the strength of the Hall term, with its peak moving toward larger scales (smaller wavenumbers) as the Hall scale is increased. By studying the detailed transfer of energy among fields and scales, we observe that the effect of the Hall term is twofold: it transfers energy towards larger scales for scales larger than the Hall length, and it transfers energy towards smaller scales for scales smaller than this length. The modification of the energy flux resulting from this transfer is consistent with the observed changes in the saturation values of energy and dissipation rate observed in our simulations." }, "1005/1005.2202_arXiv.txt": { "abstract": "{The \\object{NGC\\,1999} reflection nebula features a dark patch with a size of $\\sim$10,000\\,AU, which has been interpreted as a small, dense foreground globule and possible site of imminent star formation. We present \\herschel{} PACS far-infrared 70 and 160\\,\\micron{} maps, which reveal a flux deficit at the location of the globule. We estimate the globule mass needed to produce such an absorption feature to be a few tenths to a few \\msun. Inspired by this \\herschel{} observation, we obtained APEX LABOCA and SABOCA submillimeter continuum maps, and Magellan PANIC near--infrared images of the region. We do not detect a submillimer source at the location of the \\herschel\\ flux decrement; furthermore our observations place an upper limit on the mass of the globule of $\\sim$2.4$\\cdot 10^{-2}$\\,\\msun. Indeed, the submillimeter maps appear to show a flux depression as well. Furthermore, the near--infrared images detect faint background stars that are less affected by extinction inside the dark patch than in its surroundings. We suggest that the dark patch is in fact a hole or cavity in the material producing the \\ngc{} reflection nebula, excavated by protostellar jets from the V\\,380\\,Ori multiple system. } ", "introduction": "In the year 1774 C.E.\\ Sir Friedrich Wilhelm Herschel first noticed patches of sky in the constellation Scorpio that were devoid of stars. Unable to find even the faintest star in these regions, his sister Caroline reported him to exclaim: ``Hier ist wahrhaftig ein Loch im Himmel!'' (``Truly there is a hole in the sky here!''). These dark areas are now known to be due to obscuring material \\citep[e.g.,][]{barnard27}: clouds of molecular gas, whose dust content absorbs the light of background stars. Dark clouds, ranging in mass and size from giant molecular clouds (GMCs) to tiny globules, are the sites of star formation in our Galaxy. \\ngc{} is a small reflection nebula in the \\object{LDN\\,1641} portion of the Orion\\,A GMC, located in a small group of 22 pre--main sequence stars and protostars, including the driving source of the prototypical Herbig--Haro objects \\object{HH\\,1} and \\object{HH\\,2} (Megeath et al., in prep.). It is illuminated by the Herbig\\,Ae/Be star \\object{V\\,380\\,Ori}, a multiple system with a circumsystem disk where the primary is a 100\\,L$_{\\odot}$ B9 star exhibiting strong emission lines \\citep[e.g.,][]{leinertetal1997,hillenbrandetal1992,alecianetal2009}. \\ngc{} features a compact (20--30\\arcsec{}) dark patch (see Fig.~\\ref{fig:ngc1999_hst}), which is described in the literature as a dark globule and potential site of star formation \\citep[e.g.,][]{herbig1946,herbig1960,warrensmithetal1980}, but which has never been studied in greater detail before. \\begin{figure} \\begin{center} \\scalebox{0.325}{{\\includegraphics[angle=270]{14612fg1.ps}}} \\caption{\\ngc{} and HH\\,1/2 region (DSS, left) and HST F450W/F555W/F675W true color image of the \\ngc{} dark patch. \\label{fig:ngc1999_hst}} \\end{center} \\end{figure} We report here the serendipitous and surprising detection with PACS of a 70 and 160\\,\\micron{} flux decrement at the location of \\ngc{}. The maps show a dark patch against the nebulous far--IR emission that closely resembles the dark patch seen at visible wavelengths. The MSX, ISO, and Spitzer space telescopes have detected clouds in absorption at mid--IR (5-30\\,\\micron{}) wavelengths against the diffuse IR background of the Galaxy \\citep[e.g.,][]{bacmann00,stutz08,tobin10,stutz09a}, in a few cases even at 70\\,\\micron{} \\citep{stutz09b}. However, the detection of the \\ngc{} globule by PACS would be the first detection of a dark globule in absorption at 160\\,\\micron{}. At wavelengths $\\ge 160$\\,\\micron{}, cold globules are normally detected through the emission from cold (10--30\\,K) dust. Furthermore, absorption at these wavelengths requires extinctions in excess of 100\\,$A_V$. Motivated by the possible discovery of a 160\\,\\micron{} dark cloud, we obtained follow--up ground based submillimeter and near--IR (extinction) observations towards the globule. We did not detect the column density or mass of cold dust necessary to produce such an absorption feature in the far--IR. We therefore conclude that the dark globule is actually a cavity in the \\ngc{} cloud carved by outflows from the V\\,380\\,Ori system. Thus, the \\herschel{} telescope has discovered what truly is a hole in the sky. ", "conclusions": "Optical images of \\ngc{} show a dark patch suggestive of a small, dense globule obscuring the \\ngc{} reflection nebula. Surprisingly, our 70 and 160\\,\\micron{} images clearly show a dark patch against the bright nebula with a strikingly similar morphology to that in visible light images. These \\herschel{} measurements, combined with subsequent ground based data, lead to the conclusion that the dark feature is a hole in the \\ngc{} nebula. First, we find that the masses needed to cause the 70 and 160\\,\\micron{} dark patch --- 0.1 and 2.5\\,\\msun{}, respectively --- are inconsistent. Furthermore, the globule is not detected in emission at 350 and 870\\,\\micron{}: the SABOCA obervations place an upper limit of $2.4\\cdot 10^{-2}$\\,\\msun{} for a temperature of 10\\,K. This upper limit is far below the amount of mass needed to cause the obscuration in the PACS data. Finally, near--IR observations with PANIC detect background stars toward the dark patch; the $H-K_\\mathrm{s}$ colors of these stars are slightly bluer than stars detected outside the globule, suggesting a lower extinction toward the dark patch. Furthermore, the extinctions of the background stars are less than that required to absorb the 160\\,\\micron{} flux. Taken together, these observation show that the dark patch is not a globule, but instead a cavity in the nebula. The presence of a well delineated cavity of the size of $\\sim$10,000\\,AU deserves some attention. With the typical turbulent velocities on the order of a few km/s in clouds, such a cavity should be filled on timescales of at most a few 10,000\\,yrs and quickly disappear. The PANIC H$_2$ narrow band data (Fig.\\,\\ref{fig:NGC1999_H2}) deliver important hints about the possible origin and peculiar shape of the cavity. They reveal a previously unknown, faint H$_2$ bow shock enveloping the SMZ\\,60/HH\\,148 compact knots \\citep{stankeetal2002,corcoranray1995}, constituting the clearest evidence so far for a collimated flow running northeast to southwest through the cavity. This flow, which likely originates in the V\\,380\\,Ori multiple system, could possibly excavate the southern part of the cavity. \\object{SMZ\\,6-8} in the southeastern corner of the PANIC image resembles a small bow shock in a flow coming from the northwest; together with \\object{HH\\,35}, located northwest of V\\,380\\,Ori (Fig.\\,\\ref{fig:ngc1999_hst}), it indicates a second, northwest to southeast oriented flow, which could be responsible for digging the northwestern lobe of the cavity. We note a similar flux depression in the 160\\,\\micron{} \\herschel{} image southwest of the protostar HOPS\\,166, which drives the \\object{HH\\,147} outflow \\citep{corcoranray1995}. Optical images (e.g., Fig.\\,\\ref{fig:ngc1999_hst}) show a circular reflection nebula marking the HH\\,147 outflow cavity, coinciding with the far--IR flux depression. The \\ngc{} dark patch may therefore only be a somewhat peculiar example of a cavity carved in the ambient medium by an outflow, rendered particularly visible by the illumination and heating of the cavity walls by V\\,380\\,Ori. Sensitive far--IR maps taken with \\herschel{} may therefore provide a new tool to assess the importance of outflow feedback on cloud cores." }, "1005/1005.2943_arXiv.txt": { "abstract": "In a series of papers, we propose a theory to explain the formation and properties of rings and spirals in barred galaxies. The building blocks of these structures are orbits guided by the manifolds emanating from the unstable Lagrangian points located near the ends of the bar. In this paper, the last of the series, we present more comparisons of our theoretical results to observations and also give new predictions for further comparisons. Our theory provides the right building blocks for the rectangular-like bar outline and for ansae. We consider how our results can be used to give estimates for the pattern speed values, as well as their effect on abundance gradients in barred galaxies. We present the kinematics along the manifold loci, to allow comparisons with the observed kinematics along the ring and spiral loci. We consider gaseous arms and their relations to stellar ones. We discuss several theoretical aspects and stress that the orbits that constitute the building blocks of the spirals and rings are chaotic. They are, nevertheless, spatially well confined by the manifolds and are thus able to outline the relevant structures. Such chaos can be termed `confined chaos' and can play a very important role in understanding the formation and evolution of galaxy structures and in galactic dynamics in general. This work, in agreement with several others, argues convincingly that galactic dynamic studies should not be limited to the study of regular motions and orbits. ", "introduction": "\\label{sec:intro} In previous papers (\\citealt{RomeroGMAG06}, hereafter Paper I; \\citealt{RomeroGAMG07}, hereafter Paper II; \\citealt{AthaRGM09}, hereafter paper III; \\citealt{RomeroGMGA09}) we proposed a theory to explain the formation and structure of spirals and inner and outer rings in barred galaxy potentials. We propose that the building blocks of these structures are chaotic orbits that are guided by manifolds associated with the Lagrangian points $L_1$ and $L_2$ which are located along the direction of the bar major axis. These manifolds can be thought of as tubes that confine the orbits, so that the latter can form thin structures in configuration space. A theory, however, can be dynamically correct but still irrelevant to a particular application. We therefore have to check whether our theory is applicable to spirals and rings observed in barred disc galaxies. It is thus necessary to compare our theoretical results and predictions to observations. We started this in Paper IV (\\citealt{AthaRGBM09}, hereafter Paper IV) where we made a number of comparisons of model spirals and rings, the latter both inner and outer, to observations and found good agreement. We found that our theory can reproduce all necessary morphologies of both inner and outer rings, and produces no unrealistic morphologies. Model inner rings were found to be elongated along the bar and outer ones either along it ($R_2$, or $R_2'$), or perpendicular to it ($R_1$, or $R_1'$), or both ($R_1R_2$). Model spirals in barred galaxy potentials were found to be predominantly two-armed and trailing, although arms of higher multiplicity are possible for specific potentials. They were found to have the right shape and could reproduce the fall-back of an arm towards the bar region or towards the other arm, which is observed in many spirals. A quantitative comparison to the arm shapes of NGC 1365 proved successful. We predict that the relative strength of the non-axisymmetric forcing in the region around and somewhat beyond corotation influences the winding of the arms, in the sense that in strongly barred galaxies the spirals will be more open than in less strongly barred ones. Thus, a series of models with increasing bar strength will have a continuous sequence of morphologies from $R_1$ to $R_1'$, then to tightly wound spirals, to end with open spirals. We also compared the shape of the inner and outer rings, as well as the ratio of outer-to-inner ring major axes of our models to observations and discussed which type of potentials give best agreement. We also showed that there are correlations, or trends, between all the ratios of ring sizes discussed above and the bar strength. These correlations are very tight if only one type of model potential is used but thicken, sometimes very considerably, if other models, with different properties and analytical expressions, are included. The present paper is the fifth and last of this series. A reminder of the main theoretical results and prerequisites is given in Sect.~\\ref{sec:theory}. This is very brief, since these concepts have been described extensively in Papers I, II and III. Here we will mainly present further comparisons with observations and predictions. These include e.g. the bar shape (Sect.~\\ref{subsec:barshape}), ansae (Sect.~\\ref{subsec:ansae}), radial drift and abundance gradients (Sect.~\\ref{subsec:abundancegr}). Kinematics and behaviour of the line-of-sight velocities are discussed in Sect.~\\ref{sec:Kinematics} and pattern speed prediction is the subject of Sect.~\\ref{sec:omegap}. In Sect.~\\ref{sec:discussion} we discuss specific topics, such as time evolution, self-consistency and spiral structure formation, and we also compare with other spiral structure theories and give outlines for further work. These discussions rely on material from all five papers in this series. Finally, we summarise and conclude in Sect.~\\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} Since the early work of B. Lindblad (\\citeyear{Lindblad63}), density waves have been commonly assumed to be at the basis of any explanation of spiral structure formation. In this series of papers we presented an alternative viewpoint, which we have applied specifically to barred galaxies. This explains the formation of spiral arms, as well as of inner and outer rings, in a common theoretical framework. According to our theory, it is the unstable Lagrangian points located at the ends of the bar and the corresponding manifolds that are responsible for the formation of spirals and rings. These manifolds drive orbits, which are in fact chaotic, but are confined by the manifolds, so that they create over-densities which have the right shape to explain the spirals and the rings (Papers I and II). We showed that weaker non-axisymmetric perturbations will produce manifolds of $R_1$ ring shape, while stronger ones will produce other types of rings and spirals. It is, nevertheless, the same theory that can explain all these different morphologies. This is in good agreement with the observed morphological continuity between rings, pseudorings and spirals. This continuity is clear in our theory, but not in others. We made many different kinds of comparisons with observations. We first made sure that our theory can account for the basic morphological structures observed, namely the spirals, the inner rings and the outer rings (Paper IV). The model outer rings have the observed $R_1$, $R_2$ and $R_1R_2$ ring and pseudo-ring size and morphologies, including the dimples often observed near or at the direction of the bar major axis. Both model and observed inner rings have roughly the same relative sizes and orientations. We next turned to the main spiral arm properties (Paper IV). We showed that, as the bar grows, material gets trapped mainly in the unstable manifolds, i.e. the sense of the arms will be trailing, in good agreement with what is currently known from observations. There is, nevertheless, a very small fraction of the material concerned that will be caught by the stable manifold branch, and will thus form a very low amplitude leading spiral. If its amplitude is not too low, it may be observable in a Fourier analysis of the galaxy image, or as interference patterns on the arm amplitude. Theoretical and observational results agree also well on the number of arms in a barred galaxy (Paper IV). Our theory links one spiral arm to each of the unstable Lagrangian points in the standard case. Since galactic bars are bisymmetric, there should be two such Lagrangian points, the $L_1$ and $L_2$, and therefore two spiral arms. This morphology persists even in the non-standard case where the $L_1$ and $L_2$ are stable (Paper III). Indeed, the vast majority of barred galaxies have two spiral arms. There are, nevertheless, a few barred galaxies with four spiral arms. Our theory can account for such cases, given the right potential, and we discussed a few such potentials in Paper IV. We also compared the shape of observed spiral arms with those of manifolds (Paper IV). We showed that our theory can explain the characteristic arm winding often observed, namely that, as the angle along the arm increases, the radius first increases and then, after reaching a maximum, `falls back' towards the bar. This property has, to our knowledge, not been reproduced by any other theory so far, but follows naturally from ours. We also predict that spirals in galaxies with stronger bars will be more open than those in galaxies with weaker bars. This prediction can be made because our theory, contrary to most others, is non-linear. Finally, very tightly wound near-logarithmic spirals can also be obtained with our theory, but will rely considerably on spiral forcing. The shape of rings and the ratio of their extents, although not as straightforward to compare as other properties, reveal interesting information (Paper IV). Concerning $R_1$ rings, for which we have sufficient models to get adequate statistics, we find that the ring shapes cover the same range as observations. A comparison for inner rings is less straightforward, because in strong bar cases parts of the inner manifold branches outline parts of the bar and not an inner ring. Taking this into account, we find that there is a fair agreement between models and observations. One of the predictions made in Paper IV concerns the shape of the inner rings. Namely, we found a strong correlation between the ring shape and $Q_{t,L_1}$, in the sense that bars with a stronger forcing at or somewhat beyond $L_1$ should have more elongated $R_1$ rings. We also stressed that, in order for correlations concerning a spiral or a ring property with the strength of the non-axisymmetric forcing to be strong, the latter quantity should be measured at or beyond the Lagrangian radius, i.e. with $Q_{tL_1}$. If $Q_b$ is used instead, the correlation can be destroyed, or severely damaged. Both of these results have now been confirmed by observations \\citep{GrouchyBSL10}. We also introduced collisions and dissipation to the manifold calculations, in order to roughly model the gas properties (Paper IV). We found that this does not influence the existence of the spiral arms or rings, and makes no major modifications to their shape. The amount of dissipation, however, does influence the width of the arms. These become thinner as the dissipation is increased, so that the gaseous arm comes nearer to the lowest energy manifold. Our theory makes very clear and, for many morphological types, precise predictions about the value of the bar pattern speed, since it shows how it is possible from the morphology only to locate the positions of the $L_1$ and $L_2$, both for spirals and for the various ring morphologies (Paper V). This of course works only in the case where sufficient morphological features are present in the observed galaxy. For example if there is only a bar, with no spirals or rings, our theory can not be applied and the $L_1$ and $L_2$ can not be located. However, in cases with a reasonable amount of structure our theory allows the location of the $L_1$ and $L_2$ from morphology alone. In cases with appropriate morphological features this may allow a more accurate determination of the pattern speed then other methods used so far. Our theory also provides building blocks to explain the rectangular-like outline observed in strong barred galaxies and to explain ansae (Paper V). Concerning photometry, our results are not very constraining. The amplitude of the spiral should in general decrease outwards, but this decrease may well not be monotonic, because other arm components -- e.g. a weak, leading component --, if present, will lead to bumps on an otherwise smoothly decreasing density profile. The circulation of material within the manifolds also depends on the bar strength. In the relatively weak bars, which form $rR_1$ morphologies, the mass elements can move from the inner manifolds to the outer manifolds and then back to inner ones again. I.e. material can move from the region within corotation to the region outside it and vice versa via the neighbourhood of the $L_1$ and $L_2$. Averaged over time, this brings internal circulation, but no net motion of material inwards or outwards. The situation is more complex for the cases where the non-axisymmetric forcing around corotation is stronger and which have a spiral morphology. In such cases, material moves from the region within corotation to the region outside it. If the potential has a predominantly barred structure, this material returns eventually to the region of the bar or of the other arm. But if the potential has also a spiral component, then the arm can continue winding outwards for much longer times. Either way, this results in a considerable radial mixing and material can reach the outermost parts of the disc and can contribute to its radial extension. Of course the density wave theory had already predicted that in a spiral galaxy the radial motion is outwards outside corotation \\citep{Lin.Shu71, Kalnajs78}. Our results, however, are more than a simple restatement of the above, because our theory shows how material from well within corotation can reach regions well outside it. It can thus explain considerable radial mixing and have a more important impact on the formation of the outer disc than those of the density wave theory. Furthermore, our theory predicts the relation between bar strength and the abundance gradients found in observations \\citep{Martin.Roy94}. We also made predictions on the galaxy kinematics, by measuring the radial and tangential velocity components along the manifold loci, i.e. along the spirals and the inner and outer rings. For the tangential component we subtract the local circular velocity and thus obtain the perturbation, or peculiar tangential velocity. Plotted as a function of the azimuthal angle along the ring, both radial and tangential velocities show $m$ = 2 oscillations, the amplitude of the former being higher than that of the latter. We find that the amplitude of these oscillations increases with increasing bar strength and that they are larger in the inner than in the outer ring. The latter result is in agreement with the former, since the bar forcing is higher at the inner ring than at the outer one. For spirals, we find that the radial velocity of material moving from the Lagrangian points outwards along the arm increases with the distance to the Lagrangian point until it reaches a maximum and then it decreases, until at some point it changes sign. From that point onwards the arm will start approaching the other arm or the region of the bar. Where this occurs is a function of the bar strength and, when observed, will provide an additional confirmation to our theory. Nevertheless, this may be a difficult observation to make, since the spiral density is low in that region. Furthermore, in galaxies with a high spiral amplitude, the spiral forcing may be a considerable fraction of the non-axisymmetric forcing, so that this sign reversal may not take place. Finally, we calculated the line-of-sight velocity and found that in the case of rings its variation along the ring can be well modelled by a sinusoidal. This is not true for very strong bars and spiral morphologies. Finally, when the sinusoidal is a reasonable fit, we find that the difference between the kinematic and photometric major axes is in agreement with the observed values. As already mentioned, our models show a clear connection between bar strength and morphology of the response. Relatively weak non-axisymmetric forcings give $R_1$ and $R_1'$ morphologies, while stronger forcings give spirals and other types of rings. Yet it is possible in self-consistent calculations, in which the potential due to the material trapped in the manifold is taken into account, to form a spiral in cases where the bar on its own is too weak to sustain it. This is due to the fact that initially the only part of the ring that will be populated is the part which is near the $L_1$ or the $L_2$. This then adds a spiral-like forcing to that of the bar, so that manifold will have a more spiral-like shape. It is thus possible in self-consistent calculations to have a spiral morphology in a bar potential which is somewhat below the threshold between the $R_1$ and the spiral morphology in the rigid bar forcing case. In these series of papers, we applied our theory to barred galaxies. It can, however, be straightforwardly applied also to non-barred galaxies with other non-axisymmetric perturbations, such as spirals or ovals, provided these have unstable saddle Lagrangian points. There is thus no discontinuity between barred and non-barred spiral galaxies. On the other hand it is unclear at this point whether, and under what conditions, complex unstable Lagrangian points can have the type of manifolds that can create realistic spirals or rings. Such Lagrangian points are found e.g. at the $L_4$ and $L_5$ Lagrangian points of very strong bars \\citep[][and Paper IV]{Atha92a}. More work is necessary before this question can be answered. We discussed how material can get trapped in the manifolds, where the mass in the spirals comes from and whether the spirals are short-lived or long-lived. We also compared our theory with other theories not relying on chaotic orbits and manifolds. In particular, we discussed a number of advantages which our theory has over the density wave theory. We stressed that it need not necessarily be the same theory that explains all spirals in disc galaxies, and even in one single galaxy more than one mechanism can be at play. The properties of the resulting spiral would then result from the self-consistent interaction of all the spirals of different origin. We finally propose a number of avenues for future work within the framework of our theory. We showed that the orbits that form the building blocks of the spirals and of the inner and outer rings are chaotic. Nevertheless, they can form narrow features because they are confined by the manifolds, which act as guiding tubes. We propose to call this type of chaos `confined chaos' and suggest that such orbits should be taken into account in dynamical studies, since they could contribute, together with the regular orbits trapped around the stable periodic orbits, to galactic structures. A prerequisite for any theory is that it can be falsified \\citep[e.g.][]{Popper59}. Ours fulfils well this criterion. Indeed there are a number of observational results that could have proved it wrong. Namely, if the two armed global spirals were not preponderant in symmetric barred galaxies, or if these arms were leading, or if the major features of the observed and the theoretical morphologies were inconsistent, our theory would have been proven wrong. As we showed in this paper, this is not the case and there is good agreement between our theory and observations in these and many other points. There are more observations which could have proven our theory to be wrong. For example, our theory would be invalidated if it was clearly shown that all spirals rotate with a different pattern speed than the bar. In total over the five papers, we presented a number of predictions of our theory. Whenever the necessary data are available, we compared the results of our theory to the existing observational results. All tests we have been able to make tend to show that our theory is viable. In many cases, however, new observations, or new analysis of existing data are necessary. In such cases, we presented here simply the theoretical predictions and outlined the way the comparisons could be made. The results of these comparisons will be very important in order to confirm our theory, to bring about modifications, or to reject it. As already mentioned, one of the theoretical predictions of paper IV concerning the dependence of the inner ring shape on bar strength has already been confirmed by observations \\citep{GrouchyBSL10}. We can thus conclude that the theory we presented in this series of papers should be useful in order to explain the formation of spirals and of inner and outer rings in barred galaxies. Concerning rings, no other theory has been fully developed, while our results agree very well with those of simulations, suggesting that the motion of particles in them are guided by the manifolds we describe here. Concerning spirals, the comparisons we have made with observations are more extensive than those that have been made so far for the swing amplifier theory. Furthermore, our theory is clearly falsifiable, contrary to the classical WKBJ density wave theory (see discussion by \\cite{Kalnajs78}). We thus believe that it should be considered as a possible alternative, on a par with other theories. More work, both theoretical and observational, is necessary in order to establish in what cases each one of the so far proposed theories prevails and how they can, perhaps all together, come to an explanation of spiral formation and spiral properties." }, "1005/1005.2458_arXiv.txt": { "abstract": "The conditions of thermodynamic equilibrium, local thermodynamic equilibrium, and statistical equilibrium are discussed in detail. The equations of statistical equilibrium and the supplementary equations are shown together with the expressions for radiative and collisional rates with the emphasize on the solution for trace elements. ", "introduction": "Solution of the stellar atmosphere problem is usually considered as the determination of spatial dependence of basic macroscopic quantities \\citep[see also][\\citetalias{ja1}]{ja1} also be understood \\citep[following][]{ivandis} as aiming at finding three basic microscopic distributions, namely, the momentum distribution (distribution of velocities of all particles), distribution of particle internal degrees of freedom (populations of atomic excitation stages), and distribution of internal degrees of freedom of the electromagnetic field (radiation field for all frequencies and directions). ", "conclusions": "In this review we summarized the basic equations for the solution of the NLTE line formation problem with a focus on the solution of the problem for trace elements in stellar atmospheres. It has to be emphasized that in NLTE calculations for trace elements preference should be given to background NLTE model atmospheres, since using LTE model atmospheres is inconsistent with the condition of NLTE. It must be carefully decided which elements can be treated as a trace element and for which such treatment is questionable. After obtaining the numerical results all assumptions have to be verified, especially that of the traceness." }, "1005/1005.0388_arXiv.txt": { "abstract": "Ongoing and upcoming surveys in x-rays and SZE are expected to jointly detect many clusters due to the large overlap in sky coverage. We show that, these clusters can be used as an ensemble of rulers to estimate the angular diameter distance, \\da. This comes at no extra observational cost, as these clusters form a subset of a much larger sample, assembled to build cluster number counts \\dn. On using this \\da, the dark energy constraints can be improved by factors of 1.5 - 4, over those from just \\dn. Even in the presence of a mass follow-up of 100 clusters (done for mass calibration), the dark energy constraints can be further tightened by factors of 2 - 3 . Adding \\da from clusters is similar to adding $d_{\\rm L}(z)$, from the SNe observations; for eg., \\dn (from ACT/SPT) plus \\da is comparable to \\dn plus $d_{\\rm L}(z)$ in constraining $\\Omega_m$ and $\\sigma_8$. ", "introduction": "Large cluster surveys like the SPT, ACT, Planck and eROSITA promise to detect from a thousand to a few hundred thousand clusters in the coming decade. The abundance and redshift distribution $\\frac{dN}{dz}$ of these clusters are important probes to understand the nature of dark energy as well as to constrain other cosmological parameters like $\\Omega_m$ and $\\sigma_8$ \\citep{Holder01, WangSteinhardt, Levine, Weller02, Hu}. To deduce a cosmology from these $\\frac{dN}{dz}$ observations, one requires a precise knowledge of the limiting mass of the survey as a function of redshift. One frequently uses proxy observables such as X-Ray surface brightness and temperature \\citep{Ebeling00}, Sunyaev Zel'dovich effect (SZE) decrement \\citep{Staniszewski, Hinks}, cluster richness \\citep{Postman, Koester} and lensing \\citep{Wittman06, Zitrin09} for the masses of clusters, related through simple power-law scaling relations. These scaling parameters are highly degenerate with the cosmological parameters, and breaking this degeneracy is crucial to obtain tight constraints on cosmology. This may realized, for example, through the so called `self-calibration' techniques \\citep{MM04, Hu, LimaHu04}. Other approaches include an `unbiased' mass follow-up of a sub-sample of the survey clusters \\citep{MM03,MM04} or better theoretical modeling of clusters to predict the form of mass-observable scaling relation \\citep{YoungerHaiman, Reid06, CM09}. One can also try to optimize the cluster surveys so as to get the best possible survey yield \\citep{SatejPRL, Battye}. Measurement of the angular diameter distance, $d_{\\rm A}$, at the redshift of the cluster using a combination of SZE and X-Ray observations have been routinely made over the last 30 years. The results have suffered, in the past, from various systematics and reliable estimates have only been achieved recently with analysis of statistically significant samples of galaxy clusters \\citep{Reese02,Bonamente}. These new observations have demonstrated the power of using clusters to measure `$d_{\\rm A}$ vs $z$' and use it to study the expansion history of our Universe \\citep{Molnar}. However, these recent progress has been done with targeted observations. Since, targeted observations are costly, this approach limits the size of the sample of $\\lsim 100$. In this {\\it letter}, we show that one can build up an `ensemble' of \\da by picking a sub-sample of clusters discovered in both X-Ray and SZE surveys with overlapping area and redshift coverage. Since, the surveys are already geared towards getting clusters for $\\frac{dN}{dz}$, we get the \\da without any {\\it extra} targeted observations. Addition of \\da to \\dn helps in tightening cosmological constraints, especially on dark energy equation of state. This is not surprising, since using \\da from clusters is akin to adding $d_{\\rm L}(z)$ information from supernovae (SNe) observations. The rest of this paper is organized as follows. In \\S~2, we briefly discuss cluster number counts and the procedure to estimate \\da from SZE + X-Ray observations; in \\S~3 we describe the surveys, the choice of cosmological and cluster models and also summarize our methodology; in \\S~4, we forecast constraints on cosmological parameters; and finally, we conclude in \\S~5. ", "conclusions": "We show that just using number count observations from the ongoing SZE galaxy cluster surveys like ACT/SPT would not be sufficient to put tight constraints on the cosmological parameters, especially the dark energy parameters. However these constraints can be significantly improved by doing a joint analysis with \\da. These $d_{\\rm A}(z)$ can be constructed out of clusters detected jointly in X-Ray and SZE observations having overlapping areas. The optical follow-up providing the redshifts for \\dn will, also, naturally provide the redshifts for \\da. With the current and upcoming large area surveys one will be able to get thousands of clusters providing us with us with $d_{\\rm A}$ at various redshifts without much effort. We find that adding \\da to number count observations always improves the dark energy constraints, from \\dn alone by {\\it factors of 1.5 to 4}. This leads to better constraints not only on dark energy but also on the parameter $\\Omega_m$. Even when a targeted mass follow-up of clusters helps in breaking the cluster-cosmology degeneracies, addition of \\da helps in further tightening of the cosmological constraints. Moreover addition of \\da improves dark energy constraints comparable to the improvement brought by adding CMB priors on $\\sigma_8$ and $\\Omega_m$. Thus, our proposal of adding the \\da data to cluster number counts provides a natural way of improving the cosmological constraints using clusters alone. The authors would like to thank Gil Holder, Jonathan Dudley and Joe Mohr for many discussions during the work." }, "1005/1005.2344_arXiv.txt": { "abstract": "Nuclear hard X-ray luminosities (\\ln\\ba) for a sample of 112 early type galaxies within a distance of $67$ Mpc are used to investigate their relationship with the central galactic black hole mass \\mb (coming from direct dynamical studies or the \\mb$-\\sigma$ relation), the inner galactic structure (using the parameters describing its cuspiness), the hot gas content and the core radio luminosity. For this sample, \\ln ranges from $10^{38}$ to $10^{42}$ erg s$^{-1}$, and the Eddington ratio \\ln\\ba/\\ledd from $10^{-9}$ to $10^{-4}$, with the largest values belonging to four Seyfert galaxies. Together with a trend for \\ln to increase on average with the galactic luminosity \\lb and \\mb\\ba, there is a wide variation of \\ln (and \\ln\\ba/\\ledd\\ba), by up to 4 orders of magnitude, at any fixed \\lb$> 6\\times 10^{9}L_{B,\\odot}$ or \\mb$>10^7M_{\\odot}$. This large observed range should reflect a large variation of the mass accretion rate \\md, and possible reasons for this difference are searched for. On the circumnuclear scale, in a scenario where accretion is (quasi) steady, \\md$\\,$ at fixed \\lb (or \\mb\\ba) could vary due to differences in the fuel production rate from stellar mass return linked to the inner galactic structure; a trend of \\ln with cuspiness is not observed, though, while a tendency for \\ln\\ba/\\ledd to be larger in cuspier galaxies is present. In fact, \\md$\\,$ is predicted to vary with cuspiness by a factor exceeding a few only in hot gas poor galaxies and for large differences in the core radius; for a subsample with these characteristics the expected effect seems to be present in the observed \\ln values. \\ln does not show a dependence on the age of the stellar population in the central galactic region, for ages$>$3 Gyr; less luminous nuclei, though, are found among the youngest galaxies or galaxies with a younger stellar component. On the global galactic scale, \\ln shows a trend with the total galactic hot gas cooling rate ($L_{X,ISM}$): it is detected both in gas poor and gas rich galaxies, and on average increases with $L_{X,ISM}$, but again with a large scatter. The observed lack of a tight relationship between \\ln and the circumnuclear and total gas content can be explained if accretion is regulated by factors overcoming the importance of fuel availability, as 1) the gas is heated by black hole feedback and \\md$\\,$ varies due to an activity cycle, and 2) the mass effectively accreted by the black hole can be largely reduced with respect to that entering the circumnuclear region, as in radiatively inefficient accretion with winds/outflows. Finally, differently from \\ln\\ba, the central 5 GHz VLA luminosity shows a clear trend with the inner galactic structure, that is similar to that shown by the total soft X-ray emission; therefore it is suggested that they could both be produced by the hot gas. ", "introduction": "In the past years, high angular resolution studies of the centers of early type galaxies have been performed with the $Hubble$ $Space$ $Telescope$ ($HST$) in the optical and near infrared, and in the X-rays with $Chandra$, obtaining important results that deeply influenced our understanding of the nature and past evolution of these systems. The first, major $HST$ result was that massive black holes (MBHs) are ubiquitous in the centers of spheroids, and linked by tight relationships with the luminosity and central stellar velocity dispersion of their hosts (e.g., Magorrian et al. 1998, Ferrarese \\& Merritt 2000, Gebhardt et al. 2000), indicative of a strong mutual influence during their formation and evolution. The second $HST$ result was that the central brightness profiles of galaxies with $M_V\\lsim -19$ show either steep brightness cusps or, interior to a break radius $r_b$, they flatten markedly in a core with respect to an inner extrapolation of the outer profile. These profiles have been described respectively by a \\Ser or core-\\Ser law (Graham et al. 2003, Trujillo et al. 2004, Ferrarese et al. 2006, Kormendy et al. 2009) or alternatively by the \"Nuker law\" (Faber et al. 1997, Lauer et al. 2007a). Cores dominate at the highest luminosities and steep cusps at the lowest, with an intermediate luminosity region of coexistence ($-20.5\\gsim M_V \\gsim -23$, Lauer et al. 2007a). The shape of the brightness profile in the inner galactic region has been related to the past formation and evolution of galaxies, with cores created during dry merging events by a black hole binary ejecting stars from the center of the new system (Ebisuzaki et al. 1991, Faber et al. 1997, Milosavljevic et al. 2002, Graham 2004, Gualandris \\& Merritt 2008), and cusps being preserved or (re)generated during gaseous (wet) mergings. Recently, it was found that coreless elliptical galaxies in the Virgo cluster have extra-light at their center, above the inward extrapolation of their outer \\Ser profile (Kormendy et al. 2009), the result of a wet merger induced starburst (see also Hopkins et al. 2009a) or of AGN induced starburst activity (Ciotti \\& Ostriker 2007). Moreover, the presence of steep cusps or cores correlates with other fundamental galactic properties, even more tightly than how these properties correlate with the galatic luminosity: core galaxies generally have boxy isophotes, are slow rotators and triaxial systems, while cusp galaxies are disky, fast rotators and axisymmetric (Kormendy \\& Bender 1996, Faber et al. 1997); core galaxies show a large range of radio and X-ray luminosities, while cusp galaxies are confined below a threshold (Bender et al. 1989; Pellegrini 1999, 2005a; Capetti \\& Balmaverde 2005; Pasquali et al. 2007). Pellegrini (2005a) also attempted an investigation of the relation between the X-ray nuclear emission (\\ln) and the inner core/cusp profile, but the study was limited by the small number of nuclei with known \\ln available. Launched in 1999, the $Chandra$ satellite has now pointed a large number of early type galaxies, with an unprecedented angular resolution in the X-rays of less than $1^{\\prime\\prime}$. For the first time measurements of the nuclear X-ray emission down to values as low as $10^{39}$ erg s$^{-1}$ and out to distances of $\\sim 60$ Mpc have been obtained. The MBHs of the local universe turned out to be typically radiatively quiescent and very sub-Eddington emitters (Loewenstein et al. 2001, Soria et al. 2006a, Zhang et al. 2009, Gallo et al. 2008, 2010). In a number of cases the mass accretion rate on the MBH could be estimated (e.g., Di Matteo et al. 2003, Pellegrini 2005b) and the radiative quiescence was interpreted in terms of radiatively inefficient accretion (Narayan \\& Yi 1995), possibly with the mechanical power dominating the total output of accretion (e.g., Allen et al. 2006). However, many aspects of accretion in the local universe remain unknown: what determines \\ln ? Is there any relation of \\ln or its Eddington-scaled value with the galactic luminosity or the mass of the central supermassive black hole \\mb\\ba? or with the inner stellar profile, that has been linked to the past galactic evolution and other major global galactic properties? or with the hot gas content? Answering these questions is important for a complete understanding of the MBH-host galaxy coevolution process. In this work we have collected all early type galaxies (E and S0) out to a distance of $\\sim 67$ Mpc with known \\ln\\ba, based mostly on data coming from $Chandra$ pointings; a total of 112 galaxies resulted with \\ln measured or with an upper limit on it. The sample includes also most of the early type galaxies with a direct measurement of \\mb via dynamical studies currently available; for the other galaxies, the central stellar velocity dispersion allows for an estimate of \\mb via the \\mb$-\\sigma$ relation. For 81 of these 112 galaxies the central stellar profile shape has been measured with $HST$. The sample is described in Sect.~\\ref{sample}, the observational evidences about relationships between \\ln\\ba, \\mb\\ba, the central stellar structure and the radio luminosity are presented in Sect.~\\ref{obs}, the results are discussed in Sect.~\\ref{disc} (that examines also the relationship between \\ln and the galactic hot gas luminosity); the conclusions are summarized in Sect.~\\ref{concl}. ", "conclusions": "In this work measurements or upper limits have been collected for the hard X-ray emission at the nucleus of 112 early type galaxies (E and S0) of the local universe (within 67 Mpc). \\ln derives from $Chandra$ data for 94\\% of the nuclei, and the sample includes all the available measurements of \\ln (29 cases) for galaxies with the above characteristics and a direct estimate of the \\mb (38 objects). Using also \\mb from the \\mb$-\\sigma$ relation, and the inner stellar profile (slope $\\gamma$ and break radius $r_b$) measured with $HST$ for 81 galaxies, the relationships between \\ln and \\mb\\ba, the central stellar structure, the central age, the radio and the soft X-rays (hot gas) luminosities have been investigated, with the following results: $\\bullet$ \\ln increases on average with \\lb and \\mb\\ba, with a wide variation of \\ln and \\ln\\ba/\\ledd\\ba, up to 4 orders of magnitude, at any fixed galactic \\lb$> 6\\times 10^{9}L_{B,\\odot}$ or \\mb$>10^7M_{\\odot}$. Cusp and core galaxies both cover the whole large range of \\ln\\ba, without a clear trend with $\\gamma$ or $r_b$. Most nuclei have an Eddington ratio $-9<$log(\\ln\\ba/\\ledd\\ba)$<-5$, the brightest ones (with log(\\ln\\ba/\\ledd\\ba)$\\sim -4$) are four Seyfert nuclei. The \\ln\\ba/\\ledd is highest at the lowest \\mb\\ba, and shows an increase of its highest values with increasing \"cuspiness\" ($\\gamma$). $\\bullet$ Accretion in these MBHs may have entered the radiatively inefficient regime, where \\ln scales as the mass accretion rate \\md$^2$ at fixed \\mb\\ba; therefore reasons for \\md$\\,$ variations explaining the wide range of observed \\ln at fixed \\lb (or \\mb\\ba) are searched for. In a scenario where accretion is (quasi) steady, \\md$\\,$ could vary because of a different inner galactic structure, that determines the amount of mass shed by stars in the circumnuclear region and then available for accretion. The lack of a clear trend of \\ln with $\\gamma$ and $r_b$, and the weak evidence for an increase of \\ln\\ba/\\ledd with cuspiness, indicate that differences in the inner galactic structure do not have the dominant effect on \\ln\\ba. In fact the circumnuclear fuel production rate is estimated to vary more than a factor of a few only when the inflowing region is small (of the order of the accretion radius) and for large differences in $r_b$; in agreement with this, for galaxies with presumably a small inflowing region, a trend for \\ln to decrease for increasing $r_b$ seems to be present. $\\bullet$ The stellar mass return rate depends also on the age of the stellar population, but \\ln covers the same wide range of values at all central ages from 3 to 14 Gyr; \"youth\", indicated by a younger center, recent starformation, and/or a younger stellar component at the center (of ages $<2-3$ Gyr), seems to be more connected with a lower \\ln\\ba. An explanation could be that the heating by starformation or the dynamical effect of the merging process on the hot gas drive the available gas out in a wind, ending abruptly the starformation and accretion. However, this finding remains an indication, because of the trend of \\ln with \\lb and the small number of galaxies with information about recent starformation episodes. $\\bullet$ \\ln on average increases with the total hot gas emission $L_{X,ISM}$, but the relation is not tight: it shows a large variation of 2--3 orders of magnitude both at the lowest and highest gas contents. Nuclear emission at the highest detected levels for this sample (\\ln$\\sim 10^{40}-10^{41}$ erg s$^{-1}$) is present even when the gas content is low, and \\ln at the lowest levels ($\\sim 10^{38}-10^{39}$ erg s$^{-1}$) can be found at gas luminosities differing by 3 orders of magnitude. $\\bullet $ The mild sensitivity of \\ln to the circumnuclear and global hot gas contents, and its large variation, finds two possible explanations, both of which go in the sense of overcoming the importance of fuel availability : 1) the gas is heated due to feedback from the MBH, which could be a phenomenon not limited to high $L_{X,ISM}$ galaxies, or 2) only a small fraction of the mass entering the accretion radius actually reaches the MBH, as in RIAF models with winds/outflows, and this fraction can vary largely at fixed \\mb\\ba. $\\bullet $ A sub-sample of galaxies shows an already known trend by which cusp galaxies are confined below a threshold in the VLA 5 GHz central luminosity $L_{R,core}$, while core galaxies span a large range of $L_{R,core}$; this subsample does not show a similar trend of \\ln with the central light profile. The $L_{R,core}$ behavior is instead similar to that of the total soft X-ray emission with respect to $\\gamma$; the hot gas could then be responsible both for the soft X-ray emission and the jet confinement and propagation. While core galaxies can possess a conspicuous radio activity cycle, in cusp galaxies the variation of the radio emission keeps within a smaller range, because of a rapid jet failure due to the lack of a dense confining medium, or a smaller duty cycle, being these galaxies on average less massive systems." }, "1005/1005.1171_arXiv.txt": { "abstract": "% The effect of massive neutrinos on matter power spectrum is discussed in the context of $f(R)$ gravity. It is shown that the anomalous growth of density fluctuations on small scales due to the scalaron force can be compensated by free streaming of neutrinos. As a result, models which predict observable deviation of the equation-of-state parameter $w_\\DE$ from $w_\\DE=-1$ can be reconciled with observations of matter clustering if the total neutrino mass is $O(0.5~\\eV)$. ", "introduction": " ", "conclusions": "" }, "1005/1005.1940_arXiv.txt": { "abstract": "When studying the evolutionary stages of protostars that form in clusters, the role of any intracluster medium cannot be neglected. High foreground extinction can lead to situations where young stellar objects (YSOs) appear to be in earlier evolutionary stages than they actually are, particularly when using simple criteria like spectral indices. To address this issue, we have assembled detailed SED characterizations of a sample of 56 \\textsl{Spitzer}-identified candidate YSOs in the clusters NGC~2264 and IC~348. For these, we use spectra obtained with the Infrared Spectrograph onboard the \\textsl{Spitzer} Space Telescope and ancillary multi-wavelength photometry. The primary aim is twofold: 1) to discuss the role of spectral features, particularly those due to ices and silicates, in determining a YSO's evolutionary stage, and 2) to perform comprehensive modeling of spectral energy distributions (SEDs) enhanced by the IRS data. The SEDs consist of ancillary optical-to-submillimeter multi-wavelength data as well as an accurate description of the 9.7~$\\mu$m silicate feature and of the mid-infrared continuum derived from line-free parts of the IRS spectra. We find that using this approach, we can distinguish genuine protostars in the cluster from T Tauri stars masquerading as protostars due to external foreground extinction. Our results underline the importance of photometric data in the far-infrared/submillimeter wavelength range, at sufficiently high angular resolution to more accurately classify cluster members. Such observations are becoming possible now with the advent of the \\textsl{Herschel} Space Observatory. ", "introduction": "Finding and characterizing individual Young Stellar Objects (YSOs) in embedded clustered environments is challenging because the observed YSO spectral energy distributions (SEDs) can easily be influenced by dense material that is distributed throughout the cluster, but physically unrelated to an individual YSO or its envelope. In that situation, the extra extinction can mislead the observer since an affected source may appear to be in an earlier evolutionary stage than it actually is. For example, the SED of a T Tauri star behind foreground extinction can in some circumstances look protostellar. In a more isolated environment, this situation would only occur in case of edge-on disks, aggravated by the presence of disk flaring. To avoid confusion, \\citet{rob06} proposed that sources be assigned designations that refer to the actual evolutionary stages, rather than spectral classes defined by observations, since the latter are affected by extinction. In most instances, the evolutionary stages (I, II, III) correspond to the SED classes. However, in heavily extincted regions the SED classes may not always correspond to the appropriate evolutionary stages. In embedded clustered environments, the difference between YSO classes and evolutionary stages becomes particularly important. Infrared spectroscopy allows us to go beyond protostellar classifications based on photometry. At near-infrared wavelengths, varying degrees of veiling can be used to distinguish protostars from PMS stars \\citep{cas92,gre96,whi07}. At mid-infrared wavelengths, silicate dust and ices cause telltale spectral features that can be used to refine spectral classes. Many of these features occur in the wavelength range that was accessible with the Infrared Spectrograph (IRS) onboard the \\textsl{Spitzer} Space Telescope \\citep{hou04}. In the context of star formation, the power of this instrument has been shown most notably by studies of T Tauri stars and protostars in Taurus, Chamaeleon, and Ophiuchus \\citep{fur06,fur08,fur09}. Most recently, \\citet{mcc10} study an extensive sample of YSOs in Ophiuchus, using IRS data to determine their evolutionary stages. To find out whether the effect of intracluster or external foreground extinction can be disentangled from the actual evolutionary stage of a YSO, we have obtained IRS spectra and assembled detailed multi-wavelength photometry for a sample of 56 candidate YSOs in NGC~2264 and IC~348. We then use these data for comprehensive SED modeling with the primary aim of ascertaining the evolutionary stages for the sources in our sample. These can then finally be compared to previous classification attempts that are mainly based on less broad mid-infrared wavelength coverage. For each of the 56 sources, we have obtained mid-infrared spectra using \\textsl{Spitzer}-IRS. These data not only allow us to study spectral features that trace the evolutionary stage of a YSO (silicates and ices), but they also allow us to derive photometry from line-free parts of the spectrum to accurately reflect the underlying continuum. This technique also allows us to accurately describe the main silicate spectral feature by photometric data at selected wavelengths. Combined with ancillary photometric data spanning optical to submillimeter wavelengths, the IRS data thereby allow us to construct SEDs that are tailored toward the requirements of detailed modeling. The accuracy of the modeling depends on the SEDs not containing spectral lines that are not part of the underlying model (e.g., PAH features); these features can fall into the band passes of broadband filters. Our IRS-derived photometry attempts to depict the continuum and the silicate features. Compared to the usual approach of using photometry from \\textsl{Spitzer}-IRAC in SEDs for subsequent modeling, we discard bands 3 and 4 of the IRAC data that we have for every source in favor of corresponding line-free photometry derived from the IRS data. \\subsection{Source selection} We selected 56 candidate protostars in NGC 2264 and IC\\,348 -- 40 in the former and 16 in the latter region -- from previous \\textsl{Spitzer} photometric studies of these regions by \\citet{mue07} and \\citet{tei08}. Our target sources are listed in Table~\\ref{tab_srclist}. NGC 2264 is a prominent, clustered star-forming region located at a distance of $\\sim$913~pc \\citep{bax09}; for a recent review, see \\citet{dah08}. Here, we focus on the population of embedded protostars in the vicinity of IRS~2 \\citep{you06} where we selected 40 sources for \\textsl{Spitzer}-IRS mid-infrared spectroscopic observations. This region is one of the two most prominent regions of ongoing star formation in NGC~2264, next to IRS~1. All selected sources are located in the \\textsl{Spokes} cluster \\citep{tei06}. Figure~\\ref{mips_n2264} shows the selected sources in a mid-infrared view of the region. IC\\,348 is located in Perseus at a distance of $\\sim$260~pc \\citep[e.g.,][]{lom10}; for a recent review, see \\citet{her08}. \\citet{mue07} conducted a \\textsl{Spitzer} census of this star-forming region. The IC\\,348 sources discussed in this paper are among the most embedded in the cluster and are predominantly protostars in a narrow filamentary ridge located about 1~pc southwest of the central cluster. In Figure~\\ref{mips_ic348}, we mark the selected sources in a mid-infrared view of the region. A first comparison within our sample can be accomplished by using the IRAC spectral index $\\alpha_{\\rm IRAC}$, as determined from a least-squares fit to the photometry of all four IRAC bands \\citep{lad06}; see section~\\ref{sec_anc} for a discussion of the photometry. The sources in our NGC 2264 subsample show diverse IRAC spectral indices, ranging from $\\alpha=-1.41$ (source 27) to $\\alpha=4.21$ (source 45). Eighteen sources in this subsample have positive spectral indices. \\citet{sun09} identify 37 out of the 40 sources as candidate cluster members in various stages, 20 of them are listed as class~I sources; their classification is included in Table~\\ref{tab_srclist}. Note that while \\citet{sun09} list no membership status for the three remaining sources, our results suggest that they are cluster members, too. In the IC\\,348 sample, the IRAC spectral indices range from $\\alpha=-0.93$ (source 14) to $\\alpha=1.92$ (source 6). The classification from \\citet{lad06} and \\citet{mue07} is listed in Table~\\ref{tab_srclist}. Most of the sources in this subsample are class I objects. Since we do not find significant differences between these two cluster subsamples -- both (unevenly) sample the same YSO classes --, we will discuss them as one sample in the following, focusing more on a discussion of evolutionary stages in general. Note, however, that the source table allows the reader to identify individual YSOs in both clusters. In this paper, we first discuss the \\textsl{Spitzer}-IRS observations and their data reduction. We then describe the ancillary multi-wavelength data, ranging from optical to submillimeter wavelengths, and how we assembled composite SEDs from both IRS spectra and ancillary data. Subsequently, we perform comprehensive modeling of these SEDs before concluding with a discussion and summary of the results. ", "conclusions": "After the analysis of the IRS spectra and subsequent detailed SED modeling, we can characterize the YSOs in our sample more accurately than what was previously possible from YSO spectral classes alone by assigning evolutionary stages next to their SED classes. As we have seen, the comprehensive SED modeling suggests that the sources fall into three classes: 1) genuine protostars (stage I), 2) disk sources behind high foreground extinction (stage II(ex)), and 3) disk sources with less external foreground extinction (stage II). Of the 44 sources for which we determine evolutionary stages, 17 are stage I protostars. Most of these were already classified as class~I in the literature (one, source 4, is a class 0 source, but a stage earlier than stage~I has not been defined), with the exceptions of sources 28, 45, and 51 which have not been classified, or have been classified differently (source 35) in \\citet{sun09}. On the other hand, as summarized in Table~\\ref{tab_srclist}, the sample contains many previous class~I sources that instead appear to be in stage II(ex). Thus, a number of sources appear to be in later evolutionary stages than what was previously suspected. Most previously determined class~II sources turn out to be in stage II, although a few are in stage II(ex). With regard to the use of spectral indices as an indicator of YSO class, our sample contains a few objects that can serve as instructive examples of the limits of using a single spectral index for the classification of individual sources. Sources 3 and 56 both have negative IRAC spectral indices, but it turns out that they are stage I protostars and were classified as class~I protostars in the literature. The IRS data are a significant help to constrain the SEDs of these sources both by allowing us to make use of spectral properties in YSO characterization, and by extending the SED wavelength coverage. The latter is further improved since the IRS data allow us to avoid spectral ranges contaminated by unwanted line features for photometry and subsequent SED modeling. In particular, the youngest sources show strong silicate and ice features in their mid-infrared spectra; both features are correlated with the sources' spectral indices, both IRAC and \\textsl{K}--24\\,$\\mu$m. A correlation of the silicate optical depth at 9.7\\,$\\mu$m and the total ice column density is present, but has more scatter than the correlations of the two features with the IRAC spectral index as a measure of evolutionary stage. The main exception appear to be objects with high silicate optical depths but low total ice column densities. These might be more massive stars with radiation fields able to evaporate the ice mantles. A similar correlation is found in the c2d data and by \\citet{fur08}. We detect three sources with PAH emission (all in NGC~2264), a small subset of our sample. Presumably, these are more massive YSOs. For the crowded region of the \\textsl{microcluster} of NGC~2264, we can make use of IRS spectral maps, for example to identify regions with jet activity. Based on the IRS and modeling results, we can check whether the three evolutionary stages that were identified by SED modeling in turn have distinctive features in their IRS spectra. We have therefore marked the sources in Fig.~\\ref{fig_speccorr} according to their evolutionary stage. Indeed, the stage I sources tend to have large optical depths of the silicate feature and strong ice features. Stage~II(ex) sources have less strong silicate and ice features and stage II sources tend to show the silicate feature in emission while not showing any ice features. It is interesting to note that the IRS spectral features also appear to be distinctly different when comparing stage II(ex) and stage II sources. Fig.~\\ref{fig_speccorr} shows that while virtually all stage II sources have the 9.7~$\\mu$m in \\textsl{emission}, indicative of disks, that is \\textsl{not} the case for the stage II(ex) sources, presumably due to the additional extinction which may `cancel out' the emission. Also, none of the stage II sources show ice features, while stage II(ex) interestingly do show ice absorption, although at smaller total column densities than the stage I sources. In our interpretation, the stage II(ex) ice features would be related to the additional extinction by intracluster foreground material and not the sources themselves. In summary, while IRS data are of crucial importance, they alone are not sufficient to distinguish the evolutionary stages. For this purpose, detailed SED modeling is required. We point out that in the context of our comprehensive SED modeling, the archival submillimeter data were very important for the separation of the sample into the different evolutionary stages. Dense dust and gas in an envelope manifest themselves in submillimeter maps as concentrated emission in contrast to more large-scale foreground extinction. Future high-angular resolution observations in the far-infrared/submillimeter wavelength range, as forthcoming from the \\textsl{Herschel} Space Observatory, will be even better suited to help determining the evolutionary stages in YSO samples. Finally, we discuss the observable quantities that correlate best with the different evolutionary stages. We have seen that this is certainly not a spectral index since the correlation of the evolutionary stages with spectral indices is not unambiguous. The genuine protostars (stage I) all have submillimeter detections and the highest silicate optical depths as well as the highest ice column densities. The stage II sources are broadly characterized by a silicate feature in emission and no ice features. The stage II(ex) sources lie between these two extremes, but their SEDs can be easily mistaken for class I sources (in fact, some were previously determined to be class I sources, see above); in particular, they do not show the silicate feature in emission. Once an individual SED shape is well characterized, far-infrared and submillimeter continuum data are the most important photometric data to distinguish the different evolutionary stages." }, "1005/1005.1037_arXiv.txt": { "abstract": "We continue to explore the validity of the reflected shock structure (RSS) picture in \\snr that was proposed in our previous analyses of the X-ray emission from this object. We used an improved version of our RSS model in a global analysis of 14 CCD spectra from the monitoring program with {\\it Chandra}. In the framework of the RSS picture, we are able to match both the expansion velocity curve deduced from the analysis of the X-ray images and light curve. Using a simplified analysis, we also show that the X-rays and the non-thermal radio emission may originate from the same shock structure (the blast wave). We believe that using the RSS model in the analysis of grating data from the {\\it Chandra} monitoring program of \\snr that cover a long enough time interval, will allow us to build a more realistic physical picture and model of \\snrE. ", "introduction": "After its reappearance in radio (\\citealt{turtle_90}; \\citealt{stav_92}, 1993) and in X-rays (\\citealt{buer_94}; \\citealt{gor_94}; \\citealt{has_96}) at about 1000-1200 days after the explosion (DAE), SN 1987A entered into the phase of supernova remnant (SNR) formation \\citep{dick_07}. Observations of this phenomenon from its onset give us a chance to study in detail the physics of strong shocks as well as to investigate the distribution of the circumstellar matter (CSM) around the exploded star. The latter may help us solve the mystery of the origin of the triple-ring system observed in the optical (\\citealt{wampler_90}; \\citealt{jak_91}; \\citealt{crotts_91}; \\citealt{burr_95}). With this respect, X-ray observations are very important since they manifest most of the energetics at present and they provide us with direct information about the underlying physics of this evolving object. The richness and variety of processes involved in the birth of \\snr are manifested by observations across the entire electromagnetic spectrum. Radio observations (\\citealt{stav_92}, 1993, 2007; \\citealt{ball_01}; \\citealt{man_02}, 2005; \\citealt{gaensler_97}, 2007; \\citealt{ng_08}; \\citealt{potter_09}; \\citealt{zana_10}) tell us about the non-thermal (synchrotron) emission mechanisms and relativistic particle acceleration in strong shocks. Near-infrared observations tell us about the distribution of dust in the CSM and the physics of its interaction with the hot plasma (\\citealt{bou_04}, 2006; \\citealt{dwek_08}). Optical and UV observations (\\citealt{son_98}; \\citealt{law_00}; \\citealt{pun_02}; \\citealt{suger_02}; \\citealt{gron_06}, 2008) tell us about the physics of slow shocks and the distribution of dense clumps of gas. The early evolution of \\snr is strongly influenced by the distribution of circumstellar gas, which is concentrated in its triple-ring system. Although the origin of the ring system is not well understood, the colliding wind scenario that assumes a spherically-symmetric blue supergiant (BSG) wind interacting with an earlier emitted asymmetric red supergiant (RSG) wind seems the physically most reasonable explanation (\\citealt{luo_91}; \\citealt{wang_92}; \\citealt{morris_07}). It has been explored in detail by \\citet{bl_lu_93} who demonstrated that such a model is capable of accounting for the observed CSM characteristics, provided the RSG wind was highly asymmetric with most of its mass confined near the equator. This picture needed an additional ingredient in order to explain the reappearance of \\snr in radio and X-rays. \\citet{che_dwa_95} proposed that an extended HII region was formed inside the inner equatorial ring that was naturally produced by the photoionization of the shocked RSG wind (the inner ring) by the central progenitor BSG star. Based on this, Borkowski et al. (1997a, b) studied the complex hydrodynamics of the interaction of the SN ejecta with the CSM in vicinity of the inner ring that also predicts a continuous brightening of \\snr in UV and X-rays. Since its reappearance, \\snr has been brightening continuously in X-rays with an accelerating rate in the past few years (\\citealt{sang_05}, 2006, 2007; \\citealt{judy_09}). The first spectrum of \\snr with high spectral resolution was obtained with the high energy transmission grating on {\\it Chandra} by \\citet{eli_02}. The subsequent brightening has enabled dispersed spectra with good photon statistics to be obtained with both {\\it Chandra} (\\citealt{zhek_05}, 2006; \\citealt{dew_08}; \\citealt{zhek_09}) and {\\it XMM-Newton} (\\citealt{haberl_06}; \\citealt{heng_08}). With the superb spatial and spectral resolution of the {\\it Chandra} observatory, the first direct evidence was obtained that the distribution of the X-ray emitting plasma in \\snr is not spherically-symmetric but it is `flat', {\\it i.e.,} it is concentrated in the vicinity of the equatorial ring (\\citealt{zhek_05}, 2006; \\citealt{dew_08}; \\citealt{zhek_09}). These authors also showed that the bulk gas velocities deduced from the X-ray spectral lines are too low to account for the plasma temperatures inferred from the spectral fit, and showed that a simplified reflected-shock structure (RSS) model could match the observed spectra successfully. Superb spatial resolution of {\\it Chandra} allowed \\citet{burr_00} to obtain the first image of \\snr that showed a ring-like structure in X-rays. Subsequently, a monitoring program was established with {\\it Chandra} to track the evolution of the X-ray images and CCD spectra of \\snrE. In addition to the continuing brightening in X-rays, the spectral data showed a gradual decrease of the plasma temperature behind the fast shocks. From the imaging data, we measured the expansion velocity of the remnant (\\citealt{sang_02}, 2004, 2007). After applying a physically more realistic method to analyze the X-ray images, \\citet{judy_09} obtained a more accurate expansion velocity curve. It showed clearly that the velocity started to decelerate after $\\sim 6000$~DAE, about the same time that the X-ray light curve turned up \\citep{sang_05}. Here we analyze the evolution of the X-ray spectrum of \\snr as observed with the pulse-height spectra obtained with the ACIS CCD camera during the monitoring program. One of the main goals of our study is to test the reflected-shock structure picture that emerged from analysis of the {\\it Chandra} grating data. Since the pulse-height spectra have much better time coverage than the grating data, they can test whether is possible to match the CCD spectra and the expansion velocity curve in the framework of the RSS picture. In section \\S~\\ref{sec:rss_picture} we give a qualitative description of the RSS picture for \\snrE. In section \\S~\\ref{sec:rss_model} we present details about the global RSS model. We describe the evolution of the CCD spectra and results from the spectral fits in sections \\S~\\ref{sec:data} and \\S~\\ref{sec:spectral_fits}, respectively. In section \\S~\\ref{sec:discussion} we discuss the results, how they might be related to the non-thermal radio emission from \\snr and the consistency of the global RSS picture. We present our conclusions in section \\S~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} In this study, we continued to explore the validity of the reflected-shock structure picture in \\snrE. This picture assumes that the blast wave is interacting with the HII region in vicinity of the inner ring and when it encounters dense clump(s) a shock is transmitted into the clump(s) and a reflected shock forms that additionally compresses the shocked gas behind the blast wave. We adopted an improved version of the reflected-shock model (global RSS model) to test the RSS picture in \\snr by making use of CCD spectra from the monitoring program with {\\it Chandra}. The main results from this analysis are as follows. 1. The global RSS model is capable of matching both the X-ray expansion velocity curve \\citep{judy_09} as well as the X-ray spectra of \\snr over its evolution for the last ten years or so. Moreover, extrapolating back in time the X-ray flux, the model is able to match the observed {\\it ROSAT} fluxes from the time of the supernova reappearance in X-rays (1000-3000 DAE). 2. The evolution of the mean electron temperature behind the shocks shows a decrease with time of the electron temperature in the blast wave, while a slight increase with time is noticed for this parameter in the transmitted shock. The mean ratio of the electron to the mean plasma temperature is required to increase with the decrease of the velocity of the blast wave. 3. At early times, the soft (0.5-2 keV) X-ray flux of \\snr was dominated by the blast wave but for times after $\\sim 6000$ DAE the transmitted shock became dominant. This transition coincides with the steep upturn in the X-ray LC reported by \\citet{sang_05}. Also, the emission measure of the hot plasma in the transmitted shock (as well as in the reflected shock) is increasing steeper than that for the blast wave for times after $\\sim 6400$ DAE. All this is indicative of an ongoing penetration of the blast wave deeper in the equatorial ring debris. 4. Results from the X-ray analysis were used to explore the possibility that the X-rays and the non-thermal radio emission from \\snr originate from the same shock structure (the blast wave). Using a simplified analysis, it was shown that this can indeed be the case, provided the particle acceleration mechanisms, that operate in the shocks in \\snrE, are not very efficient in producing relativistic particles with very high energies. 5. As a next step, the global RSS model could be used to obtain further constraints on the reflected-shock structure picture in \\snr by analyzing X-ray spectra with good spectral resolution once the available data cover a long enough time interval of the \\snr evolution. We believe that this will allow us to improve our understanding of the underlying physics and to build a more realistic picture and model of this fascinating phenomenon: the birth and evolution of supernova remnant 1987A." }, "1005/1005.3817_arXiv.txt": { "abstract": "We report spectroscopic confirmation and high-resolution infrared imaging of a $z=2.79$ triply-imaged galaxy behind the Bullet Cluster. This source, a \\spitzer-selected luminous infrared galaxy (LIRG), is confirmed via polycyclic aromatic hydrocarbon (PAH) features using the \\spitzer~Infrared Spectrograph (\\irs) and resolved with \\hst~\\wfc~imaging. In this galaxy, which with a stellar mass $M_*\\approx4\\times10^9$ M$_\\sun$ is one of the two least massive ones studied with \\irs\\ at $z>2$, we also detect $H_2\\;S(4)$ and $H_2\\;S(5)$ pure rotational lines (at 3.1$\\sigma$ and 2.1$\\sigma$) -- the first detection of these molecular hydrogen lines in a high-redshift galaxy. From the molecular hydrogen lines we infer an excitation temperature $T=377^{+68}_{-84}$ K. The detection of these lines indicates that the warm molecular gas mass is $6^{+36}_{-4}$\\% of the stellar mass and implies the likely existence of a substantial reservoir of cold molecular gas in the galaxy. Future spectral observations at longer wavelengths with facilities like the \\herschel\\ {\\it Space Observatory}, the Large Millimeter Telescope, and the Atacama Pathfinder EXperiment (APEX) thus hold the promise of precisely determining the total molecular gas mass. Given the redshift, and using refined astrometric positions from the high resolution imaging, we also update the magnification estimate and derived fundamental physical properties of this system. The previously published values for $L_{IR}$, star formation rate, and dust temperature are confirmed modulo the revised magnification; however we find that PAH emission is roughly a factor of five stronger than would be predicted by the relations between $L_{IR}$ and $L_{PAH}$ reported for SMGs and starbursts in \\citet{pope2008}. ", "introduction": "\\label{sec:intro} An important legacy of the \\spitzer~{\\it Space Telescope} is that it enabled the first detection of dust emission from a large number of luminous infrared galaxies at $z>2$ \\citep[e.g.,][]{papovich2006,perez2005}, demonstrating that these galaxies dominate the massive galaxy population at high redshift. The implication is that at $z\\sim 2-3$ massive galaxies are rapidly assembling their stars and growing supermassive black holes. For the most infrared-luminous systems, spectra from the \\spitzer\\ Infrared Spectrograph (IRS) have provided important insights into the physical processes driving $L_{IR}$ \\citep[e.g.,][]{houck2005,yan2007}. \\citet{pope2008} used \\irs\\ to demonstrate that the mid-IR properties of ultraluminous infrared galaxies (ULIRGs, $L_{IR}>10^{12}$ L$_\\sun$) and submillimeter galaxies (SMGs) are distinct, with star-formation dominating the infrared emission for typical SMGs. It has been postulated by this group and others that such differences indicate different evolutionary stages, as the dominant source of $L_{IR}$ transitions from star formation to AGN emission. Existing studies like \\citet{pope2008} clearly demonstrate that \\irs~ spectroscopy provides a clean means of disentangling the AGN and star formation contributions. A limitation to such work though is that current samples only probe the bright end of the luminosity function at high-z \\citep{dey2008,dye2008}, and \\irs~observations are practical for only a small subset of the most luminous sources. Observations of luminous infrared galaxies (LIRGS, $10^{11}2$.\\footnote{The lensed LBG MS1512-cB58 has a factor of $\\sim2-4$ lower stellar mass \\citep[see][]{siana2008}.} Perhaps most promising for future investigations, our detection of rotational H$_2$ lines is indicative of a large molecular gas reservoir. We derive a temperature of $T=377^{+68}_{-85}$ K and estimate a gas mass of $M_{H_2}=2.2^{+17}_{-0.8}\\times10^8$~M$_\\sun$ in this component, roughly 2-42\\% of the stellar mass. Future spectroscopic observations of longer wavelength H$_2$ lines with \\herschel\\ and CO rotational lines with a facility such as the Large Millimeter Telescope or the Atacama Pathfinder EXperiment (APEX) therefore have the potential to precisely measure the total molecular gas content in this galaxy. Finally, we note that because dwarf galaxies of this mass are a ubiquitous population, the odds are good that programs like the \\herschel\\ Lensing Survey can detect a sizable sample of similar lensed sources behind massive clusters." }, "1005/1005.4430_arXiv.txt": { "abstract": "Measuring the full three-dimensional motions of extra-galactic objects in the Universe presents a seemingly insurmountable challenge. In this paper we investigate the application of a technique to measure tangential motion that has previously only been applied nearby within the Local Group of galaxies, to clusters of galaxies far beyond its borders. We show that mapping the mean line-of-sight motion throughout a galaxy cluster could in principle be used to detect the {\\it perspective rotation} induced by the projection of the cluster's tangential motion into the line-of-sight. The signal will be most prominent for clusters of the largest angular extent, most symmetric intrinsic velocity distribution and surveyed with the largest number of pointings possible. We investigate the feasibility of detecting this signal using three different approaches: measuring line-of-sight motions of individual cluster members; taking spectra of intracluster gas; and mapping distortions of the Cosmic Microwave Background radiation. We conclude that future spectroscopic surveys of 1000's of members of nearby galaxy clusters hold the most promise of measuring cluster tangential motions using this technique. ", "introduction": "It is our unhappy luck that even though we live in three dimensions, we are forced to view most of the Universe as a two dimensional projection with only one dimension of velocity information. As such, we know very little about the motions of the things around us in any direction other than radially outward. A more comprehensive view of motions in the Universe would provide a unique probe into the history and future of the things around us. Indeed, within the Milky Way, the HIPPARCOS \\citep{perryman97} and near-future GAIA \\citep{perryman02} missions have brought about a rebirth in the ancient field of astrometry, pushing accuracies for direct proper motion measurements down first to mas/year precision and then on to tens of $\\mu$as/year. The prospect of billions of stars with full 6-D phase-spa2$ keV ($L_{2-10 keV} \\sim 10^{40}$ ergs s$^{-1}$). %It is likely the X-ray counterpart of another strong radio and infrared source in %the central region. If X-1 is instead associated with the kinematic center, the nucleus of NGC253 is compatible with an obscured low luminosity active galactic nucleus (AGN) or a spatially resolved super star cluster (SSC) brightening up in X-rays most probably due to young supernovae or supernova remnants, a situation also observed in the nuclear starburst of M82. %we then suggest that the nucleus of NGC253 may harbor an obscured low luminosity %active galactic nucleus (AGN). %If X-1 corresponds to the kinematic center of the galaxy, %the simplest explanation accounting for these properties is that the nucleus of %NGC253 harbors a hidden LLAGN. %\\item Alternatively, X-1 may be associated with one SSC in this region: SSC %IR-11/radio-TH6. In this case the nucleus of NGC253 would then be a $10^5-10^6$ %M$_\\odot$ spatially resolved (diameter of $\\sim5$ pc) NSC in which some of its %massive stars are evolving into the SN phase. If no SSC is associated with the kinematic center, we conclude that NGC253 is a galaxy in which a strong starburst and a weak AGN (either TH2 or X-1) coexist. Results from few other high resolution studies of nearby starburst galaxies (e.g. M82, NGC5253, NGC4945) indicate that the AGN in these systems, if present, is always in the low luminosity regime. This may indicate that the onset of nuclear activity in galaxies is closely related with the occurrence of star formation, and that we are witnessing the emergence or dissaperance of an AGN. %Currently the gas in the nuclear region of NGC253 is feeding intense multiple %starbursts and the galactic nucleus is thus in a latent phase of activity. %Both, the radio core and the central peak of hard X-ray emission, are consistent %with the kinematic center location, and therefore are the most likely galactic %three Ultraluminous X-ray sources (ULXs) in the central $10\\arcsec \\times %10\\arcsec$ but no source was detected at the position of TH2. %The strongest point-like X-ray source, HX-1, lies $\\sim 1.2 \\arcsec$ from TH2, has %an unabsorbed 2-10 keV luminosity of $\\sim 10^{43}$ ergs s$^{-1}$ and is observed %through an absorbing column density of $7.5 \\times 10^{23}$ cm$^{-2}$. %can be associated with another compact radio source TH6. %The two flat spectrum radio sources %We suggest that while HX-1 is either an intermediate-mass black hole or a weakly %accreting supermassive black hole, TH2 is a near twin of our Galactic Center %SgrA$^*$. ", "introduction": "\\label{int} %In this paper we present the first sub-arcsecond resolution two-dimensional %stellar kinematics and X-ray observations of the prototypical starburst galaxy %NGC253 which define the position and nature of the galactic nucleus. %imaging observations of the 2.3 km CO absorption feature It is observationally well established that star formation in starburst galaxies manifests itself in the production of numerous young massive star clusters of extraordinary luminosity and compactness, normally referred to as super star clusters (SSCs). In fact, the four most luminous circumnuclear starburst systems (M82, M83, NGC253 and NGC4945) account for $25\\%$ of the high-mass ($>8$ M$_\\odot$) star formation within 10 Mpc \\citep{heckman98}. Nearby starburst galaxies provide thus a unique laboratory for investigating the mechanisms which cause enhanced star formation to occur and be sustained, as well as physical conditions analogous to intensely star-forming galaxies identified at high-z (e.g. Shapley et al. 2003). An intriguing issue in studies of starburst galaxies is the apparent absence of a well defined nuclear source which could be associated with the center of the galaxy. %The key problem is that there exist offset between photometric centers (e.g. the %radio core is not coincident with the IR/visible photometric center). %A key problem in studies of starburst galaxies is to establish the physical nature %of the center of the galaxy. This situation is present in spatially resolved starbursts in nearby galaxies: e.g. M82 \\citep{matsumoto01}; NGC4945 \\citep{marconi00}; NGC5253 \\citep{summers04}; NGC253 (this work). Massive black holes (BHs) and/or nuclear star clusters (NSCs) are thought to reside in the centers of most galaxies, with the first being mostly identified in massive ellipticals \\citep{richstone98, graham08} and NSCs in irregular and late spirals (e.g. B\\\"oker et al. 2002; Seth et al. 2010). Although tens of SSCs exist in the central region of starburst galaxies, the presence of a NSC in these systems is very poorly constrained. Since the central few hundred parsecs of starburst galaxies are heavily affected by large and nonuniform dust extinction, the determination of SSCs properties (age, colors, luminosities, etc.), and thus the detection of a putative NSC, has been difficult and uncertain. %The ultimate observational evidence of the presence of massive BHs in the centers %of galaxies comes from kinematic studies, by measuring the gravitational influence %of the central object on neighboring stars and gas. As kinematic BH mass %determinations require very high spatial resolution, this has been possible to %achieve only in a few number of objects, being the most believable example the %supermassive black hole (SMBH) at the center of the Milky Way \\citep{begelman03}. The presence of massive BHs in galaxy centers is normally inferred when nuclear activity is observed \\citep{filippenko89}. %However, the presence or absence of any of them in starburst galaxies is very %poorly constrained. This is due to mainly two reasons: 1) the central few hundred %parsecs of starburst galaxies are heavily affected by large and nonuniform dust %extinction, making difficult and uncertain the determination of SSCs properties %and thus the detection of a putative NSC, and 2) the photometric centers at radio, %IR and X-ray wavelengths are not spatially coincident. The presence of massive BHs %in galaxy centers is normally inferred when nuclear activity is observed (ref %Seth). Active Galactic Nuclei (AGN) are usually characterized by prominent point sources in $K$-band. Starburst galaxies do not show any IR point-like emission at the position of the assumed galactic nucleus, which in the majority of these systems has been associated to either a powerful compact radio source or a hard X-ray peak. However, in several cases it is not possible to reject/accept the AGN hypothesis on the basis of radio/X-ray emission peaks alone, particularly when the three photometric centers (radio, IR, X-ray) don't coincide. For the nearest, best studied starbursts, possible galactic center candidates include: young supernovae (SNe), supernova remnants (SNRs), intermediate-mass BHs, inactive supermassive BHs, low-luminosity AGN, etc. (see e.g. Wills et al. 1999 and Stevens et al. 1999 for the case of M82, Marconi et al. 2000 for NGC4945 and Fern\\'andez-Ontiveros et al. 2009 for NGC253). %In addition, due to the strong, patchy obscuration in these systems the nature of %the SSCs and the photometric centers is far As the center of a galaxy is likely to coincide with the bottom of the galaxy potential well, independent estimates of the kinematic center position from kinematic data can help to determine which photometric center, if any, is the most promising candidate for galactic nucleus. These important dynamical studies had not previously been widely performed because high S/N spectra over a full 2D field with high spatial resolution are required, and these are still scarce. In this paper we provide the first estimate of the kinematic center of the prototypical starburst galaxy NGC253 using sub-arcsecond resolution two-dimensional stellar kinematics. This is combined with data at radio, infrared and X-ray wavelengths to examine the nature of the galactic nucleus itself. These observations comprise some of the best probes of the inner region of NGC253 and will allow us to gain more insight into the nature of the centers of starburst galaxies. %and X-ray observations of the prototypical starburst galaxy %NGC253 which define %the position and nature of the galactic nucleus. %In this paper we investigate the nature of the nucleus of the prototypical %starburst galaxy NGC253 using %existence of AGNs in starburst galaxies %through observations of NGC253, one of the closest and hence well studied %starburst galaxies, which is an ideal laboratory to test the power of integral %field spectroscopy and stellar kinematics in the determination of kinematic %centers of galaxies, particularly those which are largely inaccessible optically %because of dust obscuration, %has served as testing ground for study starburst theories %NGC253 is an ideal laboratory for studying any potential causal connection %between starburst and AGN activity. %provide an estimate of the kinematic center of the prototypical starburst galaxy %NGC253. Understanding the nature of the nuclei of starburst galaxiesThis is a %crucial step in understanding how a putative massive BH and a strong starburst %relate to each other. %for studying any potential causal connection between starburst and AGN activity. %With a strong patchy and with tens of SSCs detectable in the near-IR, the search %for a NSC in these systems is far from a simple task. making difficult and %uncertain the detection of SSCs and thus the determination of their properties, %Over the past decade the search for central massive objects (CMOs) in local %galaxies has experienced a major break-through. Mostly thanks to high spatial %resolution techniques and the development of new theoretical tools for the %analysis of stellar dynamical data, we have learned that probably all centers of %galaxies contain CMOs in the form of massive black holes (BHs) and/or nuclear star %clusters (see Richstone et al. 1998, Bender \\& Kormendy 2003, Seth et al. 2010). %, which in addition are largely inaccessible optically because of dust %obscuration, %Over the past decade, there has been increasing interest in studying any possible %physical connection between circumnuclear starbursts and active galactic nuclei %(AGN). Thanks to high spatial resolution techniques it is now widely accepted that %many Seyfert galaxies contain both a circumnuclear starburst and an AGN (e.g. %Davies et al. 2007). To date, however, %little is known about the centers of nearby starburst galaxies ($<10$ Mpc) %it is not clear whether or not the centers of nearby starburst galaxies ($<10$ %Mpc) host massive black holes (BHs). %responsible for AGN activity. %or nuclear star clusters %or not starburst galaxies host accreting supermassive BHs. %In the Milky Way, NGC4258 and likely M31, it is virtually certain that these dark %objects are indeed supermassive BHs, because alternatives in the form of stellar %remnants can be ruled out (Maoz 1998, Bender et al. 2005). %The ultimate observational evidence of the presence of massive black holes (BH) in %the centers of galaxies comes from kinematic studies, by measuring the %gravitational influence of the central object on neighboring stars and gas. As %kinematic BH mass determinations require very high spatial resolution, this has %been possible to achieve only in a few number of objects, being the most %believable example the supermassive black hole at the center of the Milky Way %\\citep{begelman03}. %However, it is possible to perform kinematic studies at larger scales and estimate %the kinematic centers of galaxies from rotation curves, or even better, from %two-dimensional velocity fields. By associating the kinematic center with the %galactic nucleus, as it occurs in the majority of galaxies, one can then use other %methods to unveil the nature of the central object. Strong evidence of the %presence of a nuclear BH comes from its accretion-powered emission %\\citep{salpeter64}. Nuclear activity can be identified either from the detection %of broad recombination lines in the optical/near-IR spectrum of the nuclear %region, or from the presence of forbidden high ionization lines (coronal lines), %or from a point-like X-ray nuclear source, or a flat-spectrum radio core. %As consequence, a key problem in studies of nearby starburst galaxies is to %establish the physical nature of the center of the galaxy. %, which in addition are largely inaccessible optically because of dust %obscuration, is to establish the physical nature of the center of the galaxy. %The fundamental question arising here is: do the centers of starburst galaxies %contain massive dark objects %, particularly when the photometric centers at radio, IR and/or X-ray wavelengths %don't spatially coincide %Previous studies on the nature of the nuclei of starburst galaxies have not %reached a general consensus; cases were made for supernova remnants (NGC1569, %Westmoquette), SSCs, intermediate-mass black holes (NGC253, Weaver et al. 2002), %inactive supermassive black holes (M82, Wills et al. 1999; NGC253, %Fernandez-Ontiveros et al. 2009), low-luminosity AGN (M82, japones; NGC253, Turner %\\& Ho) , and powerful AGN (NGC4945, Marconi et al. 2000). %In particular, it is important to know if it is possible to hide an AGN, %contributing significantly to the bolometric emission, when optical to mid-IR %spectroscopy and imaging reveal only a starburst component. %There are several ways in which one can determine the position of the center of a %galaxy. As nuclear activity in galaxies is likely to coincide with the bottom of %the galaxy potential well, a central compact radio source is a good indicator of %the center of a galaxy \\citep{hummel84}. %One can also determine the center by fitting ellipses to the surface brightness %distribution. For those galaxies which also show a well-defined nuclear source at %infrared or visible wavelengths, one can derive the central position by fitting a %Gaussian to the central source. Similarly, nuclear star clusters can be used to %locate the center (Matthews \\& Gallagher 2002). Each of these estimates yields a %photometric center. Another very important independent estimate can be obtained %from kinematic data by constructing a model for a rotation curve or velocity %field. %a tilted-ring model for a rotation curve or velocity field. %In this case a kinematic center is derived. Ideally, these different %determinations should all coincide. Indeed, for most galaxies the kinematic center %agrees %with the position of the radio continuum source and/or the photometric center as %derived from IR/visible imagery. However, in some cases, there exist offset %between photometric centers (the radio core is not coincident with the IR/visible %photometric center) or between the kinematic and photometric centers, this being %an indication of disturbances such as merging processes, spiral arms or strong %bars. When photometric centers do not coincide, as in the protoypical starburst %galaxy NGC253 \\citep{fernandez09}, the kinematic center position can help to %determine which photometric center, if any, is the most promising candidate for %galactic nucleus. In this paper we provide such independent estimate of NGC253's %center. %In the latter case, the discrepancies indicate disturbances such as merging %processes, spiral arms or strong bars. %The ultimate observational evidence of the presence of massive black holes (BH) in %the centers of galaxies comes from kinematic studies, by measuring the %gravitational influence of the central object on neighboring stars and gas. As %kinematic BH mass determinations require very high spatial resolution, this has %been possible to achieve only in a few number of objects, being the most %believable example the supermassive black hole at the center of the Milky Way %\\citep{begelman03}. %By associating the kinematic center with the %galactic nucleus, as it occurs in %the majority of galaxies, one can then use other %methods to unveil the nature of the central object. Strong evidence of the %presence of a nuclear BH comes from its accretion-powered emission %\\citep{salpeter64}. Nuclear activity can be identified either from the detection %of broad recombination lines in the optical/near-IR spectrum of the nuclear %region, or from the presence of forbidden high ionization lines (coronal lines), %or from a point-like X-ray nuclear source, or a flat-spectrum radio core. As the nearest massive galaxy experiencing a strong starburst (3.94 Mpc, $1^{\\prime\\prime}$ = 19 pc, Karachentsev et al. 2003), %NGC253 is an ideal laboratory for studying any potential causal connection %between starburst and AGN activity. %Despite its proximity %(3.94 Mpc, $1^{\\prime\\prime}$ = 19 pc, Karachentsev et al. 2003), many aspects %of its nuclear region remain poorly understood. One of them is precisely the %exact location and physical nature of the center of the galaxy. NGC253 is one of the most studied objects in the sky, yet many aspects of its nuclear region remain poorly understood. %despite its proximity. One of the main unresolved issues is the precise location and physical nature of the center of the galaxy. %A variety of investigations of NGC 253 have provided information about the %individual compact radio sources %located within the inner 100 pc of the galaxy \\citep{turner85, ulvestad97, %lenc06, brunthaler09}. The galactic nucleus has historically been associated to a powerful compact non-thermal radio source (TH2 in Ulvestad \\& Antonucci 1997) with characteristics similar to other synchrotron sources in active galaxies \\citep{turner85}: very small size ($<100$ mas), %(VLA A-configuration observations do not resolve TH2 at any wavelength, see %Table 3), %e.g. size $<130 \\times 70 mas$ at 1.3 cm), high brightness temperature ($T>40000$ K at 15 GHz), and a relatively flat spectral index between $8.4-23$ GHz($\\alpha =-0.3$, $S_\\nu \\propto \\nu^\\alpha$). Follow up studies accross the electromagnetic spectrum, each limited by the spatial resolution available at that time, have matched this source with i) H$_2$O maser emission \\citep{nakai95}, ii) different knots of IR emission \\citep{forbes00, galliano05}, and iii) the hard X-ray emission peak \\citep{weaver02}, %and the dynamical center of the galaxy (Arnaboldi, Yo), suggesting that TH2 may be a low luminosity active galactic nucleus (LLAGN). However, it is now apparent from recent high-spatial resolution observations at different wavelengths, and using diverse observational techniques, that this picture may be incorrect. The evidences are the following: \\begin{itemize} \\item VLA and VLBA observations at 23 GHz \\citep{brunthaler09} have shown that the high velocity H$_2$O maser is not associated with TH2 but with another nearby continuum source (TH4 in Ulvestad \\& Antonucci 1997), which has been classified as a SNR embedded in an HII region \\citep{lenc06}. This suggests that a maser-AGN connection is unlikely and that the maser emission is almost surely related to star formation. \\item Very high spatial resolution VLBI observations at 2.3 and 23 GHz \\citep{lenc06, brunthaler09} did not detect TH2 on mas scales. As this source is unresolved and very bright in VLA 5-23 GHz continuum maps (beam sizes $>100$ mas), %this implies that the source is either over-resolved or is too faint. In both %cases this non-detection questions this questions the presence of a possible AGN. %However, as pointed out by \\citet{brunthaler09}, this non-detection does not rule %out other scenarios such as the existence of an inactive supermassive black hole %like Sgr A$^*$. \\item Very accurate alignment at parsec scales has been found between the near-IR and radio sources by means of Adaptive Optics VLT/NACO IR data with a comparable resolution to those in existing VLA radio maps \\citep{fernandez09}. Surprisingly, this new alignment shows that TH2 has no infrared counterpart. The newly revealed IR nature of TH2 shows no evidence of a central point source that would indicate nuclear activity. However, the non-detection of IR emission does not rule out other scenarios such as the existence of an inactive supermassive black hole (SMBH) like Sgr A$^*$ \\citep{fernandez09, brunthaler09}. %and instead suggests that TH2 may be a twin of Sgr\\,A$^*$ rather than an AGN. \\end{itemize} In addition, one needs to be cautious when associating the hard X-ray peak to TH2. It is well known that NGC253 harbors $\\sim 60$ compact radio sources \\citep{ulvestad97} and $\\sim37$ near-infrared point-like sources \\citep{fernandez09}. This significant change in morphology with wavelength has led to different registrations in the literature (see e.g. Galliano et al. 2005 and references herein). Since \\textit{Chandra} data were processed with $0.5\\arcsec$ pixel $^{-1}$ resulting in $\\sim 1 \\arcsec$ resolution and have an astrometric accuracy of $\\sim 1 \\arcsec$ rms, the hard X-ray peak could be associated with several other discrete sources in the field. %or simply just not coincident with TH2. This is demonstrated in Figure 2 of \\citet{weaver02} where a variety of IR and radio sources are contained within the size of the hard X-ray peak. Moreover, these authors followed the alignment proposed by \\citet{forbes00} which, in light of the current high spatial resolution NACO images presented in \\citet{fernandez09}, seems to be incorrect. Thus, it is still a topic of active investigation whether TH2 harbors an AGN, or even if it corresponds to the galactic center, or not. Based on the alignment proposed by \\citet{fernandez09} one could point to the infrared photometric center (or IR peak) as possible AGN candidate. The IR peak lies $\\sim 3.3 \\arcsec$ SW from TH2. It is in fact the brightest source at infrared wavelengths ($1-20\\, \\mu$m) and emits $\\sim10\\%$ of the galaxy's bolometric flux \\citep{mohan02}. The spatial distribution of the 3.3 $\\mu$m polycyclic aromatic hydrocarbons (PAH) feature-to-continuum ratio reaches a local minimum in this region \\citep{tacconi05}. This indicates that the source produces strong ionization, probably destroying the PAH molecules by photo-dissociation. Supporting this interpretation is the fact that multiband tracers of ionized gas such as Pa$\\beta$, Br$\\gamma$, [HeI] and [NeII] are seen to peak strongly at this position \\citep{boeker98, engelbracht98, keto99, forbes00, tacconi06}. %which is in contrast with the behaviour of other nuclear sources. In addition, the IR peak has an individual counterpart at radio wavelengths (TH7 in Ulvestad \\& Antonucci 1997) which also presents a relatively flat spectrum between $8.4-23$ GHz ($\\alpha=-0.21$, $S_\\nu \\propto \\nu^\\alpha$), but it is $\\sim 4$ times fainter than TH2. %It has been interpreted in terms of a mixture distribution of supernova remnants %and HII regions \\citep{ulvestad97}. Moreover, this strong IR source is located at the vertex of the rather collimated ionised-gas region seen in NGC253 (see Figure 8 of Tacconi-Garman et al. 2005). All these characteristics make this object unique among the other sources in the central region of the galaxy and could be explained either by invoking a very young SSC or an AGN. %The IR peak has been thought to host a Super Star Cluster (SSC). %If there is an AGN in this galaxy, these characteristics point directly to this %source as a candidate. NGC253 has been the subject of many detailed kinematical studies of its gaseous components. Optical rotation curves of the ionized gas have shown that, in addition to ordered rotation, large non-circular motions exist within the central region \\citep{ulrich78, munoz93, arnaboldi95}. However, because of problems of obscuration at optical wavelengths, results from H$\\alpha$ and [N II] observations are difficult to interpret. In addition, HI and H$\\alpha$ rotation curves are incomplete: either they cover the outer parts $r>100\\arcsec$ \\citep{pence80}, or they are confined to the innermost regions where peculiar features and asymmetries in the ionized gas are dominant. Most of the optical and IR observations, aimed at studying the kinematics of the molecular gas in the nuclear region of NGC 253 (e.g., Prada et al. 1996; Engelbracht et al. 1998), have relied on measuring the velocity gradients along chosen position angles passing through the nucleus (e.g., major axis and minor axis). Interferometric radio observations at millimeter and centimeter wavelengths measure two-dimensional velocity fields which provide a more complete picture of the kinematics in the nuclear region. Several radio observations in molecular emission lines such as CO \\citep{mauersberger96, canzian88, das01, paglione04}, CS \\citep{peng96}, HCN \\citep{paglione95}, or in hydrogen recombination lines \\citep{anantamariah96} exist in the literature. These two-dimensional measurements reveal a complex but systematic velocity pattern in the central region. In addition to solid-body rotation, the observed velocity fields indicate motions which may be due to a barlike potential in the center \\citep{peng96} or a kinematic subsystem which may be caused by a past merger event (Anantharamaiah \\& Goss 1996). Although most of these observations have high spectral resolution, they typically have poor spatial resolution $>2.5 \\arcsec$ and are not always useful for probing the dynamics very close to the nucleus. Because of this situation, different authors assumed different centers for their rotation curves/velocity maps: either the center is derived assuming complete symmetry of the outer parts in their fields of view, or it is coincident with the IR peak, or with the radio core TH2, or with the surface brightness peak. The highest resolution measurement of the two-dimensional velocity field available to date is that of Anantharamaiah \\& Goss (1996), who observed the H92$\\alpha$ recombination line with a beam of $1.8\\arcsec \\times 1.0 \\arcsec$ and a velocity resolution of 54 km s$^{-1}$. After trying several positions near the H92$\\alpha$ emission peak, the optimum center of the velocity field was found to be at ($\\alpha$, $\\delta$)$_{2000}$ = $00^h47^m33^s.06$, $-25\\degr 17^\\prime 18\\arcsec .3$ ($1\\sigma$ error $\\sim 0.3\\arcsec$), which is $\\sim 2.0 \\arcsec$ southwest of the H92$\\alpha$ line peak which in turn is coincident with TH2. This means that the most accurate estimate of the kinematic center based on gas kinematics available to date lies closer to the IR peak ($\\sim 1.3\\arcsec$) than to TH2. %is not consistent with TH2, but lies closer to the IR peak within their %resolution. %could be associated within the errors with the IR peak. However, kinematic studies of galaxies focused on the gaseous component (ionized and/or molecular gas) have the potential for revealing the dynamical forces that work in the gas closer to the galactic nucleus, but they cannot trace adequately the gravitational potential. The gas represents only a fraction of the total mass of the galaxy and could be driven by non-gravitational perturbations such as interactions, warps, bars, etc. Therefore, the kinematic parameters of any nuclear galactic disk (center, position angle and inclination) can be greatly influenced by these perturbations. Thus the velocity field of stars is needed to properly trace the gravitational potential. Integral-field spectroscopy is an ideal tool for studying objects with complex morphologies and kinematics, such as the nuclei of starburst galaxies. This paper is the first of two describing the conclusions from an integral-field spectroscopic study with SINFONI on the VLT to clarify the structure, kinematics and emission mechanisms of the central 150 pc starburst region of NGC253. In this part we use the CO stellar absorption features in $K-$band to determine the stellar kinematic center of the galaxy. The stellar CO features have the advantage that they trace a component of the galaxy nucleus that is not affected by non-gravitational processes such as shocks or winds. %In order to reveal its nature, We also obtained archival \\textit{Chandra} data and re-processed them using $0.125\\arcsec$ pixel$^{-1}$. The X-ray images are carefully compared with available high spatial resolution radio and near-IR data. This multiwavelength approach combined with the estimate of the kinematic center is essential to find the true nucleus of NGC253. %We carefully aligned this new X-ray images with high spatial resoultion radio %and near-IR images. As a result we will be able to identify individual %counterparts to the %As neither broad recombination lines nor forbidden high ionization lines have %been observed in the nuclear spectra of this galaxy, AGN activity can be %identified from a point-like X-ray nuclear source, or a flat-spectrum radio %core. In a forthcoming publication \\citep{mueller09} we investigate the physical properties and kinematics of the super star clusters and the molecular and ionized gas in the nuclear starburst. %We address the star formation process, stellar populations, masses of individual %star clusters and the kinematics of the superwind. This paper is structured as follows. In Section 2, we describe the SINFONI and \\textit{Chandra} observations and the processing we performed on the data. The method used to extract the kinematics from the SINFONI spectral features is presented in this Section as well as the new X-ray images and spectra. In Section 3 we examine the stellar velocity field and obtain an estimate of the position of the kinematic center. We discuss the implications of our results for the nature of the galactic nucleus in Section 4. The overall conclusions of the study are outlined in Section 5. ", "conclusions": "\\label{conclusions} SINFONI near-infrared integral field spectroscopy combined with \\textit{Chandra} imaging and spectroscopy have been used in this work to probe the galactic nucleus of NGC253. Thanks to our high S/N integral field data we were able to extract 2D stellar kinematics in the nuclear region of NGC253 and estimate the kinematic center of the galaxy with a very good accuracy. Our main results and conclusions can be summarized as follows: \\begin{itemize} \\item The stellar kinematic center is located in between the strongest compact radio source (TH2) and the IR photometric center, or IR peak. It lies $\\sim 0.7 \\arcsec$ SW from TH2 and $\\sim 2.6 \\arcsec$ NE from the IR peak. \\item The estimated confidence interval of our kinematic modeling ($r\\sim 1.2\\arcsec$) clearly discards the IR peak as kinematic center, but includes the position of TH2. \\item Newly processed \\textsl{Chandra} images reveal a central point-like hard X-ray source (X-1) lying $\\sim0.4 \\arcsec$ SW from the kinematic center. %and two off-center X-ray sources which are good Ultraluminous X-ray sources (ULX) %candidates. \\item Very accurate alignment between radio, IR and X-ray sources in the nuclear region shows that TH2 and X-1 are not associated with each other, being separated by $\\sim 1.1 \\arcsec$. As the two could be a manifestation of nuclear activity, and our current estimate of the kinematic center location is consistent with both sources, we consider the two as possible galactic nucleus candidates. \\item TH2 is the brightest radio source, with the highest brightness temperature, detected with the VLA in the central region of NGC253. However, it does not have any IR, optical or X-ray counterparts. When observed at very high spatial resolution with the VLBI, TH2 is among the weakest sources ($\\sim 6$ mJy at 2.3 GHz, Lenc \\& Tingay 2006), or it is not detected ($<1$ mJy at 23 GHZ, Brunthaler et al. 2009). If the kinematic center is associated with TH2, we conclude that the nature of the nucleus of NGC253 resembles that of SgrA$^*$. \\item X-1 is a heavily absorbed object ($N_{H} > 10^{23}$ cm$^{-2}$) only detected at energies $>2$ keV ($L_{2-10 keV} \\sim 10^{40}$ ergs s$^{-1}$). If X-1 corresponds to the kinematic center of the galaxy, the simplest explanation accounting for these properties is that the nucleus of NGC253 harbors a hidden LLAGN. \\item Alternatively, X-1 may be associated with one SSC in this region: SSC IR-11/radio-TH6. In this case the kinematic center of NGC253 would then be a $10^5-10^6$ M$_\\odot$ spatially resolved (diameter of $\\sim5$ pc) NSC in which some of its massive stars are evolving into the SN phase. \\end{itemize} %This study demonstrated the power of integral field spectroscopy and stellar %kinematics for the determination of kinematic centers of galaxies. %suggesting that this source is likely to be a SSC. %In any of the two cases, we can conclude that there is a massive central object %presenting signatures of low accretion-powered activity. %This allows us to conclude %that NGC253 is basically a galaxy in which a strong starburst %and a weak AGN (either TH2 or X-1) coexist. We speculate that the galactic nucleus %of NGC253 is in an evolutionary state where it is transitioning between a %starburst phase and an AGN phase. Currently the intense multiple starbursts are %young with ages $<35$ Myr \\citep{mueller09}. Previous studies of starburst-AGN %connection hint at a delay of $50-100$ Myr between the onset of star formation and %subsequent fuelling of the black hole \\citep{davies07}. The galactic nucleus of %NGC253 is thus in a latent phase of activity, still lacking of sufficient fuel. An important conclusion from this study is that the nature of dusty galaxies cannot be always characterized by photometric data alone but must be complemented with spectroscopic observations. In particular, we demonstrated the power of integral field spectroscopy for the determination of kinematic centers of galaxies. Further near-IR integral-field observations with high spatial resoultion would be useful to derive a more accurate position of the kinematic center, measure directly the mass of the putative central massive object, and reveal the presence of either broad recombination lines or forbidden high ionization lines. A detection of any of these would provide definite evidence of the existence of an AGN in this galaxy. %% If you wish to include an acknowledgments section in your paper, %% separate it off from the body of the text using the" }, "1005/1005.3951_arXiv.txt": { "abstract": "The emission of neutrinos within a wide energy range is predicted from very-high-energy phenomena in the Universe. Even the current or next-generation Cherenkov neutrino telescopes might be too small to detect the faint fluxes expected for cosmic neutrinos with energies exceeding the EeV scale. The acoustic detection method is a promising option to enlarge the discovery potential in this highest-energy regime. In a possible future deep-sea detector, the pressure waves produced in a neutrino interaction could be detected by a $\\gtrsim 100\\,$km$^3$-sized array of acoustic sensors, even if it is sparsely instrumented with about 100\\,sensors/km$^3$. This article focuses on the AMADEUS set-up of acoustic sensors, which is an integral part of the ANTARES detector. The main aim of the project is a feasibility study towards a future acoustic neutrino detector. However, the experience gained with the ANTARES-AMADEUS hybrid opto-acoustic set-up can also be transferred to future very large volume optical neutrino telescopes, especially for the position calibration of the detector structures using acoustic sensors. ", "introduction": "At ultra-high energies (UHE, $E \\gtrsim 10^{18}\\,$eV), neutrinos are the only viable messengers beyond the ``local'' universe. Photons interact with the microwave and infrared background light. Protons suffer energy losses due to photo-production of pions (GZK mechanism \\cite{grei,zats}) and of e$^+$e$^-$ pairs; nuclei are additionally subject to photo-nuclear reactions. All these reactions confine undisturbed propagation of UHE cosmic rays to distances below 100\\,Mpc \\cite{allard}. UHE neutrinos, on the other hand, are undisturbed and ``guaranteed'' by the GZK mechanism. The interaction rate of neutrinos originating from the propagation of UHE protons in 1\\,km$^3$ water equivalent has been estimated to be about 0.2 per year \\cite{seckel}. Consequently, target masses exceeding 100\\,km$^3$ are required to obtain a few neutrino events per year. Such detector sizes are not reachable with current technologies and new detection techniques have to be considered for the study of UHE cosmic neutrinos, e.g.\\ the acoustic method discussed here. The investigation of acoustic neutrino detection has historically evolved in the context of Cherenkov neutrino {te\\-le\\-scopes}. The method was discussed in the 1970s within the DUMAND project \\cite{bowen}. Acoustic set-ups have been integrated in the framework of existing Cherenkov experiments in the 2000s: AMADEUS in {ANTA\\-RES}, ONDE in NEMO, SPATS in IceCube and acoustic test set-ups in the lake Baikal neutrino telescope\\footnote{There is a variety of additional ongoing projects - not only in sea water but also in ice and salt, cf.\\ \\cite{arena} for an overview.}. A next step towards an acoustic neutrino detector could be the integration of acoustic sensors into a next-generation neutrino telescope like KM3NeT \\cite{km3net}. ", "conclusions": "Cosmic neutrinos are the only viable UHE messengers beyond the ``local'' universe. The acoustic neutrino detection technique is an option to extend neutrino astrophysics to the extreme energy range, either in stand-alone detectors or in combined opto-acoustic sensor arrays. The AMADEUS system, dedicated to the investigation of this technique, has been successfully integrated into the ANTA\\-RES neutrino telescope. Except for size, the system has all features required for an acoustic detector and thus allows for deciding on the feasibility of neutrino detection with a potential future large acoustic sensor array. In the context of AMADEUS it is currently being investigated whether an integrated opto-acoustic sensor configuration has the potential to combine the position calibration of an optical detector with acoustic detection studies and marine science applications." }, "1005/1005.1679_arXiv.txt": { "abstract": "In this paper we test a special-relativistic formulation of Smoothed Particle Hydrodynamics (SPH) that has been derived from the Lagrangian of an ideal fluid. Apart from its symmetry in the particle indices, the new formulation differs from earlier approaches in its artificial viscosity and in the use of special-relativistic ``grad-h-terms''. In this paper we benchmark the scheme in a number of demanding test problems. Maybe not too surprising for such a Lagrangian scheme, it performs close to perfectly in pure advection tests. What is more, the method produces accurate results even in highly relativistic shock problems. ", "introduction": "\\label{Rosswog::sec:1} Relativity is a crucial ingredient in a variety of astrophysical phenomena. For example the jets that are expelled from the cores of active galaxies reach velocities tantalizingly close to the speed of light, and motion near a black hole is heavily influenced by space-time curvature effects. In the recent past, substantial progress has been made in the development of numerical tools to tackle relativistic gas dynamics problems, both on the special- and the general-relativistic side, for reviews see \\cite{Rosswog:marti03,Rosswog:font00,Rosswog:baumgarte03}. Most work on numerical relativistic gas dynamics has been performed in an Eulerian framework, a couple of Lagrangian smooth particle hydrodynamics (SPH) approaches do exist though.\\\\ In astrophysics, the SPH method has been very successful, mainly because of its excellent conservation properties, its natural flexibility and robustness. Moreover, its physically intuitive formulation has enabled the inclusion of various physical processes beyond gas dynamics so that many challenging multi-physics problems could be tackled. For recent reviews of the method we refer to the literature \\cite{Rosswog:monaghan05,Rosswog:rosswog09b}. Relativistic versions of the SPH method were first applied to special relativity and to gas flows evolving in a fixed background metric \\cite{Rosswog:kheyfets90,Rosswog:mann91,Rosswog:mann93,Rosswog:laguna93a,Rosswog:chow97,Rosswog:siegler00a}. More recently, SPH has also been used in combination with approximative schemes to dynamically evolve space-time \\cite{Rosswog:ayal01,Rosswog:faber00,Rosswog:faber01,Rosswog:faber02b,Rosswog:oechslin02,Rosswog:faber04,Rosswog:faber06,Rosswog:bauswein10}.\\\\ In this paper we briefly summarize the main equations of a new, special-relativistic SPH formulation that has been derived from the Lagrangian of an ideal fluid. Since the details of the derivation have been outlined elsewhere, we focus here on a set of numerical benchmark tests that complement those shown in the original paper \\cite{Rosswog:rosswog09d}. Some of them are ``standard'' and often used to demonstrate or compare code performance, but most of them are more violent---and therefore more challenging---versions of widespread test problems. ", "conclusions": "We have summarized a new special-relativistic SPH formulation that is derived from the Lagrangian of an ideal fluid \\cite{Rosswog:rosswog09d}. As numerical variables it uses the canonical energy and momentum per baryon whose evolution equations follow stringently from the Euler-Lagrange equations. We have further applied the special-relativistic generalizations of the so-called ``grad-h-terms'' and a refined artificial viscosity scheme with time dependent parameters.\\\\ The main focus of this paper is the presentation of a set of challenging benchmark tests that complement those of the original paper \\cite{Rosswog:rosswog09d}. They show the excellent advection properties of the method, but also its ability to accurately handle even very strong relativistic shocks. In the extreme shock tube test 3, where the post-shock density shell is compressed into a width of only 0.1 \\% of the computational domain, we find the shock front to propagate at slightly too large a pace. This artifact ceases with increasing numerical resolution, but future improvements of this point would be desirable. We have further determined the convergence rate of the method in numerical experiments and find it first-order accurate when shocks are involved and second-order accurate for smooth flows. \\vspace*{0.5cm} {\\em Acknowledgment}\\\\ This work was supported by the German Research Foundation under grant number 50245 DFG-RO-5." }, "1005/1005.5570_arXiv.txt": { "abstract": "We present the second report of our systematic search for strongly lensed quasars from the data of the Sloan Digital Sky Survey (SDSS). From extensive follow-up observations of 136 candidate objects, we find 36 lenses in the full sample of 77,429 spectroscopically confirmed quasars in the SDSS Data Release 5. We then define a complete sample of 19 lenses, including 11 from our previous search in the SDSS Data Release 3, from the sample of 36,287 quasars with $i<19.1$ in the redshift range $0.6$ 1 M$_{\\odot}$, as classical nova models imply, the donor mass must be even higher. We therefore rule out the possibility that T Pyx has evolved beyond the period minimum for cataclysmic variables. We find that the system inclination is constrained to be $i \\approx 10$ degrees, confirming the expectation that T Pyx is a low-inclination system. We also discuss some of the evolutionary implications of the emerging physical picture of T Pyx. In particular, we show that epochs of enhanced mass transfer (like the present) may accelerate or even dominate the overall evolution of the system, even if they are relatively short-lived. We also point out that such phases may be relevant to the evolution of cataclysmic variables more generally. ", "introduction": "Binary systems containing a Roche-lobe filling main-sequence star that loses mass to a primary white dwarf (WD) are referred to as cataclysmic variables (CVs, see~\\citealp{b28} for an overview). T Pyx is a luminous CV in which the donor star is transferring mass at a very high rate onto a high-mass white dwarf, resulting in unusually frequent thermo-nuclear runaways (TNRs) on the surface of the primary. Between the years 1890 $-$ 1967, T Pyx has undergone five such nova eruptions, with an recurrence time of about 20 years between the eruptions, and was therefore classified as a member of the recurrent nova (RN) subclass. However, the last eruption was in 1966, which means that T Pyx has now passed its mean recurrence time by more than 20 years. The eruptive behavior of RNe in comparison with classical novae is thought to be due to a high mass-transfer rate in combination with a massive primary white dwarf ~\\citep{b31}. ~\\cite{b13} carried out an extensive photometric study of T Pyx and found a stable, periodic signal at $P=1.83$ hours that was interpreted as the likely orbital period. This would place T Pyx below the CV period gap and suggests a donor mass around $M_{2} \\sim 0.1$ M$_{\\odot}$. This is surprising. According to standard evolutionary models, a CV below the period gap should be faint and have a low accretion rate driven primarily by gravitational radiation (GR). Yet T Pyx's quiescent luminosity and status as a RN both imply that it has a high accretion rate of $> 10^{-8}$ M$_{\\odot}$ yr$^{-1}$ (\\citealp{b13};~\\citealp{b20}). This is about two orders of magnitude higher than expected for ordinary CVs at this period. Assuming that the determination of the photometric orbital period is correct, the existence of T Pyx is interesting for at least two reasons. First, unless we are seeing the system in a transient evolutionary state, its lifetime would have to be very short $\\tau \\sim M_{2}/\\dot{M}_{2} \\leappeq 10$~Myrs. This would imply the existence of an evolutionary channel leading to the fast destruction of at least some short-period CVs. Second, TNRs frequent enough to qualify as RNe are thought to be possible only on high-mass accreting WDs ($M_{1} \\gtappeq 1$ M$_{\\odot}$). Moreover, RNe are the only class of novae in which the WD is expected to gain more mass between eruptions than it loses during them. This would make T Pyx a strong candidate Type Ia supernova progenitor. However, the recent study of the system by~\\cite{b18} (see also~\\citealp{b20}) suggests, first, that T Pyx {\\em is}, in fact, in a transient evolutionary state, and, second, that, integrated over many nova eruptions, its WD does lose more mass than it gains. More specifically,~\\cite{b18} suggest that T Pyx was an ordinary cataclysmic variable until it erupted as a nova in 1866. This eruption triggered a wind-driven supersoft X-ray phase (as first suggested by~\\citealp{b7}), resulting in an unusually high luminosity and accretion rate. However, unlike in the original scenario proposed by~\\cite{b7}, the supersoft phase is not self-sustaining, so that the accretion rate has been declining ever since the 1866 nova eruption from $\\dot{M} \\sim 10^{-7}$ M$_{\\odot}$\\,yr$^{-1}$ to $10^{-8} $M$_{\\odot}$\\,yr$^{-1}$. As a result, T Pyx has faded by almost 2 magnitudes since the nova eruption (\\citealp{b18}). Based on this, and the fact that T Pyx has already passed its mean recurrence time by more than 20 years,~\\cite{b18} argue that T Pyx might no longer even be a recurrent nova. If these ideas are correct, T Pyx is not a viable SN Ia progenitor, and its remaining lifetime can be substantially longer than a few million years. However, if all its ordinary nova eruptions are followed by relatively long-lived (> 100 yrs) intervals of wind-driven evolution at high $\\dot{M}$, its secular evolution may nevertheless be strongly affected, with significant implications for CV evolution more generally (see also Section~\\ref{SD}). A key assumption in virtually all of these arguments is that the photometric period measured by~\\cite{b13} is, in fact, the orbital period of the system. So far, there has only been one attempt to obtain a spectroscopic period for T Pyx, by~\\cite{b26}, who reported a spectroscopic modulation with $P=3.44$ hours. Such a long orbital period above the CV period gap would be much more consistent with the high accretion rate found in T Pyx. In this study, we present the first definitive spectroscopic determination of the orbital period of T Pyx, showing that it is, in fact, consistent with Patterson et al.'s photometric period. We also use our time-resolved spectroscopy to estimate the main system parameters, such as the velocity semi-amplitude of the white dwarf ($K1$), the mass-ratio (q), the masses ($M_{1}$ and $M_{2}$) and the orbital inclination ($i$). Finally, we discuss the implications of our results for the evolution of T Pyx and related systems. ", "conclusions": "\\label{SD} The main result of our study is the spectroscopic determination of T Pyx's orbital period, $P_{orb} \\simeq 1.83$~hrs. This confirms that the system is a CV below the period gap and implies that its current accretion rate is at least 2 orders of magnitude higher than that of an ordinary CV at this period. We also find that our spectroscopic orbital period is consistent with the photometric ephemeris found for T Pyx (\\citealp{b13}, an updated version is given in Section~\\ref{PE}). This means not only that photometric timings can be used as a more precise and convenient tracer of the orbital motion, but also that the large period derivative required by the photometric ephemeris marks a genuine change in the orbital period of the system. In fact, the spectroscopic data are consistent with the photometric period only if the period derivative is accounted for. The period derivative obtained here from the latest combined photometric data ($\\dot{P} = 6.7 \\times 10^{-10}$) is slightly lower than that obtained by~\\cite{b13} from data up to 1997 ($\\dot{P} \\simeq 9 \\times 10^{-10}$). A decrease in the rate of period change would be in line with~\\cite{b18} scenario that T Pyx's days as a high-$\\dot{M}$ recurrent nova are numbered (at least until its next ordinary nova eruption). In any case, the average time-scale for period change found across all the photometric data are about $3 \\times 10^5$~yrs. We have also used our spectroscopic data to obtain estimates of other key system parameters, most notably the radial velocity semi-amplitude of the WD, $K1 = 17.9 \\pm 1.6$ km\\,s$^{-1}$, and the mass ratio, $q = 0.20 \\pm 0.03$. The latter estimate rests on three key assumptions: first, that our determination of $K1$ is correct, second, that our estimate $v(R_{disk})\\sin i$ is correct, and, third, that the accretion disk around the WD in T Pyx is tidally limited. Taken at face value, this relatively high value of the mass ratio implies that the donor star in the system is not a brown dwarf. Thus T Pyx is not a period bouncer. If we assume that the radius of the secondary is $10\\% \\pm 10\\%$ inflated relative to an ordinary main sequence star of the same mass, we find that its most likely mass is $M_{2} = 0.14 \\pm 0.03$ M$_{\\odot}$. Overall, the physical picture that emerges from our study is consistent with the scenario proposed by~\\cite{b18}. In particular, they suggest that, prior to the 1866 eruption, T Pyx was an ordinary CV. That eruption then triggered a high-$\\dot{M}$ wind-driven phase, as suggested by~\\cite{b7} to account for T Pyx's exceptional luminosity. However, this phase is not quite self-sustaining, so that T Pyx is now fading and perhaps not even a RN anymore. In line with this picture, we find that the mass ratio and donor mass we derive are not abnormally low for a CV at its orbital period. This shows that the present phase of high-$\\dot{M}$ accretion cannot have gone on for too long already. The hint of a declining period derivative may point in the same direction, but this needs to be confirmed. Does all this mean that the phase of extraordinarily high accretion rates T Pyx is currently experiencing will have no lasting impact on its evolution? Not necessarily. In a stationary Roche-lobe-filling system, the orbital period derivative and total mass-loss rate from the donor (wind loss + mass transfer) are related via \\begin{equation} \\frac{\\dot{P}_{orb}}{P_{orb}} = \\frac{3 \\zeta -1}{2} \\frac{\\dot{M}_2}{M_2} \\end{equation} where $\\zeta$ is the donor's mass-radius index. In T Pyx, which has an increasing orbital period, we will take $\\zeta \\simeq -1/3$, which is appropriate for adiabatic mass-loss from a fully convective star (e.g.~\\citealp{b7}). We thus expect that $\\dot{M_2}/M_2 \\simeq {\\dot{P}_{orb}}/{P_{orb}}$, which suggests a typical mass-loss rate from the donor of $\\dot{M_2} \\sim 5\\times10^{-7}$~M$_{\\odot}$/yr in its current state. (The accretion rate onto the WD can be lower than this, since, in the wind-driven scenario, much of this mass escapes in the form of an irradiation-driven outflow from the donor.) If every ordinary nova eruption in T Pyx is followed by $\\sim$ 100 yrs of such high-$\\dot{M}_2$ evolution, the total mass loss from the donor in the luminous phase is $\\sim 5 \\times 10^{-5}$ M$_{\\odot}$ between any two such eruptions. This needs to be compared to the mass lost from the donor during the remaining part of the cycle. If this is driven by gravitational radiation, the mass loss rate from the donor will be $\\dot{M}_2 \\simeq 5 \\times 10^{-11}$~M$_{\\odot}$/yr~\\citep{b7}. The recurrence time of ordinary nova eruptions for such a system is on the order of $10^5$ yrs~\\citep{b31}, so the total mass lost from the donor during its normal evolution (outside the wind-driven phase) is $5 \\times 10^{-6}$ M$_{\\odot}$. {\\em This shows that the long-term secular evolution may be dominated by its high-$\\dot{M}$ wind-driven phases, even if the duty cycle of these phases is very low (e.g. 0.1\\% for the numbers adopted above).} It is finally tempting to speculate on the relevance of ``T Pyx-like'' evolution for ordinary CVs. At first sight, the numbers above suggest that the evolution of a CV caught in such a state may be accelerated by about an order of magnitude. This is interesting, since it could bear on the long-standing problem that there are fewer short-period CVs and period bouncers in current samples than theoretically expected (e.g.~\\citealp{b13};~\\citealp{b14};~\\citealp{b15}; Pretorius \\& Knigge 2008ab). Moreover, this channel need not be limited to systems containing high-mass WDs that are capable of becoming RNe. After all, it is not the recurrent nova outbursts that are of interest from an evolutionary point of view, but simply the existence of a prolonged high-$\\dot{M}$ phase in the aftermath of nova eruptions. In CVs containing lower-mass WDs, such a phase may still occur, although its evolutionary significance could still depend on the WD mass. For example, the interoutburst time-scale is longer for low-mass WDs, and the duration of the high-$\\dot{M}$ phase could also scale with WD mass. One obvious objection to this idea is that, observationally, most nova eruptions are not followed by centuries- (or even decades-) long high-$\\dot{M}$ phases. However, this need not be a serious issue. Most observed novae are long-period systems, so if the triggering of a wind-driven phase requires a fully convective donor, most novae would not be expected to enter such a phase. It may be relevant in this context that at least one other short-period nova -- GQ Mus -- exhibited an exceptionally long post-outburst supersoft phase of $\\sim 10$~years~\\citep{b12}.~\\cite{b3} have also suggested that the duration of the supersoft phase in novae may scale inversely with orbital period. However, there is another important consequence to the idea that the evolution of many short-period CVs is accelerated by ``T Pyx-like'' high-luminosity phases. The orbital period-derivative is positive in the high-$\\dot{M}$ phase, but negative during the remaining times of GR-driven evolution. But since $\\dot{P}_{orb}/P_{orb} \\simeq \\dot{M}_2/M_2$ in {\\em both} phases, the sign of the secular (long-term-average) period derivative will generally correspond to the phase that dominates the secular evolution. So whenever wind-driving dramatically accelerates the binary evolution, the direction of period evolution will also be reversed. Is this a problem? Perhaps. The recent detection of the long-sought {\\em period spike} in the distribution of CV orbital periods~\\citep{b4} suggests that there is, in fact, a reasonably well-defined minimum period for CVs, as has long been predicted by the standard model of CV evolution (e.g.~\\citealp{b9}). On the other hand, the location of the observed spike ($\\simeq 83$~mins) is significantly different from the expected one ($\\simeq 65-70$~min). Is it possible that the observed period minimum corresponds to the onset of wind-driving in most CVs? T Pyx, with $P_{orb} \\simeq 110$~min would then have to be an outlier, however, perhaps because of an unusually high WD mass. The idea that T Pyx-like phases may significantly affect the evolution of many CVs is, of course, highly speculative, and we do not mean to endorse it too strongly. However, it highlights the importance of understanding T Pyx: until we know what triggered the current high-luminosity state, it will remain difficult to assess the broader evolutionary significance of this phase. Note that the apparent uniqueness of T Pyx is not a strong argument against such significance. For example, if the duty cycle of high-luminosity phases is $\\sim 0.1$\\%, as suggested by the numbers above, we should not expect to catch many CVs in this state. Thus T Pyx could be the tip of the proverbial iceberg." }, "1005/1005.3377_arXiv.txt": { "abstract": "{ The issue of the influence of coronal holes (CHs) on coronal mass ejections (CMEs) in causing solar energetic particle (SEP) events is revisited. It is a continuation and extension of our previous work \\citep{Shen_etal_2006}, in which no evident effect of CHs on CMEs in generating SEPs were found by statistically investigating 56 CME events. This result is consistent with the conclusion obtained by Kahler in 2004. In this paper, we extrapolate the coronal magnetic field, define CHs as the regions consisting of only open magnetic field lines and perform a similar analysis on this issue for totally 76 events by extending the study interval to the end of 2008. Three key parameters, CH proximity, CH area and CH relative position, are involved in the analysis. The new result confirms the previous conclusion that CHs did not show any evident effect on CMEs in causing SEP events. ", "introduction": "Gradual solar energetic particle (SEP) events are thought to be a consequence of CME-driven shocks generating plenty of SEPs which would be observed near the Earth. In our previous work in 2006, we statistically studied the effect of coronal holes (CHs) on the CMEs causing SEP events by investigating the CME source locations and their relation with the CHs identified in EUV 284 \\AA$ $\\citep[][ hereafter Paper I]{Shen_etal_2006}. It was implied that neither CH proximity nor CH relative location exhibits any evident effect on the intensities of SEP events. This result is consistent with the conclusion obtained by \\citet{Kahler_2004}, who comparatively studied the SEP events produced in the fast and slow solar wind streams and found no significant bias against SEP production in fast-wind regions which are believed to originate from CHs. These findings seem not quite fit people's `common sense', because CHs are believed to be regions with low-density and low temperature in the corona \\citep[e.g.][]{Harvey_Recely_2002}, from which the solar wind is fast and the magnetic field is open, and therefor apparently three disadvantages for a CME to produce SEP may exist when it is near a coronal hole region. These advantage are: (1) the background solar wind speed $V_{sw}$ near CHs is larger than that in other regions; (2) the plasma density near CHs is much lower than that in other regions, so that the Alfv\\'{e}n speed $V_a$ is larger \\citep{Shen_etal_2007,Gopalswamy_etal_2008}; and (3) the magnetic field lines in CHs are open. The first two disadvantages suggest that a strong shock might be hardly produced near CHs. The third one implies that particles might be able to escape from the shock acceleration process earlier and easier. Thus, it can be expected that CHs would influence the CME in producing SEP events. The work by \\citet{Kunches_Zwickl_1999} was consistent with the picture depicted above. In their paper, they found that the CH may delay the onset times of SEPs when a CH is present between Sun-observer line and the solar source of the SEP event. They also speculate that the peak intensity could be influenced by CH. However, they did not statistical study such influence. It is hard to say that their conclusion is statistically significant. In principle, CHs are open field regions, though they were first identified in observations \\citep[e.g.][]{Zirker_1977}. \\citet{Kunches_Zwickl_1999} identified CHs based on He 10830 \\AA$ $. In our 2006 work (Paper I), CHs were auto-determined based on EUV 284\\AA$ $ images taken by SOHO/EIT. Thus, it is doubtable whether or not the CHs identified in EUV wavelengths really represent open field regions. Another doubt in our 2006 work is that only frontside CHs are taken into account. In order to remove the doubt and get a more reliable result, we look into this topic again by extrapolating coronal magnetic field instead of analyzing EUV images. The term `CHs' in this paper therefore actually refers to open field regions. The magnetic field extrapolation and determination of CHs are introduced in Section 2. Section 3 presents the statistical analysis. A brief summary and conclusions are given in Section 4. ", "conclusions": "In order to study the influence of CHs on CMEs in producing SEP events, a total of 76 west-side fast halo CMEs during 1997 - 2008 are investigated, as well as their associated CHs. Different from the CHs obtained by brightness method based on EIT 284\\AA$ $ data in paper I, the CHs we investigated in this paper are obtained with the aid of the extrapolation of coronal magnetic field by CSSS model, in which the MDI daily-updated synoptic magnetic field charts are adopted as the bottom boundary condition. By using this method, all the CHs, defined as the regions consisting of open magnetic field lines only, over the entire solar surface are inferred. After analyzing three parameters, CH proximity, area of corresponding CHs and relative position between CHs and CMEs, it is found that all of the statistical results do NOT have significance exceeding the 1$\\sigma$ level. These parameters do NOT show any evident influence on SEP occurrence probability, and the speed of SEPYCMEs also do NOT show any difference between different groups binarized by these parameters. These results confirmed the conclusion we got in Paper I and \\citet{Kahler_2004} that no evident influence of CHs on CME in producing SEP events. An expanding CME may drive a quasi-parallel shock at its flank as discussed by \\citet{Kahler_2004}. The condition of CME in driven shock in this situation is $V_{cme}$ larger than local alf\\'{v}en speed $V_a$ or sound speed $C_s$ only. Thus, the fast flow speed near CHs may show no influence on producing strong shock. Beside, not only the plasma density but also the magnetic field strength in fast solar wind region is smaller than them in slow solar wind region\\citep{Ebert_etal_2009}, so the alf\\'{v}en speed in fast solar wind region may not obvious faster than it in slow solar wind region. Based on the these analysis, it could be expected that shock can also be produced in fast solar wind region near CH and no evident fast of the CME needed. In addition, the shock interact with background solar wind may generate a turbulence. Such turbulence could be treated as the main mechanism that makes particles back to shock acceleration process to produce SEP events\\citep{Reames_1999}. The close magnetic topology could only provide an addition method to make the particle back to shock acceleration\\citep{Shen_etal_2008}. So, the influence of open magnetic field topology may weak in shock producing SEP events. \\normalem" }, "1005/1005.4557_arXiv.txt": { "abstract": "{While sub-micron- and micron-sized dust grains are generally well mixed with the gas phase in protoplanetary disks, larger grains will be partially decoupled and as a consequence have a different distribution from that of the gas. This has ramifications for predictions of the observability of protoplanetary disks, for which gas-only studies will provide an inaccurate picture. Specifically, criteria for gap opening in the presence of a planet have generally been studied for the gas phase, whereas the situation can be quite different in the dust layer once grains reach mm sizes, which is what will be observed by ALMA.} {We aim to investigate the formation and structure of a planetary gap in the dust layer of a protoplanetary disk with an embedded planet.} {We perform 3D, gas$+$dust SPH simulations of a protoplanetary disk with a planet on a fixed circular orbit at 40 AU to study the evolution of both the gas and dust distributions and densities in the disk. We run a series of simulations in which the planet mass and the dust grain size varies.} {We show that the gap in the dust layer is more striking than in the gas phase and that it is deeper and wider for more massive planets as well as for larger grains. For a massive enough planet, we note that cm-sized grains remain inside the gap in corotation and that their population in the outer disk shows an asymmetric structure, a signature of disk-planet interactions even for a circular planetary orbit, which should be observable with ALMA.} {} ", "introduction": "\\label{sec:introduction} While we understand the general scenario of planet formation via accretion \\citep{Mizuno1980, Wetherill1980, Lissauer1987, Wetherill1990, Pollack_etal1996}, the devil is in the detail and the processes by which tiny sub-micron grains grow into planetesimals, the building blocks of planets, is not well understood. With over 400 extrasolar planets detected to date as listed in the extrasolar planet encyclopedia\\footnote{http://exoplanet.eu/}, we can start to do some meaningful statistics on the types of planets that exist in our galaxy \\citep{AMBW05,M07} to help constrain theories of planet formation. A wealth of analytical as well as numerical studies of the formation of gaps by a planet in a gas disk have shown how the shape of the gap depends on properties of the planet (mass) as well as the gaseous disk (pressure scale height, viscosity). See \\cite{CMM06} for the case of gap shapes and \\cite{PNKMA07} for a more general review on planet migration and gap formation. The dust phase, however, has been shown to behave differently to the gas. Dust experiences a headwind from the pressure-supported sub-Keplerian gas and the induced drag force slows the dust and makes it settle to the midplane and migrate inwards. The magnitude of these effects depends strongly on the grain size and the disk density \\citep{W77,SV96,SV97,GBFL04,GL04,BF05}. In recent years, gap formation by a planet embedded in disks of gas and dust has been studied by several authors, for different planet masses and different sizes of the solid particles. \\citet{PM04,PM06} determined the spatial distribution of 1~mm grains with 2D simulations in order to derive the smallest planet mass that would result in an observable gap with ALMA (Atacama Large Millimeter Array). In their work the dust was strongly coupled to the gas and responded indirectly to the planet gravity through the radial pressure gradients caused by a Neptune-mass planet in the gas disk that led to the formation of a gap. \\citet{Muto09} investigated the effect of a low-mass planet on the dust distribution by injecting one dust grain at a time and derived criteria for gap opening. \\citet{Ciecielag07} as well as \\citet{Marzari00} focused on already formed planetesimals while the gas phase is still present. \\citet{Ciecielag07} considered planetesimals down to 1~m in size. They studied the effect of spiral structures in the circumprimary gas disk triggered by a secondary companion in a tight binary system in order to derive relative velocities between planetesimals and determine whether planet formation is possible in such systems. The presence of planets, and the gap they create when they are massive enough, can also help constrain the global properties of the gas (temperature, density, viscosity) as well as those of the dust (grain size distribution, degree of settling) in the disks. This can be achieved by measuring the width and brightness of the gap, ideally for each phase \\citep{W05,CMM06,Fouchet07}. It has been shown that ALMA will be able to observe a planetary gap opened by a 1~$M_J$ planet at 5.2~AU from a 1~$M_\\odot$ star at a distance of 140~pc under the naive assumption that gas and dust are well mixed \\citep{W02}. It will certainly be possible to observe other features related to the presence of the planet, such as warm dust in its vicinity or even spiral waves for distances not exceeding 100~pc \\citep{W05}. Here we focus on the formation and features of the gap itself in the case of a massive protoplanet. Our previous simulations of dust evolution in a typical Classical T-Tauri Star (CTTS) disk \\citep[hereafter BF05]{BF05}\\defcitealias{BF05}{BF05} showed that the thickness of the dust layer depends on grain size because different sized grains fall to the midplane at different rates. We distinguished three dynamical regimes for the dust: (1) almost uncoupled for large grains where the dust component follows slightly perturbed Keplerian orbits and keeps its 3D distribution (if initially 3D); (2) weakly coupled for intermediate-sized grains for which settling is very efficient; (3) strongly coupled for small grains where grains are forced to follow the gas motion. We have also investigated the formation of a gap by a planet immersed in a Minimum Mass Solar Nebula (MMSN) and showed that dust settling makes the gap much more striking in the dust layer than in the gas phase because of its reduced vertical extension for weakly coupled grains \\citep[hereafter F07]{Maddison07,Fouchet07}\\defcitealias{Fouchet07}{F07}. Indeed, the criterion for gap formation depends on the disk scale height \\citep{CMM06}. Because of the size of particles in the weakly coupled regime for the particular case of the MMSN (1~m, see Sect.~\\ref{sec:theory}), that study had no direct application to observations. Images of dusty disks at infrared wavelengths do not probe such large particles, but instead trace the smaller, strongly coupled dust grains whose distribution is similar to that of the gas. In this paper, we examine the gap formation in the dust layer of CTTS disks, which are spatially more extended and less dense than the theoretical MMSN case. Our ultimate goal is to use the results of our hydrodynamical simulations to produce synthetic images for ALMA. We consider in particular the effects of a massive planet in the outer cooler regions of the disk. Indeed, several planets at large distances from their star have been detected, such as those recently announced orbiting Fomalhaut \\citep{Kalas08}, or HR~8799 \\citep{Marois08}. The extrasolar planet encyclopedia lists 10 planets with a semi-major axis larger than 20~AU, and 7 of them have minimum masses larger than 5~$M_\\mathrm{J}$. We study the dust distribution in the weakly coupled regime, which corresponds for these CTTS disks to a size range (100 $\\mu$m to 1 cm) that can directly be probed by current and future (sub)millimetre instruments. The present paper is the first part of this work, presenting the hydrodynamical simulations. In Sect.~\\ref{sec:theory}, we discuss the gas-dust interaction in the presence of a planet. In Sect.~\\ref{sec:simulations}, we describe the numerical method and simulation suite. Results are described in Sect.~\\ref{sec:results}, while the analysis and explanations are presented in Sect.~\\ref{sec:discussion}. We conclude in Sect.~\\ref{sec:conclusion}. In a forthcoming companion paper, we will use the simulations presented here to produce synthetic images for ALMA and present the most favorable observing configurations to detect the gap and associated structures. ", "conclusions": "\\label{sec:conclusion} We have run 3D, two-phase (gas and dust) SPH simulations of a typical CTTS disk of mass 0.02~$M_{\\sun}$ with an embedded giant planet on a circular orbit at 40~AU. We vary the grain size (100~$\\mu$m, 1~mm, 1~cm) and the planet mass (0.1, 0.5, 1, 5~$M_\\mathrm{J}$) and study the formation of the planetary gap. We confirm that gap opening is stronger in the settled dust layer than in the flared gas disk. Gaps are deeper and wider for (1) larger, more efficiently settled grains and (2) more massive planets. Larger planet masses are required to open a gap in the gas phase than in the dust: while a 0.5~$M_\\mathrm{J}$ planet only slightly affects the gas phase, it carves a deep gap in the dust. For the most massive 5~$M_\\mathrm{J}$ planet, 1~cm grains remain trapped in corotation with the planet while their distribution in the outer disk shows an asymmetric structure, even though the planet's orbit is circular. We find that this is not caused by the periastron alignment of a coherent set of orbits but rather by the pile-up of dust in the pressure maximum of the gas phase caused by the spiral density wave triggered by the planet. This global asymmetry does not appear for less massive planets because the spiral perturbation is substantially weaker. The variety of structures that we obtain in the dust phase for various grain sizes and planet masses has implications on the appearance of protoplanetary disks at (sub)millimetre wavelengths and show how important it is to go beyond the gas-only disk description proposed by, e.g., \\citet{W05}. The observability of these disks with ALMA is the subject of a forthcoming companion paper." }, "1005/1005.5551_arXiv.txt": { "abstract": "{ Comprehensive VLBI and multi-waveband monitoring indicate that a single superluminal knot can cause a number of $\\gamma$-ray flares at different locations. However, the often very rapid variability timescale is a challenge to theoretical models when a given flare (perhaps the majority of those observed) is inferred from observations to lie near the 43 GHz core, parsecs from the central engine. We present some relevant observational results, using the BL Lac object AO~0235+164 as an example. We propose a turbulent cell model leading to a frequency-dependent filling factor of the emission region. This feature of the model can provide a solution to the timescale dilemma and other characteristics of blazar emission. } ", "introduction": "There exists a crisis in the interpretation the results of multi-waveband monitoring of blazars. Comparisons of dates of peak $\\gamma$-ray flux measured by EGRET with epochs of ejections of superluminal radio knots (Jorstad et al.\\ \\cite{jor01a}), onsets of millimeter-wave flares (L\\\"ahteenm\\\"aki \\& Valtaoja \\cite{laht03}), and changes in centimeter-wave polarization (Jorstad et al.\\ \\cite{jor01b}) lead to the conclusion that most $\\gamma$-rays outbursts are coincident with radio events. Yet we know that the jets of blazars are opaque to radio emission within $\\sim 1$ pc or more from the central engine. This is difficult to reconcile with the shortest time scales of GeV variability observed --- a few hours in PKS~1622$-$297 (Mattox et al.\\ \\cite{mat97}) and 3C~454.3 (Foschini et al.\\ \\cite{fos10}). However time scales of variability reflect size scales, not distance from the central engine. Furthermore, jets are very narrow, with opening half-angle $\\sim 10^\\circ/\\Gamma$, where $\\Gamma$ is the bulk Lorentz factor of the flow in the jet (Jorstad et al. \\cite{jor05}). Because of this, a jet can have a cross-sectional radius much less than its distance from the central engine. Nevertheless, this does not solve the entire problem. The core on 43 GHz Very Long Baseline Array (VLBA) images of PKS~1510$-$089 has been estimated to lie $\\sim 20$ pc from the central engine (Marscher et al.\\ \\cite{mar10a}). Offsets of parsecs are implied for other luminous blazars as well (e.g., Marscher et al.\\ \\cite{mar08}; Chatterjee et al.\\ \\cite{chat08}). The cross-sectional radius of a luminous blazar is $\\sim 0.1$ pc near the 43 GHz core. With a Doppler factor $\\delta \\sim 20$ and redshift $\\sim 0.5$, the shortest time scale of variability should be $\\sim 1$ week, not a fraction of a day. Because of this discrepancy, a number of authors have insisted that the $\\gamma$-ray emitting region lies within $\\sim 10^{16}$ cm of the central engine (e.g., Tavecchio et al.\\ \\cite{tav10}). This has the advantage for high-energy emission models that optical-uv photons from the broad emission-line region are available for scattering to $\\gamma$-ray energies by highly relativistic electrons in the jet. Poutanen \\& Stern (\\cite{ps10}) find that pair production off the uv photon field can then cause the observed sharp break in the spectrum of some blazars at GeV energies. However, such an interpretation requires that the timing of the radio and $\\gamma$-ray events is a chance coincidence in every case. This seems unlikely, especially with singular events such as the ultra-high-amplitude optical flare, very sharp $\\gamma$-ray flare, and passage of a superluminal knot through the 43 GHz core in PKS~1510$-$089 at essentially the same time (Marscher et al.\\ \\cite{mar10a,mar10b}). Furthermore, the number of $\\gamma$-ray outbursts observed with {\\it Fermi} that are either coincident with or follow the passage of a new superluminal knot through the centroid of the core is becoming high enough to conclude that a large fraction of the $\\gamma$-ray emission originates parsecs away from the central engine (Jorstad et al. \\cite{jor10} and these proceedings). We propose to solve this dilemma by developing a model in which much of the optical and high-energy radiation in a blazar is emitted near the 43 GHz core in VLBA images, parsecs from the central engine. The model allows for short time scales of optical and $\\gamma$-ray variability by restricting the highest-energy electrons radiating at these frequencies to small sub-regions of the jet. That is, in our model the filling factor at high frequencies is relatively low, while that of the electrons radiating at $\\sim 10^{10-13}$ Hz is near unity. Such a model is consistent with other prominent features of optical vs.\\ lower frequency emission. ", "conclusions": "" }, "1005/1005.2237_arXiv.txt": { "abstract": "We investigate effects of hardness of primordial binaries on whole evolution of star clusters by means of $N$-body simulations. Using newly developed code, GORILLA, we simulated eleven $N=16384$ clusters with primordial binaries whose binding energies are equal in each cluster in range of $1-300kT_0$, where $1.5kT_0$ is average stellar kinetic energy at the initial time. We found that, in both soft ($\\le 3kT_0$) and hard ($\\ge 300kT_0$) limits, clusters experience deep core collapse. In the intermediate hardness ($10-100kT_0$), the core collapses halt halfway due to an energy releases of the primordial binaries. The core radii at the halt can be explained by their energy budget. ", "introduction": "Recently, observational informations concerning on the binary systems in the globular cluster have been accumulated from photometric observations of the eclipses (e.g. \\cite{Mateo96}), spectroscopic observations (e.g. \\cite{Albrow+01}), low-mass X-ray binaries (see \\cite{Liu+07}), and also the color-magnitude diagrams (e.g. \\cite{Rubenstein+97}). \\citet{Davis+08} also constrain the binary fraction in NGC 6397 by the method of the color-magnitude diagrams, and review the binary fractions in galactic globular clusters. The binaries in the globular cluster play important roles for its dynamical evolution, since the binaries work as energy source through interactions with other stars or binaries (\\cite{Heggie75}). Even if the globular cluster has no binary at initial, the binaries are formed in due course through three-body encounter, and energy generated from these binaries halts core collapse of the globular cluster (\\cite{Henon75}). Furthermore, if the globular cluster contains a nonnegligible fraction of primordial binaries, the dynamical evolution could be affected. \\citet{Goodman+89} (hereafter GH89) first showed theoretically difference of evolutions with and without primordial binary. GH89 estimated that the core of the cluster with primordial binaries at the halt of core contraction is larger than that of the cluster without primordial binaries by order of magnitudes. GH89's estimate is based on the model that the cluster core stop contracting when energy generated by the primordial binaries is balanced with energy outflowing from the inner region of the cluster to the outer region through two-body relaxation. McMillan et al. (1990; 1991) first performed $N$-body simulations of clusters with primordial binaries. They clearly showed that the cluster cores stop contracting at larger cores than those without primordial binaries. By means of Fokker-Planck model, \\citet{Gao+91} investigated post-collapse evolution of cluster with the primordial binary. They showed that the cores of the clusters continues to contract slowly after rapid core contraction and that the gravothermal oscillations occur after the several ten half-mass relaxation time of the slow core contraction. In order to set a population of the primordial binaries in the cluster, several parameters concerning the binary and its distributions, such as mass fraction, and distribution of binding energies and eccentricities, need to be specified. Although these parameters should be derived from theories of star and cluster formations or be limited observationally, sufficient informations have not yet been provided at present. Only a few studies investigated the effect of these parameters on the evolution. \\citet{Heggie+92} performed $N$-body simulations of clusters with two different mass fractions, $6$ and $12$ \\%, of primordial binaries, and \\citet{Heggie+06} investigated the evolutions of clusters with $0-100$ \\% mass fraction of primordial binaries by means of $N$-body simulations. \\citet{Vesperini+94} (hereafter VC94) extended the model of GH89, taking into account the core mass fraction of primordial binaries and the distribution of the binding energies of the binaries. In this paper, we focus on the dependence on binding energies of the primordial binaries. By means of $N$-body simulations, we systematically investigate the dynamical evolution of clusters with different initial hardness of binding energies. We set the distribution of binding energies of primordial binaries by delta function, $\\delta(x-E_{{\\rm bin},0})$, where $E_{{\\rm bin},0}$ is the initial binding energies of the primordial binaries and $E_{{\\rm bin},0}=1,3,10,30,100,300kT_0$. Here, $1.5kT_0$ is the average kinetic energy of stars in the cluster at the initial time. In previous simulations, the distribution of the binding energy of the primordial binaries are usually fixed at uniform distributions in $\\log E_{{\\rm bin},0}$ (see \\cite{McMillan+90}; \\cite{Gao+91}; \\cite{Heggie+92}; \\cite{Heggie+06}). We found that the evolutions of the cores are different according to the distribution of the binding energies. If primordial binaries consist of softer binaries, the binaries are disrupted before core collapse through binary-single and binary-binary encounters, and do not affect the core evolution. If primordial binaries consist of harder binaries, the binaries escape from the cluster. This is because energy generated by the harder binaries per one encounter is larger than the cluster potential. They also do not affect the core evolution and the cluster exhibits deep core collapse. In intermediate range, the binaries efficiently heat the clusters, and the core collapses of the clusters halt halfway. These behaviors are consistent with the theoretical estimate by VC94. These behaviors were not shown by previous works (\\cite{McMillan+90}; \\cite{Gao+91}; \\cite{Heggie+92}; \\cite{Heggie+06}; \\cite{Fregeau+07}; \\cite{Trenti+07}) who followed the dynamical evolution of clusters with primordial binaries. This is mainly because the previous works fixed the distributions of the binding energies of the primordial binaries at uniform distributions in $\\log E_{{\\rm bin},0}$, and their fixed distributions include binaries with hardness in soft, intermediate, and hard ranges. The structure of this paper is as follows. We describe simulation methods in section \\ref{sec:sim}. In section \\ref{sec:results}, we present results of simulations and investigate the effect of the hardness of primordial binaries on the dynamical evolution of clusters. In section \\ref{sec:summary}, we summarize this paper. ", "conclusions": "\\label{sec:summary} We study systematically the dependence of cluster evolution on the binding energy of primordial binaries. By means of GORILLA, we simulate the core evolution of the clusters, each of which contains primordial binaries with equal binding energy. When the initial mass fraction of the primordial binaries is fixed to $0.1$, we find that the dynamical evolutions of the clusters are divided into three ranges according to hardness of the primordial binaries as follows. \\begin{enumerate} \\item In soft range ($<3kT_0$), the clusters experience core collapse in similar way to those without primordial binaries. The ratios of core radii to half-mass radii at the halt of the core collapse are about $0.006$. The primordial binaries do not heat the clusters. This is because the primordial binaries are destroyed through encounters with single stars, and do not generate energy. \\item In intermediate hard range ($10kT_0-100kT_0$), the core collapses in the clusters halt halfway. The ratios of core radii to half-mass radii at the halt of the core collapse are $0.05-0.1$. The primordial binaries release energy, and the energy heats the clusters. \\item In super hard range ($>300kT_0$), the clusters experience core collapse, and the ratios of core radii to half-mass radii at the halt of the core collapse is about $0.02$. The primordial binaries do not so much heat the clusters. Although the primordial binaries release energy through encounters, the energy is so large that binaries and single stars involved with the encounters are ejected from the clusters. \\end{enumerate} The dependences of the boundaries between the soft and intermediate hard ranges and between the intermediate and super hard ranges on the initial mass fraction of the primordial binaries are as follows. \\begin{enumerate} \\item The boundary between the soft and intermediate hard ranges depends on the initial mass fraction of the primordial binaries. When the mass fraction of the primordial binaries is $0.3$, the core contraction in the cluster with $3kT_0$ primordial binaries halts at large ratio of core radius to half-mass radius $\\sim 0.07$, and the intermediate hard range includes $3kT_0$. \\item The boundary between the intermediate and super hard ranges is not changed, when the initial mass fraction of the primordial binaries ranges from $0.03$ to $0.3$. \\end{enumerate} We compared the pairs of the ratios of core radii to half-mass radii and the core mass fraction of the binaries at the halt of the core contraction in our simulations with those of theoretical estimates. We found a good agreement between $N$-body simulations and the theoretical values." }, "1005/1005.4231_arXiv.txt": { "abstract": "We describe in detail the general methodology and numerical implementation of consistent N-body simulations for coupled scalar field cosmological models, including the background cosmology and the generation of initial conditions (with the different couplings to different matter species taken into account). We perform fully consistent simulations for a class of coupled scalar field models with an inverse power-law potential and negative coupling constant, for which the chameleon mechanism does not operate. We find that in such cosmological models the scalar-field potential plays a negligible role except in the background expansion, and the fifth force that is produced is proportional to gravity in magnitude, justifying the use of a rescaled gravitational constant G in some earlier N-body simulations of similar models. We study the effects of the scalar coupling on the nonlinear matter power spectra and compare with linear perturbation calculations to investigate where the nonlinear model deviates from the linear approximation. For the first time, the algorithm to identify gravitationally virialized matter halos is adapted to the scalar field cosmology, and then used to measure the mass function and study the properties of virialized halos. We find that the net effect of the scalar coupling helps produce more heavy halos in our simulation boxes and suppresses the inner (but not the outer) density profile of halos compared with those predicted by lambda-CDM, while this suppression weakens as the coupling between the scalar field and dark matter particles increases in strength. ", "introduction": "\\label{sect:intro} The nature of the dark energy \\cite{Copeland:2006} driving an apparent acceleration of the universe has been a cosmological puzzle for more than a decade. Amongst the models proposed to explain it, those incorporating scalar fields are by far the most popular, not only because of their mathematical simplicity and phenomenological richness, but also because the scalar field is a natural ingredient of many high-energy physics theories. A scalar field contributes a single dynamical degree of freedom which can interact indirectly with other matter species through gravity or couple directly to matter, producing a fifth force on the matter which creates violations of the Weak equivalence principle (WEP). This second possibility was introduced with the hope that such a coupling could potentially alleviate the coincidence problem of dark energy \\cite{Amendola:2000} and has since then attracted much attention (see, for example,\\emph{~}\\cite% {Bean:2001, Amendola:2004, Koivisto:2005, Boehmer:2008, Bean:2008, Bean:2008b, Boehmer:2009} and references therein for some recent work). If there is a direct coupling between the scalar field and baryons, then the baryonic particles will experience a fifth force, which is severely constrained by observations, unless there is some special mechanism suppressing the fifth-force effects. This is the case in chameleon models, where the scalar field (the chameleon) gains mass in high-density regions (where observations and experiments are performed) and the fifth force effects are confined to undetectably small distances \\cite{Khoury:2004a, Khoury:2004b, Mota:2006, Mota:2007}. A common approach which avoids such complications is to assume that the scalar field couples only to the dark matter, an idea seen frequently in models with a coupled dark sector (\\emph{% e.g.}~\\cite{Maartens:2009, Valiviita:2010, Peacock:2010}). In this work our scalar field will not be chameleon-like as this case has been investigated elsewhere \\cite{Li:2009sy, Zhao:2010, Li:2010}. The formation of cosmological structure in the presence of coupled scalar fields at the linear perturbative level has already been studied in great detail. Here, we extend these studies into the nonlinear regime. In particular, we want to know how structure formation on the scales of galaxies and galaxy clusters is modified. As we shall see below, there are principally four effects of the scalar field coupling, namely through (i) the modification of the background expansion rate, (ii) the action of a fifth force on dark-matter particles, (iii) the reduced contribution of dark matter to the Poisson equation, and (iv) the change of the initial condition at early times. Of these four, (ii) can actually be subdivided into (a) an essential \"rescaling\" of gravitational constant and (b) a velocity-dependent acceleration term, while (iv) is a combined consequence of (i, ii, iii) before $z\\sim 50$. We cannot track all these effects into the nonlinear regime using linear analysis and the tool we will employ to obtain quantitative predictions is $N$-body simulation \\cite{Bertschinger:1998}. There has been some earlier work in this area (see for example,\\emph{~}\\cite% {Linder:2003, Mainini:2003, Springel:2007, Kesden:2006, Farrar:2007, Keselman:2009, Maccio:2004, Baldi:2008, Rodriguez-Meza:2007, Rodriguez-Meza:2008, Rodriguez-Meza:2009, Hellwing:2009} and also \\cite% {Oyaizu:2008a, Oyaizu:2008b, Laszlo:2008, Schmidt:2009, Chan:2009, Zhang:2010} for related works). However, in all those scalar-field simulations the scalar field equation is not solved explicitly, but rather it is assumed either that the fifth force is proportional to gravity so that its effect is a simple rescaling of the gravitational constant, or that the fifth force takes the Yukawa form with exponential cut-off beyond some scale. On the other hand, the recent fully consistent simulations performed in \\cite{Li:2009sy, Zhao:2010, Li:2010} have shown that, at least for the chameleon scalar field models, the above simplifying approximations are not good and raises questions about the extent to which we should trust them. It is these concerns that motivates this paper, in which we test the accuracy of those approximations, and study in more detail the qualitative and quantitative effects of a coupled scalar field on the formation and evolution of the nonlinear cosmic structure. As in \\cite{Li:2010, Maccio:2004}, we shall consider a universe which at late times is dominated by a scalar field and two matter species, namely the dark matter, which couples to the scalar field in a way prescribed in Sect.~% \\ref{subsect:basiceqn} and the so-called \"baryons\", which are essentially the dark matter \\emph{without} scalar-field coupling. The paper is organized as follows: in Sect~\\ref{sect:eqn} we list the essential equations to be implemented by our $N$-body simulations and describe briefly their differences from standard $\\Lambda $CDM. Sect.~\\ref% {sect:simu} presents a comprehensive description of the methodology used in this paper and its implementation in the numerical code. Sect.~\\ref% {subsect:code} introduces the code, Sect.~\\ref{subsect:param} lists the physical and simulation parameters we adopt, Sect.~\\ref{subsect:baryons_inc} illustrates how we distinguish between baryons and dark matter particles, Sect.~\\ref{subsect:bkgd} summarizes the results for the background expansion and linear perturbation evolution, which will be referred to subsequently from time to time, and finally Sect.~\\ref{subsect:ic} and Appendix~\\ref% {appen:Zeldovich} explain in detail how to generate initial conditions for the $N$-body simulations which take into account the effects of the coupling between dark matter and the scalar field. As we aim to set up a general framework for $N$-body simulations of coupled scalar field models, we have tried to include all the main ingredients in this section. Our numerical results are presented in Sect.~\\ref{sect:results}, within which Sect.~\\ref% {subsect:snapshots} displays some general results, showing that the approximation of rescaling gravitational constant as used in previous literature is a very good one for the model studied here (but not necessarily so for other models!), and Sects.~\\ref{subsect:Pk}, \\ref% {subsect:mf} and \\ref{subsect:halo} discuss, respectively, how the coupling modifies the nonlinear matter power spectrum, mass function and the internal density profiles of halos. We present our conclusions in \\ref{sect:con}. All through the paper a subscript $_{\\mathrm{B}}$ ($_{\\mathrm{D}}$) denotes a corresponding quantity for baryons (dark matter), unless otherwise stated. ", "conclusions": "\\label{sect:con} Couplings between any scalar field and (some of) the matter species can affect cosmology in various ways. There are two principal effects: modified source terms in the gravitational field equations (\\emph{channel I}) and direct new interactions between particles of the coupled matter species (% \\emph{channel II}). The effects through channel II arise in an inhomogeneous universe, but not in background cosmology. In principle, there is only effect of this sort -- the fifth force, due to the exchanges of scalar quanta between matter particles. In practice, (\\emph{e.g.}, in $N$-body simulations), the scalar field $\\varphi $ is computed in the fundamental observer's frame while the fifth force must be evaluated in individual particle frames; consequently, the fifth force is split into two parts: the part from the spatial gradient of the scalar field $\\delta \\varphi $ in the fundamental observer's frame (which is independent of the particle's peculiar velocity $v$) and the part due to frame transformation which is proportional to $v$. In a homogeneous universe, there is no spatial gradient and all matter particles are comoving with the fundamental observers, so $v=0$, and therefore the channel II effects vanish. In an inhomogeneous universe both terms are nonzero in general, leading to a net force which strengthens the attraction between particles, and enhances their gravitational clustering. The effects arising through channel I appear both in homogeneous (via the Friedmann and Raychaudhuri equations) and inhomogeneous (via the Poisson equation) universes. They arise mainly via a renormalization of the contributions to the total density from the coupled matter species and by a new contribution from the scalar field itself. \\ The exact effects depend on the specific forms of the scalar potential and coupling function. Here, we list our main findings for an inverse power-law potential Eq.~(\\ref% {eq:potential}), with an exponential coupling Eq.~(\\ref{eq:coupling_function}% ), and the model parameters given in Sect.~\\ref{subsect:param}. \\footnotetext{% Note in Fig.~\\ref{fig:background} (lower left panel) the expansion rate of the coupled scalar field model finally catches up with that of $\\Lambda $CDM at $a=1$. This is because at late times the dark energy (scalar field) density is higher than that in $\\Lambda $CDM.} For the background cosmology, the contribution to the Friedmann equation from dark matter is renormalized by $C(\\varphi )<1$, and $C(\\varphi )$ decreases as the coupling constant $|\\gamma |$ increases. As a result, a stronger coupling (larger $|\\gamma |$) produces lower cosmic expansion rate during the matter-dominated era (Fig.~\\ref{fig:background})\\footnotemark[5] , which will in turn helps enhance the clustering of matter and promotes structure formation. Such an enhancement is just what we have observed from our linear perturbation analysis (Figs.~\\ref{fig:linpert} and \\ref{fig:linpert2}): on very large scales, which are beyond the range of the fifth force effects, the matter power spectrum increases with $|\\gamma |$. On small scales, the fifth force helps to enhance the matter power, making it increase further than predicted in $\\Lambda $CDM (comparing the left to the right panels of Fig.~\\ref{fig:linpert2}, where we have separated baryons, which have no scalar coupling, from dark matter, which has). This enhancement starts at a rather early, and so the scalar field does leave imprints on the matter power spectrum at $z\\sim 50$, which means that the initial condition for $N$% -body simulations also needs to modified. For example, in the model with $% | \\gamma |=0.15$, we have found that : (i) The density perturbation at $z\\sim 49$ is about $10\\%$ larger than the corresponding result in $\\Lambda $CDM for baryons (due to slower cosmic expansion) and about $12\\%$ larger for dark matter (due to the slower expansion \\emph{and} the fifth force effects); (ii) The average velocity at the same time is about $8\\%$ larger than the corresponding result in $\\Lambda $CDM for baryons and about $10\\%$ larger for dark matter. We introduce a quick method to take these changes into account when generating initial conditions for $N$-body simulations. One of the most important results from our $N$-body simulations is the magnitude of fifth force. Fig.~\\ref{fig:force} shows that everywhere and at any time (\\emph{i.e.}, in both high and low density regions) the fifth force (or more precisely, the velocity-\\emph{in}dependent part of it) is proportional to gravity in magnitude (with a coefficient $2\\gamma ^{2}$ when only considering dark matter). This is the approximation used in many previous simplified simulations, and our results confirm numerically that this approximation works fairly well, at least for those cases where the scalar potential is unimportant, such as ours. We emphasize, however, that the velocity-dependent part of the fifth force should \\emph{not} be dropped (as in some previous simulations) unless there is good reason to do so (see \\cite{Li:2010}). For the nonlinear matter power spectrum (Fig.~\\ref{fig:power}), we find that the $N$-body simulations show agreement with linear perturbation analysis on large scales. On intermediate scales, our simulations predict significant enhancement in the matter power over that predicted both by linear perturbation analysis and by the $\\Lambda $CDM paradigm. This enhancement is strongest at early epochs but gradually weakens at late times. One possible reason for this is that the scalar coupling, $C(\\varphi ),$ decreases with time, and reduces the contribution of dark matter to the gravitational potential, so weakening further clustering, but this still needs to be investigated in more detail. The bias between dark matter and baryons increases with time and with coupling strength $|\\gamma |$. This is because larger $|\\gamma |$ implies stronger fifth forces and therefore stronger total forces act on dark matter particles compared to baryons. As time passes, the bias has more time to develop and so increases as well (Fig.~\\ref{fig:powerbc}). Using the same reasoning, the fifth force increases the attraction of dark matter particles and so heavier halos are expected to form during the structure formation (Fig.~\\ref{fig:mf}). In order to identify gravitationally bound and virialized halos in our models we must also take into account this scalar coupling (Sect.~\\ref{subsect:halo}). The scalar field coupling also has impacts on the internal density profiles of the halos (Figs.~\\ref{fig:profile} and \\ref{fig:profiler}). We average ten of the heaviest halos and find that (i) When compared with $\\Lambda $CDM, the overdensities in the inner regions can be lower while those in the outer regions is higher, showing the failure of the halos to retain particles in the inner region, either because the particles move too fast or because the attractive potential at the centre is too weak; (ii) The suppression of inner density compared with $\\Lambda $CDM weakens, rather than strengthens, as $|\\gamma |$ increases. Indeed, for $|\\gamma | =0.2 $ the density is higher than $\\Lambda $CDM result throughout the halos. More detailed studies are needed to clarify the leading effect responsible for this observed pattern. (iii) There is a bias between density profiles for baryons and dark matter (Fig.~\\ref{fig:profilebc}): the dark matter density is always higher because it experiences the fifth force which boosts its clustering. In summary, in this paper we have been given a comprehensive description of the methodology and implementation of general $N$-body simulations for coupled scalar field models. Some important issues in these simulations have been addressed here for the first time: the consistent solution of the scalar field and fifth force, the fifth force effects on generating initial conditions, and the effects of scalar field on identifying virialized halos. Although the situation is complex, we have identified interesting new features in these models. We hope these developments will lead to more detailed study of nonlinear structure formation in this class of models and facilitate their subsequent confrontation with observational data." }, "1005/1005.3885_arXiv.txt": { "abstract": "We present a CCD photometric study of the star with ASAS ID 134738$+$0410.1 using V band observations obtained from the $IUCAA$ Girawali Observatory (IGO) 2-metre telescope, India.~The star was selected from the $\\delta$ Scuti database of All Sky Automated Survey (ASAS) (Pojmanski 2002). Our analysis reveals that the star is not a $\\delta$ Scuti variable but is in fact a W UMa type contact binary with an orbital period of 0.2853067 day. Two new times of primary and secondary minima were determined from the observed data. A preliminary solution obtained using the Wilson-Devinney light curve modelling technique indicates that the star is more likely a partially-eclipsing W UMa type contact binary. However, the determination of actual subtype of this binary is quite impossible from the photometry alone, as the observed light curve can fitted for both A- and W-type solutions. The exact classification of this binary needs to be determined from high resolution spectroscopy. ", "introduction": "\\label{sec:intro} In the ASAS\\footnote{http://www.astrouw.edu.pl/asas/} database, the star has maximum V magnitude of 13.47 mag and is located at $\\alpha = 13^{\\rm h}\\, 47^{\\rm m}\\,38^{\\rm s}.0$ and $\\delta = 04^{\\rm o}\\,10^{\\rm '}\\,05^{\\rm ''}.9$ (Epoch J2000). The star has been classified as $\\delta$ Scuti/Contact/ Semi-Detached binary with a period of P$_{\\rm ASAS}$ = 0.142654 day. The classification is thus ambiguous. In order to know true nature of this object, we have obtained high precision CCD photometric data from new observations. The new data indicates that the star is a W UMa contact binary with an orbital period of P$_{\\rm IGO} = 0.2853067$ day which is nearly twice the period P$_{\\rm ASAS}$ = 0.142654 day. \\begin{table*} \\begin{center} \\caption{Basic information of ASAS 134738$+$0410.1, the comparison star and the check star} \\label{Table1} \\scalebox{0.82}{ \\begin{tabular}{lccccc} \\hline Star & RA (J2000) & DEC (J2000) & J [mag] & H [mag]& K [mag] \\\\ \\hline ASAS 134738$+$0410.1 (Target) & $13^{\\rm h}\\,47^{\\rm m}\\,38^{\\rm s}.00$ & $04^{\\rm o}\\,10^{\\rm '}\\, 06^{\\rm ''}.00$ &11.894 &11.472 &11.407 \\\\ 2MASS 13473475$+$0408167 (Comparison) & $13^{\\rm h}\\,47^{\\rm m}\\,34^{\\rm s}.76$ & $04^{\\rm o}\\,08^{\\rm '} 16^{\\rm ''}.72$ &12.408&12.026 &11.951\\\\ 2MASS 13472306$+$0413420 (Check) & $13^{\\rm h}\\,47^{\\rm m}\\,23^{\\rm s}.06$ & $04^{\\rm o}\\,13^{\\rm '}\\,42^{\\rm ''}.10$ &12.213 &11.870 &11.789 \\\\ \\hline \\end{tabular}} \\end{center} \\end{table*} The V band CCD photometric observations of the star were carried out with the IGO 2-m telescope, located about 80 km from Pune, India during two nights on March 31 and April $02$ in 2009.~The IUCAA Faint Object Spectrograph Camera (IFOSC) equipped with EEV 2 K$\\times$ 2 K thinned, back-illuminated CCD with 13.5 $\\mu m$ pixels was used. The CCD used for imaging provides an effective field of view of $\\sim 10.5^{'} \\times 10.5^{'} $ on the sky corresponding to a plate scale of 0.3 arcsec pixel$^{-1}$. The gain and read out noise of the CCD camera are 1.5 e$^{-}$/ADU and 4 e$^{-}$ respectively. The FWHM of the stellar image varied from 3 to 5 pixels during the observations. We took a total of 153 frames in the V band with the exposure times varied between 100 s and 180 s for a good photometric accuracy. The co-ordinates of the variable, comparison star and the check star along with the infrared JHK magnitudes taken from the 2MASS catalogue (Cutri et al. 2003) are listed in Table 1. The comparison and the check star are so close to the variable that they are in the same field during the observations. Image pre-processing and data reduction was carried out using IRAF\\footnote {IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation.} and MIDAS softwares. Instrumental magnitudes were obtained using the DAOPHOT package (Stetson 1987, 1992). The various tasks, e.g., $find, phot, daogrow, daomatch$ and $daomaster$ were applied in order to obtain the instrumental magnitudes of stars in all the frames. Extinction corrections were ignored as the target star is very close to the comparison star. In Fig.~1, we show the plots of the differential V band magnitude of (Variable - Comparison), (Comparison - Check) versus Heliocentric Julian Day (HJD) in the left, right upper and lower panels respectively for observations on March 31 and April 02, 2009. The reduced results show that the difference between the magnitude of the check star and that of the comparison star was constant with a probable error of $\\pm$0.004 mag in the V band \\footnote{The observational data presented in this paper can be obtained from the authors on request.}. ", "conclusions": "We have discovered that star having ASAS ID 134738+0410.1 identified by ASAS is a W UMa type overcontact binary. We have determined new period and times of minima from very accurate and precise CCD data. Based on the WD code as implemented in the software PHOEBE, we have done preliminary modelling of the light curve. However, unique value of the mass ratio could not be obtained from the modelling, due to its partial-eclipsing nature. Hence, the actual subtype of the system could not be determined. This star deserves high resolution spectroscopic time series radial velocity measurements to determine its true subtype and mass ratio which should be combined with the photometric light curve data to obtain various geometrical and physical parameters accurately." }, "1005/1005.5147_arXiv.txt": { "abstract": "{We present results from the earliest observations of DEBRIS, a {\\it Herschel} Key Programme to conduct a volume- and flux-limited survey for debris discs in A-type through M-type stars. PACS images (from chop/nod or scan-mode observations) at 100 and 160 \\micron\\ are presented toward two A-type stars and one F-type star: $\\beta$ Leo, $\\beta$ UMa and $\\eta$ Corvi. All three stars are known disc hosts. {\\it Herschel} spatially resolves the dust emission around all three stars (marginally, in the case of $\\beta$ UMa), providing new information about discs as close as 11 pc with sizes comparable to that of the Solar System. We have combined these data with existing flux density measurements of the discs to refine the SEDs and derive estimates of the fractional luminosities, temperatures and radii of the discs. } ", "introduction": "\\label{intro} Debris discs are flattened distributions of planetesimals and dust located at radii of 1-1000~AU around main-sequence stars (see \\citep[see][for a recent review]{wyatt08}. The dust cannot be primordial since its lifetime in orbit is significantly less than the age of the host stars. Instead, dust is replenished from a population of colliding km-sized planetesimals \\citep{wyatt02,the07}. Over time, the dust distribution is shaped by any planetary-sized bodies in the system \\citep[e.g.,][]{dom03,wyatt07}. Therefore, resolved images of discs constrain models of the structure and evolution of planetary systems. Far-infrared and submillimetre observations are the best way to search for dust around nearby stars due to the favorable contrast of the disc relative to the star. At these wavelengths, the disc emission is optically thin and is sensitive to the large (up to $\\sim 1$ mm), cold grains which dominate the disc's dust mass. The {\\em Herschel} Space Observatory offers three major advantages for the detection and characterization of debris discs: far-infrared sensitivity, angular resolution and wavelength coverage. With its 3.5 m mirror, its sensitivity at far-infrared wavelengths is superior to any previous instrument. With its resolution of 6\\farcs7 at 100 \\micron, {\\it Herschel} has the potential to resolve many debris discs, particularly toward nearby stars. Finally, with detectors at 100, 160, 250, 350 and 500 \\micron, {\\it Herschel} has the means to sample the spectral energy distribution (SED) of disc emission across the peak, meaning models can be better constrained even for discs which are not resolved. DEBRIS (Disc Emission via a Bias-free Reconnaissance in the Infrared/Submillimetre) is an Open Time Key Programme which uses PACS (Photodetector Array Camera \\& Spectrometer) and (for appropriate targets) SPIRE (Spectral and Photometric Imaging Receiver) to detect, resolve and characterize debris discs around a volume-limited sample of 446 A through M type stars. The goals of DEBRIS include establishing the incidence and evolution of debris discs as a function of stellar type, age, multiplicity, etc.; the characterization of discs in terms of size, temperature, dust mass and morphology (where the disc asymmetries could indicate the presence of planetary companions); and the understanding of our own Solar System in the context of the larger debris disc population. Full details of the DEBRIS survey and goals will be presented in a forthcoming paper (B.~Matthews et al. 2010, in preparation). Here, we present PACS observations toward three of the first targets of the DEBRIS survey. We briefly summarize the observations and targets in Sect.~\\ref{obs}, present the results in Sect.~\\ref{res} and discuss three sources in detail in Sect.~\\ref{disc}. We summarize the paper in Sect.~\\ref{sum}. \\begin{figure*} \\centering \\includegraphics[bb=20 335 640 645,width=20cm,clip]{matthewsb_fg1.ps} \\caption{Images of the 100 and 160 \\micron\\ emission from three DEBRIS targets: $\\beta$ Leo, $\\beta$ UMa and $\\eta$ Corvi. Contours are shown at 0, 10, 30, 50, 60, 70, 80, 90 and 99\\% of the peak in each map. The 1-$\\sigma$ rms noise levels are given in Table \\ref{table}. Circles in the upper left corner of each panel mark the nominal beam sizes for a scan speed of 20\\arcsec/s, i.e., 6\\farcs7 and 11\\arcsec\\ at 100 and 160 \\micron, respectively. The negative images created by the chop/nod observing mode are visible in the $\\beta$ UMa images. Striping in the $\\eta$ Corvi image at 160 \\micron\\ is due to high 1/f noise and filtering artefacts. } \\label{images} \\end{figure*} ", "conclusions": "\\label{disc} \\subsection{$\\beta$ Leo} \\label{betaleo} Figure \\ref{images} shows the first resolved images of the disc around $\\beta$ Leo. The blackbody temperature and dust luminosity results given in Table \\ref{table} are consistent with the values found in previous works \\citep{su06,holmes03}. The fractional dust luminosity of $2.3 \\times 10^{-5}$ is 15\\% higher than the estimate from \\cite{su06}. The increase is due to a slight increase in dust to match the PACS flux densities. The radius estimates found for $\\beta$ Leo are comparable to that of the Kuiper Belt ($\\sim 50$ AU). This makes the $\\beta$ Leo disc one of the smallest disc radii yet resolved at any wavelength (see, for instance, the 'Circumstellar Disks Database'\\footnote{cirumstellardisks.org}) although smaller characteristic orbital radii have been derived based on single temperature blackbody fits to the dust components \\citep[e.g.,][]{rhee07}. The difference between $R_{obs}$ and $R_{dust}$ provides an opportunity to learn about the grains within this disc. Because differently sized grains can have the same temperature at different distances from a star, the SED models in Fig.\\ \\ref{SEDs} are degenerate. This degeneracy is broken by the resolved imaging. For example, the $\\sim$40 AU radius for the $\\beta$ Leo disc is larger than the 23 AU suggested by the blackbody fit. Therefore, the grains do not emit as blackbodies, but maintain a $\\sim$112 K temperature at a greater distance from the star as expected for small grains that emit inefficiently at far-IR wavelengths. The inferred characteristic particle radius $a$ is $ < \\lambda/2 \\pi = 16$ \\micron. Future modeling work that combines {\\it Spitzer} IRS spectra and submillimetre images with the {\\it Herschel} data will constrain these grain properties and the spatial dust distribution (L.~Churcher et al. 2010, in preparation). \\subsection{$\\beta$ UMa} \\label{betauma} Figure \\ref{images} shows that the disc emission around $\\beta$ UMa is very compact at 100 and 160 \\micron. The disc is marginally resolved at 100 \\micron\\ and not resolved at 160 \\micron. The apparent asymmetry in the 160 \\micron\\ disc image is likely artificial; it is an effect of interpolation applied to the image at native (Nyquist sampled) resolution. The flux densities measured for $\\beta$ UMa confirm the earlier 100 and 160 \\micron\\ detections. The disc component of the $\\beta$ UMa SED is well fit by blackbody grains with a temperature comparable to that of $\\beta$ Leo, requiring a bigger disc around the more luminous star. Therefore, assuming black body grains the radial estimate is 51AU, equivalent to the deconvolved disc radius from the 2D Gaussian fit to the 100 \\micron\\ image. The resolved size thus suggests an absence of small grains such as that inferred for beta Leo in Sect.~\\ref{betaleo}. More detailed modeling will be forthcoming in a future paper. \\subsection{$\\eta$ Corvi} The new {\\it Herschel} images in Fig.\\ \\ref{images} show that $\\eta$ Corvi is resolved at both 100 and 160 \\micron, as expected based on the $\\sim 300$ AU submillimeter size derived by \\cite{wyatt05}. The variation in morphology from centrally peaked emission at 100 \\micron\\ to a double-peaked limb brightened ring at 160 \\micron\\ (as observed at 450 \\micron) is consistent with an outer cool ring filled in by warmer dust which dominates the emission at 100 \\micron. This could be evidence of the third temperature component proposed by \\cite{chen06} and observed in $\\epsilon$ Eri by \\cite{back09}, although this was tentatively ruled out in mid-IR imaging by \\cite{smith08}, and more generally suggests the radial distribution of material is broader than the two ring system originally envisaged by \\cite{wyatt05}. The $R_{obs}$ estimate from Table \\ref{table} is equivalent to the submillimetre size. The two intensity maxima in the 160 \\micron\\ image are roughly a beamwidth (11\\arcsec) apart, identical to the 450 \\micron\\ SCUBA imaging of \\cite{wyatt05} who inferred that the emission arises from a ring at moderate inclination. Fitting a 2D Gaussian to the 100 \\micron\\ image of Fig.\\ \\ref{images} gives a position angle of 102\\degr $\\pm 7$\\degr and an inclination of $\\sim$ 50\\degr\\ from the line of sight. The position angle of the two peaks at 160 \\micron\\ is $\\sim 135$\\degr. These inclinations are consistent with the 450 \\micron\\ measurement of $130$\\degr\\ $\\pm 10$\\degr. A two component model of the SED of Fig.\\ \\ref{SEDs} shows similar results to \\cite{wyatt05} who found disc components of 40 K and 370 K. The warm component \\citep{smith09,smith08,chen06} shown in Fig.\\ \\ref{SEDs} has a blackbody temperature of 346 K, corresponding to a radial distance of 1.4 AU from the star. The cold component has a temperature of 33 K, corresponding to a radial separation of 160 AU from the star, consistent with $R_{obs}$. As for $\\beta$ UMa, the resolved size suggests an absence of the small grains implied for $\\beta$ Leo in Sect.~\\ref{betaleo}. Most importantly, the images in Fig.\\ \\ref{images} provide an estimate of the disk size at wavelengths intermediate between the submillimetre (which shows emission at $\\sim 150$ AU) and mid-IR (which shows emission at $<$ 3.5 AU). This will be crucial for modeling the origin of the far-infrared morphology, which most resembles the submillimetre emission. Simultaneously modeling these several images will constrain in more detail the dust properties of the disc system." }, "1005/1005.2365.txt": { "abstract": "In the past few years, several disks with inner holes that are relatively empty of small dust grains have been detected and are known as transitional disks. Recently, {\\it Spitzer} has identified a new class of ``pre-transitional disks\" with gaps based on near-infrared photometry and mid-infrared spectra; these objects have an optically thick inner disk separated from an optically thick outer disk by an optically thin disk gap. A near-infrared spectrum provided the first confirmation of a gap in the pre-transitional disk of LkCa~15 by verifying that the near-infrared excess emission in this object was due to an optically thick inner disk. Here we investigate the difference between the nature of the inner regions of transitional and pre-transitional disks using the same veiling-based technique to extract the near-infrared excess emission above the stellar photosphere. However, in this work we use detailed disk models to fit the excess continua as opposed to the simple blackbody fits used previously. We show that the near-infrared excess emission of the previously identified pre-transitional disks of LkCa~15 and UX~Tau~A in the Taurus cloud as well as the newly identified pre-transitional disk of Rox~44 in Ophiuchus can be fit with an inner disk wall located at the dust destruction radius. We also present detailed modeling of the broad-band spectral energy distributions of these objects, taking into account the effect of shadowing by the inner disk on the outer disk, but considering the finite size of the star, unlike other recent treatments. The near-infrared excess continua of these three pre-transitional disks, which can be explained by optically thick inner disks, are significantly different from that of the transitional disks of GM~Aur, whose near-infrared excess continuum can be reproduced by emission from sub-micron-sized optically thin dust, and DM~Tau, whose near-infrared spectrum is consistent with a disk hole that is relatively free of small dust. The structure of pre-transitional disks may be a sign of young planets forming in these disks and future studies of pre-transitional disks will provide constraints to aid in theoretical modeling of planet formation. ", "introduction": "\\label{intro} Several disks which have nearly photospheric near-infrared emission but substantial excesses above the stellar photosphere at wavelengths beyond $\\sim$20~{\\micron} have been observed and are referred to as ``transitional disks'' \\citep{strom89}. Using data from the {\\it Spitzer} Infrared Spectrograph \\citep[IRS;][]{houck04}, detailed modeling has demonstrated that this flux deficit at near-infrared wavelengths relative to full disks can be explained by optically thick disks with inner holes of $< 40$~AU. In most cases these inner holes are not completely devoid of material (e.g. GM~Aur, TW Hya, CS Cha, CVSO 224); a minute amount of micron- or sub-micron-sized optically thin dust exists within the hole, producing a small infrared excess over the photospheric flux, still well below the median excess of Class II objects, as well as silicate emission \\citep{calvet02,calvet05, espaillat07a, espaillat08b}. Gas has also been detected within the inner holes of transitional disks \\citep[e.g.][]{najita03,bergin04,salyk07}. Recently, the {\\it Spitzer Space Telescope} \\citep{werner04} identified a new class of disks called ``pre-transitional disks\" around LkCa~15 and UX~Tau~A \\citep{espaillat07b}. These disks have deficits of mid-infrared flux (5--20~{\\micron}) and substantial excesses at longer wavelengths, as is seen in the transitional disks. However, in contrast to the small or absent near-infrared (2--5~{\\micron}) excesses exhibited by transitional disks, pre-transitional disks have significant near-infrared excesses relative to their stellar photospheres, similar to the median spectral energy distribution (SED) of disks in Taurus \\citep{dalessio99}. The distinctive shapes of these SEDs indicate that pre-transitional disks have an inner disk separated from an outer disk and that we may be seeing the development of gaps within protoplanetary disks. While the truncation of LkCa~15's outer disk has been imaged in the millimeter \\citep{pietu06}, \\citet{espaillat07b} showed that the substantial near-infrared excess of LkCa~15 could be explained by either optically thick material or by ${\\sim}10^{-11}$ M$_{\\odot}$ of optically thin dust mixed with the gas in the inner disk. In order to resolve this issue, \\citet{espaillat08a} obtained a medium resolution near-infrared spectrum spanning the wavelength range 2--5~{\\micron}. This near-infrared spectrum had absorption lines that were weaker relative to the spectrum of a standard star of the same spectral type. This phenomenon, known as ``veiling\" \\citep{hartigan89}, is also observed in similar spectra of full disks and is due to emission from dust located at the dust sublimation radius \\citep{muzerolle03}. \\citet{espaillat08a} measured a veiling factor (r$_{K}$) of 0.3$\\pm$0.2 for LkCa~15 at $\\sim$2.2~{\\micron} and fit the near-infrared excess from 2--5~{\\micron} with a single-temperature blackbody of 1600~K. This behavior can be explained by an optically thick wall located at the dust sublimation radius, as is seen in full disks \\citep{muzerolle03}. These data confirmed that LkCa~15 has an inner optically thick disk, making this observation the first independent verification of a gap in a protoplanetary disk. Here we expand our sample to include the pre-transitional disks of UX Tau A \\citep{espaillat07b} and Rox~44 \\citep{furlan09}, and the transitional disks of GM~Aur and DM~Tau \\citep{calvet05} in order to explore the structure of the inner regions of pre-transitional and transitional disks. To do this we use veiling measurements in the K-band to extract the near-infrared (-IR) excess emission of these objects and then fit this emission with disk models \\citep{dalessio05,calvet02}. Given the sensitivity of veiling measurements on the adopted spectral type, we redetermined spectral types for our targets using optical spectra ({\\S}~\\ref{sec:spt}). We also tested our veiling measurement methods with the spectrum of the diskless, weak-line T Tauri star (WTTS) LkCa~14 (\\S~\\ref{sec:lkca14}). We find that the near-infrared spectra of the pre-transitional disks of LkCa~15, UX~Tau~A, and Rox~44 are well-explained by the wall of an optically thick inner disk ({\\S}~\\ref{sec:lkca14}, {\\S}~\\ref{sec:ux}, {\\S}~\\ref{sec:rox}). In contrast, our data shows that the inner hole of the transitional disk of GM~Aur contains a small amount of optically thin sub-micron-sized dust while DM~Tau's hole is relatively free of small dust ({\\S}~\\ref{sec:gm} \\& {\\S}~\\ref{sec:dm}). Our results are consistent with veiling and interferometric measurements found in the literature. We also perform detailed model fits to the broad-band SEDs of our pre-transitional disk sample, and explore the effect of shadowing of the outer disk by the inner disk ({\\S}~\\ref{sec:models}), taking into account that the star is not a point source, as has been adopted in other studies \\citep{espaillat07b, mulders10}. The structure of pre-transitional disks suggests that of the disk clearing mechanisms proposed to date, planet formation \\citep[e.g.][]{goldreich80, rice03, varniere06} is most likely a dominant factor in clearing these disks. ", "conclusions": "\\label{sed:sum} In this paper we presented spectral type measurements as well as veiling and near-infrared excess measurements for the diskless LkCa~14, the pre-transitional disks of LkCa~15, UX~Tau~A, and Rox~44, and the transitional disks of DM~Tau and GM~Aur. Using near-IR SpeX spectra from 1--5 {\\micron}, we extracted the near-IR excess continua of our pre-transitional disk sample and fit their excess emission with optically thick inner disk walls, supporting the interpretation that these objects contain gaps in their disks. We modeled the broad-band SEDs of these pre-transitional disks and measure gap sizes of $\\sim$40--70~AU. In the case of LkCa~15 and Rox~44 there is $\\sim$10$^{-11}$~{\\msun} of ISM-sized optically thin dust within the gap. The near-infrared emission of our pre-transitional disks differs significantly from that of the transitional disks of DM~Tau and GM~Aur. DM~Tau has no veiling and no excess emission in the near-IR, indicating that its disk hole is relatively empty of small dust grains. GM~Aur has a small amount of veiling and displays a flat excess above the stellar photosphere in the near-IR, which can be fit by a model of optically thin dust emission. This is in contrast to the pre-transitional disks which have large, blackbody-like near-IR excess continua that can be fit with models of a wall located at the dust sublimation radius. We also studied the effects of shadowing of the outer disk wall by the inner disk. We found that when the finite size of the star is taken into account, a significant portion of the outer wall is either in the prenumbra or completely out of the shadow of the inner wall. The predicted height of the wall is consistent with that of the outer disk, which is derived independently. Based on currently known disk clearing mechanisms, we propose that the gaps in pre-transitional disks are indicators of planet formation, making these disks a promising location for young planet searches. {\\it ALMA} will be pivotal in extending the sample of known pre-transitional disks and has the potential to detect planets in these gaps. Near-IR interferometry will play an important role in further understanding the innermost regions of these disks. Future studies of this class of objects may bring us a few steps closer to understanding the origin of our own solar system." }, "1005/1005.5001_arXiv.txt": { "abstract": "s{ Excitation of multicomponent dark matter in the galactic center has been proposed as the source of low-energy positrons that produce the excess 511 keV $\\gamma$ rays that have been observed by INTEGRAL. Such models have also been promoted to explain excess high-energy $e^\\pm$ observed by the PAMELA, Fermi/LAT and H.E.S.S.\\ experiments. We investigate whether one model can simultaneously fit all three anomalies, in addition to further constraints from inverse Compton scattering by the high-energy leptons. We find models that fit both the 511 keV and PAMELA excesses at dark matter masses $M < 400$ GeV, but not the Fermi lepton excess. The conflict arises because a more cuspy DM halo profile is needed to match the observed 511 keV signal than is compatible with inverse Compton constraints at larger DM masses. } ", "introduction": " ", "conclusions": "We have found that annihilating multistate DM can explain two out of three galactic cosmic ray anomalies, either PAMELA/Fermi or PAMELA/INTEGRAL, but not all three simultaneously. Although it is possible to marginally predict all the correct rates using Einasto profile parameter $\\alpha=0.20$, the angular distribution of 511 keV $\\gamma$ rays is too wide in this case. Of the two possibilities, the PAMELA/INTEGRAL combination seems preferable from the standpoint of the required DM halo parameters, since in this case we are able to adopt standard values that are quite compatible with $N$-body simulations of galactic structure evolution. Moreover we can match the anomalous lepton rates well for PAMELA/INTEGRAL. The PAMELA/Fermi possibility requires stretching the halo parameters to their maximal values, while only marginally giving a large enough rate of leptons, yet a small enough rate of associated inverse Compton $\\gamma$ rays." }, "1005/1005.2621_arXiv.txt": { "abstract": "We present the discovery of an extremely bright and extended lensed source from the second Red Sequence Cluster Survey (RCS2). RCSGA 032727-132609 is spectroscopically confirmed as a giant arc and counter-image of a background galaxy at $z=1.701$, strongly-lensed by the foreground galaxy cluster RCS2 032727-132623 at $z=0.564$. The giant arc extends over $\\sim 38$\\,\\arcsec and has an integrated $r$-band magnitude of 19.1, making it $\\sim 20$ times larger and $\\sim 3.5$ times brighter than the prototypical lensed galaxy MS1512-cB58. This is the brightest distant lensed galaxy in the Universe known to date. % We have collected photometry in 9 bands, ranging from $u$ to $K_s$, which densely sample the rest-frame UV and optical light, including the age-sensitive 4000\\AA\\ break. A lens model is constructed for the system and results in a robust total magnification of $2.04 \\pm 0.16$ for the counter-image; we estimate an average magnification of $17.2 \\pm 1.4$ for the giant arc based on the relative physical scales of the arc and counter-image on the sky. Fits of single-component spectral energy distribution (SED) models to the photometry result in a moderately young age, $t=80\\pm40$\\,Myr, small amounts of dust, $E(B-V) \\le 0.11$, and an exponentially declining star formation history with \\textit{e}-folding time $\\tau = 10-50$\\,Myr. After correcting for the lensing magnification, we find a stellar mass of M$_* \\sim 10^{10}$\\,M$_\\odot$ and a current star formation rate SFR$\\le77$\\,M$_\\odot$\\,yr$^{-1}$. Allowing for episodic star formation, an underlying old burst could contain up to twice the mass inferred from single-component modeling. % RCSGA 032727-132609 is typical of the known population of star-forming galaxies near this redshift in terms of its age and stellar mass. Its large magnification and spatial extent provide a unique opportunity to study the physical properties of an individual high-redshift star-forming galaxy in great detail, opening up a new window to the process of galaxy evolution between $z=1.7$ and our local Universe. \\subjectheadings{galaxies: evolution, galaxies: high-redshift, gravitational lensing} ", "introduction": "Significant progress has been made in recent years towards the study of the formation of galaxies and their evolution into the population of objects we observe around us today. The redshift range $1.0 \\lesssim z \\lesssim 3.0$ represents a crucial period in this process since the peak of star formation in the Universe is known to occur near $z \\sim 2$ \\citep{Blain:99, Chapman:03, Reddy:09}. Through the development of pre-selection color criteria \\citep{Adelberger:04, Steidel:04, Daddi:04a} and advances in the near-IR and UV spectroscopic capabilities of 8-10\\,m class telescopes, growing samples of optically or near-IR selected galaxies are now spectroscopically confirmed in this redshift range and studied extensively (see e.g. Shapley et al. 2005; Erb et al. 2006; Kriek et al. 2008). For a complete understanding of the process of galaxy formation, it is crucial to complement these statistical results based on larger samples with detailed study of the stellar populations and dynamics of individual objects. The majority of galaxies in the current samples are too faint for this purpose. The few galaxies that are bright enough to be studied individually represent outliers drawn from the extreme bright tail of the luminosity function, and are therefore not necessarily representative of the bulk of the population. The ability to study the properties of faint, high-redshift galaxies is one of the main science drivers for the construction of 30\\,m class telescopes. Until such instruments become available we can make a head start using gravitational lensing to increase the power of the current generation of telescopes. The lensing magnification induced by foreground galaxy clusters and individual galaxies has been succesfully used to identify galaxies out to $z \\sim 10$, opening up new windows into the very distant Universe \\citep{Richard:08, Bouwens:09}. At more moderate redshifts it brings individual galaxies from the samples at $1.0 \\lesssim z \\lesssim 3.0$ to a flux level amenable for extensive follow-up observations at various wavelengths. The first and most notable example in this class is MS1512-cB58, a Lyman Break Galaxy (LBG) at $z = 2.73$ \\citep{Yee:96, Ellingson:96}, found to have a lensing magnification of $\\sim$ 30 \\citep{Williams:96, Seitz:98}. It has been studied extensively since its discovery and provides a wealth of information on the stellar population and dynamics of a young star-forming galaxy at this redshift \\citep{Pettini:00, Pettini:02, Siana:08}. Other examples of particularly bright lensed galaxies at similar redshifts include the `Cosmic Eye' at $z=3.07$ \\citep{Smail:07,Coppin:07}, the `8 o'clock arc' at $z=2.73$ \\citep{Allam:07, Finkelstein:09} and two strongly-lensed $z\\sim3$ LBGs from the SDSS Giant Arcs Survey \\citep{Koester:10}. These highly magnified sources currently represent the best places to study the individual properties of high-redshift star-forming galaxies in great detail. % In this paper we present the discovery of a very bright and extended galaxy, RCSGA 032727-132609, spectroscopically confirmed at $z = 1.701$ and highly magnified by a foreground cluster at $z = 0.564$ from the second Red Sequence Cluster Survey (RCS2; D. Gilbank et al. 2010, in preparation). The RCS surveys were designed for the purpose of cluster finding via the identification of the linear color-magnitude relation present for early-type galaxies in clusters, known as the red-sequence technique \\citep{Gladdersyee:00}. The RCS2 survey has imaged $\\sim700$ square degrees of sky in $g$,$r$ and $z$ with the MegaCam camera at the Canada-France-Hawaii Telescope (CFHT) on Mauna Kea in 4, 8 and 6\\,min exposures respectively. Data acquisition for the survey finished in 2008 and preliminary cluster catalogs over the entire area have been created and visually inspected for strong lensing signatures; the details of this search will be published elsewhere. The brightest and most obvious strong lensing system found in this search is RCSGA (RCS Giant Arc) 032727-132609. The system consists of a counter-image and a giant arc extending over $\\sim 38$\\,\\arcsec at an Einstein-radius of $\\sim 17.8$\\,\\arcsec, estimated from the distance between the brightest knot in the arc and the brightest cluster galaxy. The arc has an apparent magnitude of $r = 19.1$, making it $\\sim 3.5$ times brighter than cB58 and the brightest distant lensed galaxy in the Universe known to date. Its large spatial extent provides unique opportunities to look inside a high-redshift galaxy and spatially resolve its substructure. \\npar The paper is organized as follows. \\S~\\ref{sec:data} presents the multitude of photometric and spectroscopic data we have assembled on this system: broadband observations in 9 bands ranging from $u$ to $K_s$, a medium-resolution optical spectrum of the lensed galaxy to obtain the source redshift and redshift measurements of 49 cluster members to estimate the virial mass of the foreground cluster. \\S~\\ref{sec:phot} describes an innovative method used to obtain accurate photometry of the source. A lens model is constructed for the cluster in \\S~\\ref{sec:lensmodel}. We fit spectral energy distribution (SED) models to the broadband photometry of the source to explore its star formation history and stellar population parameters; the SED modeling procedure and results are discussed in \\S~\\ref{sec:sedall}. RCSGA 032727-132609 is compared to the known galaxy population at $z\\sim 2$ in \\S~\\ref{sec:compall}. Throughout this paper we will assume $\\Omega_M = 0.3$, $\\Omega_\\Lambda = 0.7$ and H$_0 = 70$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$. All magnitudes are quoted in the AB system. ", "conclusions": "This paper reports on the discovery of an exceptionally bright and extended star-forming galaxy at $z=1.701$, strongly-lensed by a foreground cluster discovered in the RCS2 survey. The giant arc is $\\sim 3.5$ times brighter than cB58 and extends over 38\\,\\arcsec on the sky. We measure a velocity dispersion of 988$\\pm$122\\,km~s$^{-1}$ for the cluster and estimate a virial mass of $M_{200} \\sim 1.1 \\times 10^{15}$\\,M$_\\odot$\\,$h_{70}^{-1}$, in accord with the large Einstein-radius of $\\sim 17\\farcs8$. A lens model is constructed for the cluster using the publicly-available software {\\tt LENSTOOL} \\citep{Jullo:07} and results in a magnification factor of 2.04$\\pm$0.16 for the counter-image and an estimate of 17.2$\\pm$1.4 for the average magnification of the giant arc, based on the relative sizes of the arc and the counter-image on the sky. Higher resolution imaging is required to create a robust lens model for the giant arc which correctly explains the apparent substructure. Careful measurements of consistent magnitudes for 9 bands of photometry ranging from $u$ to $K_s$ produce a well-constrained spectral energy distribution. SED fitting is based on CB07 models with a Chabrier IMF and Calzetti dust extinction. The systematic uncertainties which arise from the choice of stellar evolution model, IMF and extinction law can be large and often exceed the statistical uncertainties from photometric errors. Our best-fit model places RCSGA 032727-132609 at a metallicity of 0.4\\,Z$_\\odot$ with a moderately young age, $t=80\\pm40$\\,Myr and a relatively small amount of dust extinction, $E(B-V) \\le 0.11$. Taking into account the lensing magnification, we find a stellar mass of M$_* \\sim 10^{10}$\\,M$_\\odot$ and a current star formation rate $SFR \\le 77$\\,M$_\\odot$\\,yr$^{-1}$. The agreement between the stellar population parameters independently inferred from the giant arc and counter-image, both images of the same background galaxy, is a good consistency check of the photometry, SED fitting procedure and lens modeling. Single-component SED models report ages and stellar mass measurements for the current episode of star formation. When allowing for an episodic star formation history, twice this mass could be present in an older underlying burst without affecting the SED detectably. A comparison to the characteristic L$_*$ and M$_{\\mathrm{star}}^*$ of the Schechter luminosity and mass functions at $z\\sim1.7$ places RCSGA 032727-132609 at a luminosity of $6.5\\pm1.8$\\,L$_*$ and a stellar mass of $0.16\\pm0.04$\\,M$_{\\mathrm{star}}^*$, suggesting a low-mass starbursting galaxy. The Salpeter stellar mass estimate of RCSGA 032727-132609 is consistent with the average stellar mass for a sample of `BM' galaxies at $z = 1.72 \\pm 0.34$ \\citep{Reddy:06a} and its age, reddening and current star formation rate generally fall into the broad range of stellar population parameters found for this comparison sample. The large magnification and spatial extent of RCSGA 032727-132609 provide a unique opportunity to spatially resolve one representative example of the variety of stellar populations found at these redshifts and study it in great detail. This paper has presented only our first level of understanding of this unique galaxy." }, "1005/1005.0630_arXiv.txt": { "abstract": "% We present deep {\\it Spitzer} mid-infrared spectroscopy, along with 16, 24, 70, and 850\\,$\\micron$\\ photometry, for 22 galaxies located in the Great Observatories Origins Deep Survey-North (GOODS-N) field. The sample spans a redshift range of $0.6\\la z \\la 2.6$, 24~$\\mu$m flux densities between $\\sim$0.2$-$1.2 mJy, and consists of submillimeter galaxies (SMGs), X-ray or optically selected active galactic nuclei (AGN), and optically faint ($z_{AB}>25$\\,mag) sources. We find that infrared (IR; $8-1000~\\micron$) luminosities derived by fitting local spectral energy distributions (SEDs) with 24~$\\micron$ photometry alone are well matched to those when additional mid-infrared spectroscopic and longer wavelength photometric data is used for galaxies having $z\\la1.4$ and 24~$\\micron$-derived IR luminosities typically $\\la 3\\times 10^{12}~L_{\\sun}$. However, for galaxies in the redshift range between $1.4\\la z \\la 2.6$, typically having 24~$\\micron$-derived IR luminosities $\\ga 3\\times 10^{12}~L_{\\sun}$, IR luminosities are overestimated by an average factor of $\\sim$5 when SED fitting with 24~$\\micron$ photometry alone. This result arises partly due to the fact that high redshift galaxies exhibit aromatic feature equivalent widths that are large compared to local galaxies of similar luminosities. Through a spectral decomposition of mid-infrared spectroscopic data, we are able to isolate the fraction of IR luminosity arising from an AGN as opposed to star formation activity. This fraction is only able to account for $\\sim$30\\% of the total IR luminosity among the entire sample. ", "introduction": "The mid-infrared ($5-40~\\micron$) SED is a complex interplay of broad emission features thought to arise from polycyclic aromatic hydrocarbon (PAH) molecules, silicate absorption features at 9.7 and 18~$\\micron$, and a mid-infrared continuum from very small grains \\citep{lp84,atb85}. Mid-infrared luminosities measured near $\\sim$8~$\\micron$ have been found to correlate with the total infrared (IR; $8-1000$) luminosities of galaxies in the local Universe \\citep{ce01, de02}, which itself is a measure of a galaxy's SFR. While a correlation is found, there does exist a large amount of scatter \\citep{dd05,jd07,la07} and systematic departures for low metallicity systems \\citep{ce05,sm06}. It is therefore uncertain if these empirical relations between mid-infrared and total IR luminosities ($L_{\\rm IR}$) for galaxies in the local Universe are applicable for galaxies at higher redshifts. Recently, data from deep far-infrared surveys such as the Far-infrared Deep Legacy Survey (FIDEL; PI: M. Dickinson) are confirming the same redshift evolution of the SFR density seen in the mid-infrared and do not show any evolution in the SED of infrared luminous galaxies \\citep[e.g.][]{bm09}. In contrast, \\citet{jr08} have provided evidence that the high redshift lensed mid-infrared selected galaxies might show a factor of $\\sim$2 stronger rest-frame 8~$\\micron$ emission compared to their total IR luminosities. Using mid-infrared spectroscopy for 22 sources selected at 24~$\\micron$ in the GOODS-N field, along with existing 850~$\\micron$ and additional 70~$\\micron$ imagery obtained as part of FIDEL, we aim to improve our understanding of the star formation and AGN activity within a diverse group of $0.6\\la z \\la2.6$ galaxies. This is done through a more proper estimate of IR luminosities and the ability to decompose these measurements into star-forming and AGN components. For the full study, please see \\citet{ejm09}. ", "conclusions": "In the present study we have used observations from the mid-infrared to the submillimeter to properly characterize the IR luminosities for a diverse sample of 22 galaxies spanning a redshift range of $0.6 \\la z \\la 2.6$. In addition, we have used the mid-infrared spectra of these sources to estimate the fractions of their IR luminosities which arise from an AGN. Our conclusions can be summarized as follows: \\begin{enumerate} \\item IR ($8-1000~\\micron$) luminosities derived by SED fitting observed 24~$\\micron$ flux densities alone are well matched to those when additional mid-infrared spectroscopy and 16, 70, and 850~$\\micron$ photometry are included in the fits for galaxies having $z \\la 1.4$ and $L_{\\rm IR}^{\\rm 24}$ values typically $\\la$$3\\times10^{12}~L_{\\sun}$. In contrast, for galaxies lying in a redshift range between 1.4 and 2.6 with $L_{\\rm IR}^{\\rm 24}$ values typically $\\ga$$3\\times10^{12}~L_{\\sun}$, IR luminosities derived by SED template fitting using observed 24~$\\micron$ flux densities alone overestimate the true IR luminosity by a factor of $\\sim$5, on average, compared to fitting all available data. A comparison between the observed mid-infrared spectra with that of the SEDs chosen from fitting 24~$\\micron$ photometry alone and from fitting all available photometric data demonstrates that local high luminosity SED templates show weaker PAH emission by an average factor of $\\sim$5 in this redshift range and do not properly characterize the contribution from PAH emission. \\item After decomposing the IR luminosity into star forming and AGN components, we find the AGN luminosity to be increasing with increasing difference between the 24~$\\micron$-derived and our best-fit IR luminosities. Such a trend suggests that the AGN power increases with mid-infrared luminosity. However, we also find that the median fraction of the AGN to the difference between the 24~$\\micron$-derived and best-fit IR luminosities is only 16\\% suggesting the AGN power is almost negligible compared to the bolometric correction necessary to properly calibrate the 24~$\\micron$-derived IR luminosities. \\end{enumerate}" }, "1005/1005.0295_arXiv.txt": { "abstract": "{ We study a coupled dark energy--dark matter model in which the energy-momentum exchange is proportional to the Hubble expansion rate. The inclusion of its perturbation is required by gauge invariance. We derive the linear perturbation equations for the gauge invariant energy density contrast and velocity of the coupled fluids, and we determine the initial conditions. The latter turn out to be adiabatic for dark energy, when assuming adiabatic initial conditions for all the standard fluids. We perform a full Monte Carlo Markov Chain likelihood analysis of the model, using WMAP 7-year data.} \\preprint{IFT-UAM/CSIC-10-28 \\\\ FTUAM-10-07\\\\ULB-TH/10-15} \\begin{document} ", "introduction": "The true substance of dark energy and dark matter is unknown although it should account for about 95\\% of the matter--energy content of our universe today~\\cite{Komatsu:2010fb}. While the couplings of dark fluids to photons and normal matter are severely constrained~\\cite{Carroll:1998zi}, nothing prevents dark matter--dark energy interactions~\\cite{Damour:1990tw,Damour:1990eh,Wetterich:1994bg,Amendola:1999er,Zimdahl:2001ar,Farrar:2003uw,Das:2005yj,Zhang:2005jj,delCampo:2006vv,Bean:2007nx,Olivares:2007rt,Valiviita:2008iv,He:2008si,Gavela:2009cy,Jackson:2009mz,Majerotto:2009np,Valiviita:2009nu,Koyama:2009gd,Boehmer:2009tk}. At the level of the background evolution equations, it is customary to parametrize the coupling between the two dark sectors~\\cite{Kodama:1985bj} as: \\begin{eqnarray} \\label{eq:EOMm} \\dot{\\bar{\\rho}}_{dm}+ 3{\\mathcal{H}}\\bar{\\rho}_{dm} &=&a \\overline{Q}_{dm}\\,,\\\\ \\label{eq:EOMe} \\dot{\\bar{\\rho}}_{de}+ 3 {\\mathcal{H}}\\bar{\\rho}_{de}(1+ w)&=&a \\overline{Q}_{de}\\,, \\end{eqnarray} where $\\bar\\rho_{dm}, \\, \\bar\\rho_{de}$ denote the dark matter and dark energy energy densities, respectively, and $\\overline{Q}_{dm}=-\\overline{Q}_{de}$ encodes the coupling between those two dark sectors and drives the energy exchange between them. The dot indicates derivative with respect to the conformal time $d\\tau = dt/a$, with $\\mathcal{H}={\\dot a}/a \\equiv a \\overline{H}$ denoting the background expansion rate, while $w \\equiv w_{de} = \\bar p_{de}/\\bar \\rho_{de}$ stands for the background dark energy equation of state and pressureless dark matter is assumed: $w_{dm}=\\bar p_{dm}/\\bar \\rho_{dm}=0$. From now on, barred quantities are to be considered as the background quantities. The initial conditions for the several components populating the early universe have been explored to a large extent. They were first analyzed for all cosmic fluids but dark energy (see {\\it e.g.} Ref.~\\cite{Ma:1995ey} and references therein), with the result that adiabatic initial conditions were one possibility. It was also noticed that the choice of gauge could be a delicate issue: a safe alternative proposed was to use a gauge invariant formalism \\cite{Bardeen:1980kt,Kodama:1985bj, Mukhanov:1990me,Durrer:2001gq}. The initial conditions for the case of dynamical dark energy as an uncoupled quintessence field have been also derived~\\cite{Viana:1997mt,Dave:2002mn, Malquarti:2002iu,Abramo:2001mv,Kawasaki:2001nx,Perrotta:1998vf,Doran:2003xq}, including a gauge invariant treatment~\\cite{Doran:2003xq}: they turned out to be adiabatic if those for the traditional fluids were adiabatic. Furthermore, the formalism in Ref.~\\cite{Doran:2003xq} has been recently applied to the case of a coupled dark energy-dark matter systems which mimic uncoupled models at early times, both at the background and perturbation levels~\\cite{Majerotto:2009np} for the viable parameter space: as expected, adiabatic initial conditions for dark energy naturally resulted then. Here we consider a different class of dark couplings, not negligible at early times. It is also illustrated that the gauge invariant formalism is particularly illuminating for the determination of the correct perturbation equations, for a general coupled theory. The structure of the paper is as follows. In Section~\\ref{sec:gauge-invar-pert}, the notation is set and the gauge invariant equations -at linear order in perturbation theory- for a coupled fluid are derived. In particular, we study in Section~\\ref{sec:coupl-prop-h} the case of a (covariant) dark matter--dark energy interaction proportional to the Hubble rate. In Section~\\ref{sec:initial-conditions}, following the method proposed in Ref.~\\cite{Doran:2003xq}, we derive the corresponding initial conditions for dark energy. Then in Section~\\ref{sec:data-constraints}, we constrain the type of coupled models analyzed, using several data sets. Section~\\ref{sec:conclusion} contains the conclusions. ", "conclusions": "\\label{sec:conclusion} Interacting dark energy-dark matter cosmologies in which the coupling term is proportional to the Hubble expansion rate are revisited. While in previous works the perturbation in the Hubble expansion rate was neglected, it is illustrated here how the inclusion of such a term is mandatory to satisfy the gauge invariance of the theory. It also serves as a guide to define a covariant formulation of the dark sector interaction. In this work, the latter has been chosen to be expressed in terms of the expansion rate associated to the total fluid. This choice is however not unique, we could have used the expansion rate of any other fluid. For the case under study, we compute the linear perturbation evolution using a gauge invariant formalism. After imposing adiabatic initial conditions on the matter and radiation fluids, we find that the initial conditions for the coupled dark energy fluid are also adiabatic. This result is independent of the choice in the covariant formulation of the expansion rate. The new terms arising from the expansion rate perturbation have negligible quantitative impact on the constraints on cosmological parameters previously obtained in the literature. A new analysis has been performed using the latest WMAP7 data." }, "1005/1005.2290_arXiv.txt": { "abstract": "{} {FIR imaging of interacting galaxies allows locating even hidden sites of star formation and measuring of the relative strength of nuclear and extra-nuclear star formation. We want to resolve the star-forming sites in the nearby system of the Antennae.} {Thanks to the unprecedented sharpness and depth of the PACS camera onboard ESA's {{\\em Herschel}} Space Observatory, it is possible for the first time to achieve a complete assessment of individual star-forming knots in the FIR with scan maps at 70, 100, and 160\\,$\\mu$m. We used clump extraction photometry and SED diagnostics to derive the properties related to star-forming activity.} {The PACS 70, 100, and 160\\,$\\mu$m maps trace the knotty structure of the most recent star formation along an arc between the two nuclei in the overlap area. The resolution of the starburst knots and additional multi-wavelength data allow their individual star formation history and state to be analysed. In particular, the brightest knot in the mid-infrared (K1), east of the southern nucleus, exhibits the highest activity by far in terms of dust heating and star formation rate, efficiency, and density. With only 2\\,kpc in diameter, this area has a 10--1000\\,$\\mu$m luminosity, which is as high as that of our Milky Way. It shows the highest deficiency in radio emission in the radio-to-FIR luminosity ratio and a lack of X-ray emission, classifying it as a very young complex. The brightest 100 and 160\\,$\\mu$m emission region (K2), which is close to the collision front and consists of 3 knots, also shows a high star formation density and efficiency and lack of X-ray emission in its most obscured part, but an excess in the radio-to-FIR luminosity ratio. This suggests a young stage, too, but different conditions in its interstellar medium. Our results provide important checkpoints for numerical simulations of interacting galaxies when modelling the star formation and stellar feedback.} {} ", "introduction": "We have set up a photometric imaging programme of a sample of nearby interacting galaxies as part of the {\\em Herschel} Guaranteed Time Key Project SHINING\\footnote{http://www.mpe.mpg.de/ir/Research/SHINING/} to trace and resolve the youngest star formation sites that are still embedded in dust, making them strong far-infrared (FIR) emitters. The \\object{Antennae} (\\object{Arp~244}, \\object{NGC~4038}/\\object{NGC~4039}) is a key object in this sample, since it is an interactive system, where strong extranuclear star formation was found previously in the so-called overlap area of the two disks between the two nuclei. Optical images show dark dust lanes interspersed with a few HII regions in this area \\citep[e.g.][]{whitmore95}. It houses several supergiant molecular cloud complexes \\citep{stanford90,wilson00,gao01,schulz07}, which provide a large reservoir for bursts of star formation. Prominent 6 and 20\\,cm radio emission \\citep{hummel86}, which spatially matches the CO emisison peaks, suggests there are ongoing starbursts. Compressed magnetic fields are found towards the northern edge indicating pre-starbursts \\citep{chyzy04}. Mid-infrared (MIR) observations have discovered embedded starbursts at the southeastern border of the overlap region \\citep{vigroux96,mirabel98,wilson00}, while submm 850 $\\mu$m observations suggest a large amount of cooler dust in the northern overlap region, arguing in favour of relatively less powerful starbursts there \\citep{haas00,zhu03,schulz07}. In the case of buried starbursts, one expects to see the continuum re-emission of the hiding dust at mid- and far-infrared wavelengths. However, previous FIR space missions have not reached the spatial resolution required to separate the overlap region from the nuclei. {\\em Herschel}-PACS now offers the capability of such spatially resolved observations for the first time. ", "conclusions": "New {\\em Herschel}-PACS scan maps of the Antennae at 70, 100, and 160\\,$\\mu$m provide a high spatial resolution complement to earlier MIR maps (ISOCAM, {\\em Spitzer}-MIPS), high-resolution CO (1-0), and radio 6\\,cm maps and thus allow detailed study of the star formation state in individual emission knots. We confirm that the highest star formation activity takes place in the overlap area concentrated in the two emission complexes K1 and K2. Our star formation diagnostics strongly indicates that these knots differ in evolutionary stage." }, "1005/1005.0589_arXiv.txt": { "abstract": "Many global circulation models predict supersonic zonal winds and large vertical shears in the atmospheres of short-period jovian exoplanets. Using linear analysis and nonlinear local simulations, we investigate hydrodynamic dissipation mechanisms to balance the thermal acceleration of these winds. The adiabatic Richardson criterion remains a good guide to linear stability, although thermal diffusion allows some modes to violate it at very long wavelengths and very low growth rates. Nonlinearly, wind speeds saturate at Mach numbers $\\approx 2$ and Richardson numbers $\\lesssim 1/4$ for a broad range of plausible diffusivities and forcing strengths. Turbulence and vertical mixing, though accompanied by weak shocks, dominate the dissipation, which appears to be the outcome of a recurrent Kelvin-Helmholtz instability. An explicit shear viscosity, as well as thermal diffusivity, is added to ZEUS to capture dissipation outside of shocks. The wind speed is not monotonic nor single valued for shear viscosities larger than about $10^{-3}$ of the sound speed times the pressure scale height. Coarsening the numerical resolution can also increase the speed. Hence global simulations that are incapable of representing vertical turbulence and shocks, either because of reduced physics or because of limited resolution, may overestimate wind speeds. We recommend that such simulations include artificial dissipation terms to control the Mach and Richardson numbers and to capture mechanical dissipation as heat. ", "introduction": "Strongly irradiated extrasolar giant planets---``hot Jupiters''---are expected to rotate synchronously with their orbits but to have strong longitudinal winds that redistribute heat from the day to the night side. The efficiency of redistribution is important for direct observables including infrared phase curves, the depth of secondary eclipses in transiting systems (i.e., eclipses of the planet by its star), and planetary spectra. Turbulence associated with the winds may contribute to chemical mixing of the atmosphere \\citep{Spiegel_Silverio_Burrows09}, and might even inject heat into the convective interior of the planet \\citep{Guillot_Showman02}, thereby perhaps explaining why some of these planets have lower densities than expected by standard evolutionary models at their estimated ages. In a previous paper \\citep[hereafter Paper I]{Goodman2009}, one of us has argued that these thermally driven winds can be understood as natural heat engines, which convert a fraction of the thermal power into mechanical work: namely, the work expended to accelerate the wind. As with any other heat engine, the continual production of work must be balanced by mechanical dissipation, else the kinetic energy in the winds would grow without bound. Paper I offered a brief discussion of possible dissipation mechanisms. Here we explore those mechanisms in greater detail. The goal is to lay the foundation for subgrid models of the dissipative process suitable for use by global circulation codes. The outline of this paper is as follows. \\S\\ref{sec:overview} gives an overview of the dissipation mechanisms we consider. \\S\\ref{sec:linear} presents a linear analysis of hydrodynamic Kelvin-Helmholtz instabilities of thermally stratified shear flows. We show that such flows can be destabilized by thermal diffusivity even at Richardson numbers greater than the well-known critical value $Ri_{\\rm crit}=1/4$. Under conditions relevant to exoplanet winds, however, we estimate that the associated growth rates are too slow and the longitudinal wavelengths too long to be important for dissipation, unless $Ri<1/4$, which requires transsonic flow unless the vertical width of the shear layer is small compared to a pressure scale height. Hence shocks may be as important as shear-driven turbulence for dissipation. To investigate this, as described in \\S\\ref{sec:nonlinear}, we have used the ZEUS code in two dimensions to simulate a part of the atmosphere with horizontal and vertical dimensions comparable to the pressure scale height. Thermal diffusion and viscosity have been added to the code, and thermal driving of the wind is simulated by a horizontal body force with nonzero curl. We study the velocity and dissipation rate of the wind in a statistical steady state as a function of the strength of the driving compared to the acceleration of gravity. \\S\\ref{sec:discussion} discusses our results in the context of previous work on winds in jovian planets both within and beyond the solar system. \\S\\ref{sec:conclusions} summarizes our main conclusions. ", "conclusions": "\\label{sec:conclusions} We have investigated the balance between horizontal forcing and mechanical dissipation in model shear flows designed to imitate thermally driven winds of jovian exoplanets. Our main conclusions are as follows. \\begin{enumerate} \\item Following Paper I, some form of mechanical dissipation is required to offset the production of mechanical energy by the longitudinal entropy gradient. Plausible dissipation mechanisms include shocks, turbulence, and MHD drag. \\item Entropy stratification has a stabilizing influence on Kelvin-Helmholtz (KH) modes, but thermal diffusivity ($\\chi$) tends to undercut this influence to a degree that can be quantified by the Peclet number $Pe\\equiv\\cs H_p/\\chi$ based on the sound speed ($\\cs$) and pressure scale height ($H_p$). We estimate $Pe\\sim O(10^3)$ at the altitude where the thermal timescale is comparable to the circumferential flow time at Mach~1, but $Pe\\lesssim O(1)$ near the infrared photosphere. \\item Linear analysis indicates that for $Pe\\gtrsim 10$, thermal diffusion has little effect on the stability of KH modes whose horizontal wavelength is comparable to the vertical thickness of the shear layer; the usual adiabatic Richardson criterion $Ri\\equiv N^2/(\\partial U/\\partial z)^2>1/4$ is sufficient for the stability of these short-wavelength modes. For typical conditions on the day sides of hot Jupiters, $Ri<1/4$ implies transsonic or supersonic winds, presuming that the shear layer is no narrower than a pressure scale height. \\item Nevertheless, at any finite $Pe$ and $Ri$, inviscid flows with an inflection point in the shear profile are unstable at sufficiently long horizontal wavelengths, at least when the flow is confined within a channel of finite vertical extent. We doubt that these long-wavelength KH modes are important for hot-Jupiter winds because their growth rates are very small (inversely proportional to wavelength) and because the planetary circumference is not large enough compared to $H_p$ to accommodate them at $Pe\\gtrsim 10^3$. \\item We estimate a turbulent diffusivity for composition in hot Jupiter atmospheres. For $Pe\\approx 100$, we find $K_{\\rm t}/c_sH_p\\approx 10^{-3}$, depending on the strength of the horizontal forcing. However, these values pertain to regions of maximum vertical shear; turbulent diffusivities may be much smaller elsewhere. \\item Nonlinear, two-dimensional (azimuth and depth), horizontally forced local simulations saturate at time-averaged Mach numbers $\\sim 2$ when the Reynolds number $Re\\equiv \\cs H_p/\\nu\\gtrsim 10^3$. Given the widths of our shear layers ($\\sim 2H_p$) and background stratification, this corresponds to minimum Richardson numbers slightly below the adiabatic critical value of $1/4$. In this regime, which is relevant to hot Jupiters, we find that the saturated wind speed depends only weakly on $Pe$ and on the strength of the forcing. \\item Dissipation in the high-$Re$ regime is due mainly to the work expended to overturn the stratification rather than shocks, though weak shocks are present. It is highly intermittent, at least in our two-dimensional calculations, and appears to be triggered by a recurrent linear instability of Kelvin-Helmholtz type. \\item At $Re\\lesssim 10^3$ or at low numerical resolution (significantly fewer than 10 cells per pressure scale height), multiple sequences of solutions appear in overlapping ranges of Reynolds number, differing in their wind speed and predominant dissipation mechanisms. The fastest have Mach numbers at least twice as large as what we believe to be the correct inviscid values for our forcing strengths. \\end{enumerate} The last three points suggest that some global simulations of hot Jupiters may have overestimated wind speeds due to incomplete physics (e.g. neglect of vertical accelerations and sound waves) and/or insufficient spatial resolution. Pending sufficient computational resources to resolve the physical dissipation mechanisms directly in global simulations, we suggest that artificial dissipative terms be added to the equations of motion so as to prevent strongly supersonic shears and small Richardson numbers. The appropriate form for these terms will depend upon the equation sets and algorithms used, but presumably it is desirable that they conserve energy and momentum, and probably that they have a sharp threshold in Mach number or Richardson number." }, "1005/1005.3858_arXiv.txt": { "abstract": "The fragile structure of chondritic-porous interplanetary dust particles (CP- IDPs) and their minimal parent-body alteration have led researchers to believe these particles originate in comets rather than asteroids where aqueous and thermal alteration have occurred. The solar elemental abundances and atmospheric entry speed of CP-IDPs also suggest a cometary origin. With the return of the Stardust samples from Jupiter-family comet 81P/Wild 2, this hypothesis can be tested. We have measured the Fe oxidation state of 15 CP-IDPs and 194 Stardust fragments using a synchrotron-based x-ray microprobe. We analyzed $\\sim$300 nanograms of Wild 2 material -- three orders of magnitude more material than other analyses comparing Wild 2 and CP-IDPs. The Fe oxidation state of these two samples of material are $>$2$\\sigma$ different: the CP-IDPs are more oxidized than the Wild 2 grains. We conclude that comet Wild 2 contains material that formed at a lower oxygen fugacity than the parent body, or parent bodies, of CP-IDPs. If all Jupiter-family comets are similar, they do not appear to be consistent with the origin of CP-IDPs. However, comets that formed from a different mix of nebular material and are more oxidized than Wild 2 could be the source of CP-IDPs. ", "introduction": "\\label{intro} Chondritic-porous interplanetary dust particles (CP-IDPs), collected in the stratosphere by high-altitude aircraft, are widely thought to originate in comets \\citep{Schramm:1989p3433,Bradley:1994p2672}. The collisional history of asteroids makes it unlikely that these particles (fragile, porous aggregates of small grains) are asteroidal \\citep{Brownlee:1985p2339}. The atmospheric entry speed of CP-IDPs, as deduced from thermal release profiles of solar wind He \\citep{Nier:1992p3404}, is consistent with cometary, rather than asteroidal, orbits \\citep{Brownlee:1995p2673}. Infrared and electron beam studies of IDPs show that those consisting of mostly anhydrous minerals are usually chondritic-porous, whereas IDPs rich in phyllosilicates are usually chondritic-smooth \\citep{Sandford:1985p3456, Bradley:1986p3458}. The presence of GEMS (glass with embedded metal and sulfides) in IDPs indicates a lack of thermal alteration \\citep{Bradley:1994p2672}. CP-IDPs, therefore, show very little parent-body alteration and must originate from either anhydrous objects or hydrous objects that have been kept at very low temperature -- again, consistent with cometary origin. Nevertheless, a cometary origin of CP-IDPs is not universally accepted \\citep{Flynn:1992p3174, Thomas:1995p3175}. With the return of samples from comet 81P/Wild 2 by NASA's Stardust mission \\citep{Brownlee:2006p1376}, it is now possible to compare CP-IDPs with material from this Jupiter-family comet. We seek to prove or disprove, with a known level of confidence, the hypothesis that CP-IDPs originate from parent bodies with the composition of comet Wild 2. The oxidation state of Fe is a clear mineralogic indicator of the oxidation state of meteorites \\citep{Rubin:1988p2665}. Meteorite groups are in fact distinguished from each other by their differing Fe oxidation states, as Urey and Craig first reported more than fifty years ago \\citep{Urey:1953p1396}. Parent body processing can drastically change the Fe oxidation state of some of the meteorite groups, obscuring information about the original oxidation state the material acquired in the solar nebula \\citep{Rubin:1988p2665}. Carbonates and Fe-bearing crystalline silicates, possible products of aqueous alteration on asteroids \\citep{Krot:1995p3169}, are found alongside anhydrous minerals in the same Stardust track \\citep{Flynn:2008p2675}, and therefore co-existed in one large aggregate in comet Wild 2 . No phyllosilicates have been unambiguously identified in the Stardust samples \\citep{Zolensky:2008p1420}. With these and other pieces of evidence, \\citet{Wooden:2008p2674} argues that instead of selective aqueous alteration on submicron scales in the comet itself, grains which formed in different regions of the solar nebula under varying reduction-oxidation conditions (e.g. Mg- and Fe-rich crystalline silicates) migrated, aggregated, and formed comet Wild 2. Likewise, the oxidation state of anhydrous CP-IDPs was unaffected by parent body processing \\citep{Schramm:1989p3433,Zolensky:1995p1397}, so a comparison of the Fe oxidation state of comet Wild 2 with anhydrous CP-IDPs yields insight into the nebular environment in which they formed. We report on the oxidation state of Fe in 15 anhydrous CP-IDPs, and directly compare these measurements to the Fe oxidation state of comet Wild 2, deduced from 194 Stardust fragments in 11 aerogel tracks. ", "conclusions": "The presence of Fe in its different oxidation states is one of the clearest mineralogic indicators of the oxidation state of meteoritic material \\citep{Rubin:1988p2665}. As discussed earlier, the Wild 2 samples and CP-IDPs are unlikely to have experienced much parent-body processing that would drastically change their oxidation state, and therefore a measurement of the Fe oxidation state of these two materials is a probe into the oxygen fugacity of the environment in which they formed. Figure \\ref{figure:ternary} shows that comet Wild 2 and CP-IDPs differ in their Fe oxidation state at more than the 2$\\sigma$ level. \\citet{Ishii:2008p962} reported several differences between CP-IDPs and the Wild 2 material: the Stardust material lacks GEMS (glass with embedded metal and sulfides) and also lacks elongated enstatite whiskers and platelets, all of which are present in CP-IDPs. They conclude from these transmission electron microscopy observations that the Wild 2 material is dissimilar to CP-IDPs and more closely resembles asteroidal material. Our results, analyzing $\\sim$1000$\\times$ more material, indicate that comet Wild 2 differs at more than the 2$\\sigma$ confidence level from CP-IDPs, and Wild 2 is also distinct from meteorites. CP-IDPs are 2$\\sigma$ inconsistent with all meteorite groups except CI and R. The significant difference between the Fe oxidation state of Wild 2 and CP-IDPs is that a significant fraction of Fe in Wild 2 exists as Fe$^{0}$ whereas CP-IDPs contain little Fe$^{0}$ compared to Fe sulfide and oxidized Fe. We can compare these results only qualitatively to, for example, TEM observations of mineral phases present in CP-IDPs: the Fe XANES technique used in this work can determine how many Fe atoms in the entire CP-IDP are in a given Fe oxidation state, whereas the the entire volume of the IDP is not always examined by TEM, and the amount of material in a given phase is seldom recorded. \\citet{MacKinnon:1987p2660} tabulate a collection of phases observed in 30 CP-IDPs. Of the 30, Fe-bearing mineral phases are reported in 23 of the CP-IDPs: 21 contain oxidized Fe, 15 contain Fe sulfide, and 4 contain Fe$^{0}$. Qualitatively, this is consistent with the relative paucity of Fe$^{0}$ we see by the Fe XANES technique employed in this work. \\subsection{Selection bias in CP-IDP collection?} Since Fe metal is denser than Fe sulfides or oxidized Fe, particles containing substantial amounts of Fe metal may have a shorter residence time in the stratosphere and are therefore less likely to be collected. Particles containing Fe metal are also more likely to vaporize than less-dense particles \\citep{Flynn:2008p3176}. Can the effect of these biases skew the composition of collected IDPs such that our observations of the 15 IDPs studied in this work are actually consistent with IDPs of Wild 2 composition before atmospheric entry? To answer this question, we recalculate the Fe oxidation for the 11 Stardust tracks from comet Wild 2 as if they were IDPs, accounting for the effects of atmospheric heating and residence time in the stratosphere. First, we first calculate the residence time of 11 IDPs in the range of 15--25 km of altitude from the formulation given in \\citet{Kasten:1968p3375} and \\citet{Rietmeijer:1993p3303}. The IDPs have the Wild 2 Fe composition as given by the Stardust tracks in Table \\ref{table:fefrac}. The speed of the particle falling through the atmosphere is proportional to its density, so the less-dense particles are more likely to be collected than the more-dense, Fe-metal-rich particles. Secondly, we consider the atmospheric entry heating of IDPs as described by \\citet{Fraundorf:1980p3177}. We calculate the probability that an IDP is destroyed by heating to a temperature above the melting point of Fe metal, 1811$^{\\circ}$ K. The probability of collecting IDP $i$ of all $j$ types of IDPs can be expressed simply as: \\begin{equation} P_i=\\frac{\\left(N_i-N_iV_i\\right)T_i}{\\displaystyle\\sum_j{\\left(N_j-N_jV_j\\right)T_j}} \\label{entryeqn} \\end{equation} where $N_i$ is the number of IDPs of type $i$, $V_i$ is the fraction of IDPs $i$ that vaporize, and $T_i$ is the residence time of IDP $i$ in the stratospheric-collection region. For each of the 11 IDPs of Wild 2 composition, we calculate the probability of collection. Then we calculate the mean and 2$\\sigma$ confidence limits of the Fe oxidation state of these 11 particles using the method described in Section \\ref{confidencelimits}, but here each IDP is weighted by its probability of collection relative to the others. The amount of Fe metal in IDPs of Wild 2 composition, taking into account atmospheric heating and residence time in the stratosphere, is $<$1\\% less than the abundance of Fe metal in the IDPs before atmospheric entry. This cannot explain the observed lack of Fe metal in CP-IDPs compared to comet Wild 2, so we conclude that these possible selection biases in IDP collection are insignificant. \\subsection{Grain sizes in CP-IDPs and Stardust Fragments} CP-IDPs are aggregates of mostly submicron grains whereas the Stardust fragments contain many mono-mineralic grains larger than a micron. Thermal modification or destruction of submicron grains in the bulb \\citep{Leroux:2009p3050,Brownlee:2006p1376} precludes the possibility of comparing the Fe oxidation state of CP-IDPs and Stardust at the same grain size. It may be the case that cometary metal exists only in larger grains and is not present in the submicron components that make up CP-IDPs. While we cannot rule out this possibility with the measurements presented here, we can compare comet Wild 2 and CP-IDP material by looking at smaller grain sizes of Wild 2 (What is the Fe$^{0}$ fraction of the smaller Stardust fragments we analyzed?) and larger grain sizes of CP-IDPs (Do larger mineral grains of CP-IDPs exist and what is their Fe oxidation state?). For each Stardust fragment we analyzed, we know its approximate, relative Fe mass by the absorption edge jump across the Fe K-edge. We compute Fe$^{0}/\\sum\\,$Fe for all Stardust fragments less than a given Fe mass, shown in Figure \\ref{figure:FeMassFrac}. The Fe$^{0}$ fraction increases for the intermediate-mass particles in this study, but is still inconsistent with the CP-IDPs over the entire Fe-mass range of Stardust fragments we analyzed (using the particle-based bootstrap uncertainty calculation). \\begin{figure} \\centering \\includegraphics[width=\\textwidth]{fefracsize} \\caption{Fe$^{0}/\\sum\\,$Fe for all Stardust fragments less than a given Fe mass (computed by the Fe K-edge XANES jump). The 2$\\sigma$ confidence interval for each set of Stardust fragments less than a given mass is shown by the red shaded region. The green region indicates the 2$\\sigma$ confidence interval for the CP-IDPs measured in this work (CP-IDPs consist of mostly submicron grains--smaller grain size than the Stardust fragments analyzed in this work). Both the Stardust and CP-IDP uncertainties shown here are calculated with the particle-based approach described in Section \\ref{confidencelimits}.} \\label{figure:FeMassFrac} \\end{figure} Stratospheric collection yields large ($>$5$\\mu$m) mono-mineralic grains of non-chondritic composition in addition to IDPs. Fine-grained chondritic material (reminiscent of small bits of CP-IDPs) is sometimes found adhering to the surfaces of these grains, indicating they originate from the same parent body or bodies as CP-IDPs \\citep{Flynn:2007p3428,Flynn:2008p3176}. The sizes of these grains are closer to the range of grain sizes of the Stardust fragments we analyzed. The mono-mineralic IDP grains consist predominantly of olivine, pyroxene, and sulfide. They probably do not contain a significant amount Fe metal. Therefore, these larger mono-mineralic IDP grains which likely sample the same parent body as the CP-IDPs, probably do not contain enough Fe$^{0}$ to be consistent with the comet Wild 2 material. If a cosmic dust particle is highly enriched in Fe metal to the exclusion of most other phases, its chemical composition would not be close to chondritic and therefore the particle would not be considered a CP-IDP. Similarly, a particle made mostly of Fe metal may not appear fluffy, and so we may not select it from the Cosmic Dust Catalogs for inclusion in a study such as this. There is an inherent bias in the selection of CP-IDPs for comparison with other samples of matter. Stratospheric-collected particles with high Fe metal abundances have been classified as ``Cosmic\", but the fraction of the collected Fe metal particles which are of true extraterrestrial origin or possibly cometary has not been established. However, if Fe metal were present as submicron grains at the 10\\% level, a particle could still have approximately chondritic composition and appear fluffy, like a CP-IDP. This, we have shown, is not the case. For this type of selection bias to significantly affect our results, the Fe metal distribution in IDPs would have to be very heterogenous: the Fe metal would need to be sequestered into particles which contain enough Fe metal to appear either non-chondritic or non-porous. Our analysis compares sub-micron-grained CP-IDPs with larger-grained Stardust fragments, but as the Stardust fragments show more Fe$^{0}$ than CP-IDPs over the entire range of particle sizes we measured, and large mono-mineralic grains which likely originated from the same source material as IDPs also show a dearth of Fe$^{0}$, we conclude it is likely that the CP-IDPs contain less Fe$^{0}$ than the Wild 2 samples at comparable grain sizes. We acknowledge a possible selection bias in choosing chondritic and porous IDPs for this study, though our results are valid under the assumption that Fe metal is not distributed heterogeneously, with scarce, very Fe-metal-rich particles containing the Fe metal missing in CP-IDPs. \\subsection{Fe metal in Stardust samples} The Fe metal we observe in the Wild 2 samples is unlikely to be a capture effect, as discussed earlier. \\citet{Zolensky:2006p1378} report that the high abundance of Ni in metal droplets in Stardust tracks shows that there must have been metal intrinsic to the comet. Much of the Fe metal we measure exists in large particles in the terminus of the track. One of the terminal particles in Track 77 was observed in TEM to be an Fe metal grain encased in olivine (D. Joswiak, private communication), reminiscent of Fe metal encapsulated in silicate grains in unequilibrated chondrites \\citep{Rambaldi:1980p2671}. The terminal particle of Track 41 is a large piece of Fe metal (as shown by the Fe XANES, and an x-ray diffraction match to kamacite and taenite) with a bit of Fe sulfide and Fe$^{2+}$ (as seen in the chemical map of the particle, Figure \\ref{figure:simeio}). The Fe metal in large terminal particles like these almost certainly existed as metal in the comet, which makes Wild 2 distinct from CP-IDPs. \\begin{figure} \\centering \\includegraphics[width=\\textwidth]{simeio_chemical_map_figure_blue_gamma} \\caption{Iron chemical map \\citep{Ogliore:2008p3135} of the the large terminal particle in Track 41, ``Simeio\", in two orientations that differ by a 90$^{\\circ}$ rotation. Simeio is mostly a large mass of Fe$^{0}$. The mineral phase of this material was identified as kamacite and taenite by x-ray diffraction (XRD). Iron sulfide is present in small amounts throughout the particle, as well as both Fe$^{2+}$ and sulfide on the particle perimeter in both orientations, possibly indicating a thin, partial shell of sulfide and Fe$^{2+}$. } \\label{figure:simeio} \\end{figure} \\subsection{Nebular mixing, CP-IDPs, and comet Wild 2} The oxidation state of Fe can be used as an indicator of the oxygen fugacity in which the grain formed \\citep{Rubin:1988p2665}. These metal grains likely formed in the high-temperature, low-O-fugacity conditions of the inner region of the early solar nebula (a reducing environment \\citep{Wooden:2008p2674}). Refractory inclusions from the inner nebula have been identified in the Stardust samples \\citep{Simon:2008p2065}. Our Fe XANES survey of Stardust tracks imply that a fraction between 0.5 and 0.65 (2$\\sigma$) of Wild 2 material came from the inner nebula \\citep{Westphal:2009p1874}, suggesting that there was large-scale mixing in the solar nebula \\citep{Ogliore:2009p2695}. Stardust olivines show a broad distribution of Mg/Fe composition \\citep{Zolensky:2006p1378} which indicates the comet consists of grains that formed in a wide range of reducing/oxidizing conditions \\citep{Rubin:1988p2665}. Therefore, it is not unexpected that the oxidation state of Fe in Wild 2 should reflect both the inner nebula (Fe metal grains) and the outer nebula (Fe-rich silicates \\citep{Wooden:2008p2674}). The fact that CP-IDPs show much less Fe metal than Wild 2 indicate that these particles did not originate from a body, or bodies, that sampled the regions of low oxygen fugacity that Wild 2 did. Comets can contain different amounts of low and high oxygen fugacity material, as \\citet{Wooden:2008p2674} states: ``Hence, the efficiency of incorporation of low vs. high oxygen fugacity products into cometary nuclei depends on the interplay between the duration of low vs. high water vapor content, the time-dependent outward mass transport rate, and the distance(s), range of radial migration, and duration of cometesimal accretion and nuclei accumulation\". CP-IDPs could be derived from comets that did not sample the highly mixed nebula as Wild 2 did, perhaps one that formed in a different time or place. Our results are inconsistent with the hypothesis that CP-IDPs originate from parent bodies with the composition of comet Wild 2 at $>$2$\\sigma$ confidence level. We have not, however, disproved the hypothesis that CP-IDPs originate from Jupiter-family comets because we do not know if all Jupiter-family comets have an Fe oxidation state that is consistent with comet Wild 2. Comet Tempel 1, the target of NASA's Deep Impact mission, shows significant differences from Wild 2 in composition \\citep{Brownlee:2006p1376} and morphology of the cometary nucleus \\citep{Thomas:2007p3161}. These differences could be due to the longer period of time Tempel 1 spent in the inner solar system, the different regions from which the samples originated, the different size-range of analyzed material, and many other complicating factors. However, it is also possible that the two comets are composed of fundamentally different material that formed in different places and at different times in the solar nebula. From orbital dynamics, \\citet{Liou:1996p3165} show that dust from Tempel 1 could contribute significantly to the population of low-entry-speed CP-IDPs collected at Earth. This work has shown that a cometary source of CP-IDPs must contain less inner nebula material (where the oxygen fugacity is low) than comet Wild 2--we do not rule out a CP-IDP source from comet Tempel 1 or the numerous other comets of unknown composition." }, "1005/1005.3543_arXiv.txt": { "abstract": "{We present five band imaging of the Vega debris disc obtained using the \\emph{Herschel Space Observatory}. These data span a wavelength range of 70--500\\,$\\mu$m with full-width half-maximum angular resolutions of 5.6-36.9\\arcsec. The disc is well resolved in all bands, with the ring structure visible at 70 and 160\\,$\\mu$m. Radial profiles of the disc surface brightness are produced, and a disc radius of 11\\arcsec\\ ($\\sim 85$\\,AU) is determined. The disc is seen to have a smooth structure thoughout the entire wavelength range, suggesting that the disc is in a steady state, rather than being an ephemeral structure caused by the recent collision of two large planetesimals.} ", "introduction": "Debris discs, of which the \\object{Vega} ($\\alpha$ Lyrae) disc is the archetype, are characterised as discs of dusty material generated by the collision of planetesimals in belts surrounding main sequence stars. The ages of the stars which exhibit these discs ($\\sim$ 350\\,Myr in the case of Vega; \\citealt{Song2000}) precludes the possibility for this dust to be primordial, as the time scale to remove such dust is $\\la$10\\,Myr \\citep{Backman1993,Wyatt2008}. The debris disc around Vega was first detected by \\cite{Aumann1984} as an infrared excess using the Infrared Astronomical Satellite (\\emph{IRAS}; \\citealt{IRAS1984}), and has been extensively studied in the infrared and submillimetre over the subsequent 25 years \\citep[e.g.][]{Holland1998, Wilner2002, Su2005, Marsh2006}. The appearance of the disc has been found to vary significantly across this wavelength regime, changing from a smooth axisymmetric structure in the infrared \\citep[hereafter S05]{Su2005}, to a structure in the submillimetre, wherein the majority of the emission lies in two discrete clumps \\citep{Holland1998}. In order to understand the reason for the variation in structure with wavelength it is important to first understand the origin of the clumps seen in the submillimetre. The recent collision of two massive planetesimals is one option, however, given the age of Vega, the statistical likelihood of this occurring with two bodies of sufficient mass to explain the submillimetre observations is low \\citep{Wyatt2002}. A more favourable alternative, first proposed by \\cite{Wilner2002} and modelled by \\citet[hereafter W06]{Wyatt2006} and \\cite{Reche2008}, is that the clumps are dust grains trapped in resonance with a planet near to the disc. In this scenario the large dust grains (larger than a few mm) are trapped in these resonances, while smaller intermediate sized grains (a few $\\mu$m--mm), having been perturbed by radiation pressure, have a more uniform distribution in the disc. Recent analysis and modelling of \\emph{Spitzer} \\citep{Werner2004} mid-infrared data have reached contradictory conclusions. S05 find the disc to be ephemeral; in this scenario the disc is the result of a recent massive collision of planetesimals, and the subsequent collisional cascade. This results in a high mass of very small grains (less than a few $\\mu$m) which are blown out of the system by radiation pressure immediately upon creation, resulting in the large disc extent observed. Conversely, \\cite{Muller2010} succeed in reproducing the surface brightness radial profile using intermediate size grains in elliptical orbits around the parent planetesimal ring, and therefore conclude that it is consistent with a steady-state model. In the steady-state model, dust that is destroyed, either by being drawn in to the star due to Poynting-Robertson drag or blown out of the disc by radiation pressure, is continuously replenished by a steady collisional cascade within the planetesimal belt. If the small blown-out grains are the origin of the emission observed in the mid-infrared then W06 predicts that spiral features, emanating from the submillimetre clumps, should be visible with high-resolution imaging; a smooth structure would support the steady-state model. In this paper we present five-band far-infrared imaging of the Vega debris disc obtained with the \\emph{Herschel} \\citep{Pilbratt2010} Photodetector Array Camera and Spectrometer \\citep[PACS;][]{Poglitsch2010} and Spectral and Photometric Imaging Receiver \\citep[SPIRE;][]{Griffin2010}. We discuss the initial analysis and disc parameterisation, and relate these results to the ephemeral and steady-state disc models. In Section 2 we present the \\emph{Herschel} data, outline the processing performed, and analyse the disc structure and properties. These data are then compared with results from the recent \\emph{Spitzer} observations (S05) and disc modelling of W06 and \\cite{Muller2010}, with our conclusions summarised in Section 3. ", "conclusions": "The structure observed in the \\emph{Herschel} data shows no sign of clumps. There are also no visible spiral arm features, predicted by W06, if the disc emission at mid-to far-infrared wavelengths is dominated by small blown-out dust grains (W06 Fig.~3; right hand panels, $\\beta=$1--10). The smooth structure observed is most consistent with the steady-state model, wherein the emission is dominated by intermediate size dust grains in elliptical orbits about the parent planetesimal belt. This model was found to simultaneously give good agreement to the data in all bands, which is unexpected, as the more distant grains should have a lower temperature, and suggests that the mean grain size decreases with distance from the star. This is in-keeping with observational data which shows a larger disc at shorter wavelengths. Full modelling of the radial grain size distribution will be presented in Sibthorpe (2010, in prep). \\begin{enumerate} \\item We presented resolved images of the Vega debris disc system in five bands ranging from 70--500\\,$\\mu$m obtained using the \\emph{Herschel} PACS and SPIRE instruments. \\item The peak surface brightness of the dust disc was identified at 70 and 160\\,$\\mu$m at a radius of 11\\arcsec\\ (85 AU). \\item The surface brightness profile was found to be well fit in the outer disc by a $\\log_{10}(S_{\\nu}) \\propto -0.63r$ distribution, with a different scale factor at each band. The inner profile ($r \\le r_{0}$) was likewise modelled, with a Gaussian profile of FWHM = 20\\arcsec\\ found to provide a good fit. The change in surface brightness distribution, occuring at a radius of $\\sim$14\\arcsec\\ ($\\sim$109 AU), is used to observationally define the distinction between the inner and outer disc. This model was found to simultaneously give good agreement to the data in all bands. \\item The structure of the disc was found to be smooth, with no clumpy structure to the sensitivity limit of these data. \\item While these data cannot preclude the option that the Vega disc is the result of a large planetesimal collision, making it ephemeral in nature, these data support the hypothesis that the Vega disc is steady-state in nature. \\end{enumerate}" }, "1005/1005.1770_arXiv.txt": { "abstract": "One of the main tasks for present and future dark energy surveys is to determine whether the dark energy is dynamical or not. To illustrate this from data, it is commonly used to parameterize the dark energy equation of state $\\omega$ as several piecewise constant $\\omega_{i}$s using the principal component analysis (PCA) method over finite redshift bins. We show that there is only $j-1$ free parameters $\\omega_{i}$s if we choose the redshift as $j$ bins. Without this constrain, one obtains the inconsistent results from the data analysis. Furthermore, if $\\omega$ decreases with non-negligible ratio as $z$ does, then PCA fails to reproduce the original behavior of $\\omega$. Also, time varying $\\omega$ can be confused with the incorrect value of constant one when the decreasing (or increasing) ratio of $\\omega$ is small but not negligible. ", "introduction": " ", "conclusions": "" }, "1005/1005.3296_arXiv.txt": { "abstract": "We report on the results from \\halpha\\ imaging observations of the eastern limb of Tycho's supernova remnant (SN1572) using the Wide Field Planetary Camera-2 on the Hubble Space Telescope. We resolve the detailed structure of the fast, collisionless shock wave into a delicate structure of nearly edge-on filaments. We find a gradual increase of \\halpha\\ intensity just ahead of the shock front, which we interpret as emission from the thin ($\\sim1\\arcsec$) shock precursor. We find that a significant amount of the \\halpha\\ emission comes from the precursor and that this could affect the amount of temperature equilibration derived from the observed flux ratio of the broad and narrow \\halpha\\ components. The observed \\halpha\\ emission profiles are fit using simple precursor models, and we discuss the relevant parameters. We suggest that the precursor is likely due to cosmic rays and discuss the efficiency of cosmic ray acceleration at this position. ", "introduction": "\\label{sec:intro} The shock transition in fast astrophysical shocks is intrinsically a ``collisionless'' process, and energy is dissipated via plasma turbulence and/or electromagnetic fields. An important consequence of the collisionless nature of the shocks is cosmic ray acceleration \\citep[e.g.,][]{1987PhR...154....1B}. While there is increasing evidence of cosmic ray acceleration in supernova remnants, the details of the process are still not well understood, and the question of whether supernova remnants (SNRs) are the primary acceleration sites of Galactic cosmic rays is still open \\citep{2009Natur.460..701B}. Cosmic ray acceleration models require a precursor in which accelerated particles can scatter back to the postshock region for further acceleration. Observations of the cosmic ray precursor can constrain the two key parameters of acceleration models; the diffusion coefficient and the injection efficiency \\citep{1987PhR...154....1B,1988ApJ...333..198B}. The Balmer-dominated filaments that are produced when fast SNR shocks propagate into partially neutral gas are potential sites where such a cosmic ray precursors can be observed. Most of the Balmer emission comes from a very narrow zone behind the shock, where the hydrogen atoms swept up by the shock are excited before they are ionized \\citep{1978ApJ...225L..27C,1980ApJ...235..186C}. However, in the presence of the precursor, additional Balmer emission is expected from the precursor region where the preshock gas is compressed and heated. Using long-slit \\halpha\\ spectroscopy along the shock normal of a Balmer filament in Tycho's SNR, \\citet[][Lee07 hereafter]{2007ApJ...659L.133L} found that there is an increase of the \\halpha\\ narrow component intensity in a small region ($\\sim 0.4\\arcsec$) ahead of the shock front, which they proposed as potential emission from the precursor. However, the angular resolution of the observation by Lee07 is $\\sim 0.5\\arcsec$ and their results needed to be verified with high resolution observations. In this \\emph{Letter}, we report \\halpha\\ imaging observations of Balmer-dominated filaments in Tycho's SNR using the Wide Field and Planetary Camera 2 (WFPC2) on the Hubble Space Telescope (HST), which resolves the detailed structure of the shock. \\S~\\ref{sec:observations} presents the observations and reports the detection of the precursor. The precursor is modeled in \\S~\\ref{sec:model} and its characteristics are discussed in \\S~\\ref{sec:analysis}. Finally in \\S~\\ref{sec:summary}, we discuss the efficiency of cosmic ray acceleration in this region. \\begin{figure}[b] \\plotone{f1.eps} \\label{fig:kpno-hst} \\caption{ Hubble Space Telescope image of the Balmer-dominated filaments in the northeastern part of Tycho's SNR. The WFPC2 detector with F656N filter was used. The inset shows an \\halpha\\ image toward Tycho's SNR from a ground-based telescope (KPNO). The area observed by HST is marked. } \\end{figure} ", "conclusions": "The existence of the cosmic ray precursor does not necessarily imply efficient cosmic ray acceleration. For the Balmer-dominated filaments to be observable, some neutral hydrogen needs to survive ionization in the precursor. As efficient cosmic ray acceleration tends to make the precursor wider and hotter, the Balmer-dominated filaments may not trace the shocks having efficient cosmic ray acceleration \\citep[cf.][]{2009Sci...325..719H}. For the shock studied in this paper, a more direct suggestion of inefficient acceleration can be inferred from the results of Lee07. From the difference in radial velocity of the preshock gas and the \\halpha\\ narrow component emitted in the postshock region, they estimated the amount of gas deceleration in the precursor. The preshock gas is decelerated in the precursor due to the gradient of the cosmic ray pressure by about a few hundred \\kms\\ based on the radial velocity measurements. The value is not sensitive to the assumed distance, and is significantly smaller than the measured line width of the broad component. % In this shock, the thermal pressure of the ordinary gas still dominates over the cosmic ray pressure and the acceleration is not likely to be very efficient. Throughout the discussion, we have assumed that the distance to Tycho's SNR is 2.1 kpc. The measurement is based on the estimated proton temperature (from the observed line width of the broad component) and the optical proper motion, assuming no cosmic ray acceleration at the shock. The cosmic ray acceleration, if significant, can effectively reduce the postshock temperature \\citep[see][and references therein]{2009Sci...325..719H}; thus the distance may have been underestimated. However, as has been discussed above, the optical observations are consistent with cosmic ray acceleration not being very efficient in this region, so the distance of 2.1 kpc, derived assuming no cosmic ray acceleration, remains a reasonable value. Using Chandra observations, \\citet{2005ApJ...634..376W} found the locations of the shock front (SF) and the contact discontinuity (CD) along the boundary of the remnant. They interpreted the small separation between the two as an indication of efficient cosmic ray acceleration. The region of knot g is one of the regions where the SF-CD separation is smallest (except those regions of ejecta protrusion). A simple extrapolation of their argument will lead to the most efficient cosmic ray acceleration in this region, being inconsistent with our results. However, the region around knot g is where the remnant could be interacting with dense ambient clouds \\citep{1997ApJ...491..816R,2004ApJ...605L.113L}. Thus, the small SF-CD separation in this region could be due to a recent encounter of the shock with the dense ambient gas, instead of efficient cosmic ray acceleration. In conclusion, we have presented high resolution \\halpha\\ imaging observations of Tycho, revealing the existence of a thin precursor which we interpret as a cosmic ray precursor. While the current observation is consistent with inefficient acceleration, the observation of the precursor itself provides an important opportunity to constrain the key parameters of the acceleration, such as the diffusion coefficient and the injection parameters. A comparison with detailed numerical simulations will be critical to study the detailed physics of cosmic ray acceleration." }, "1005/1005.3405_arXiv.txt": { "abstract": "{{\\it Context:} The expansion of network magnetic fields with height is a fundamental property of flux tube models. A rapid expansion is required to form a magnetic canopy.\\\\ {\\it Aims:} We characterize the observed expansion properties of magnetic network elements and compare them with the thin flux tube and sheet approximations, as well as with magnetoconvection simulations. \\\\ {\\it Methods:} We used data from the Hinode SOT NFI NaD$_{1}$ channel and spectropolarimeter to study the appearance of magnetic flux concentrations seen in circular polarization as a function of position on the solar disk. We compared the observations with synthetic observables from models based on the thin flux tube approximation and magnetoconvection simulations with two different upper boundary conditions for the magnetic field (potential and vertical). \\\\ {\\it Results:} The observed circular polarization signal of magnetic flux concentrations changes from unipolar at disk center to bipolar near the limb, which implies an expanding magnetic field. The observed expansion agrees with expansion properties derived from the thin flux sheet and tube approximations. Magnetoconvection simulations with a potential field as the upper boundary condition for the magnetic field also produce bipolar features near the limb while a simulation with a vertical field boundary condition does not. \\\\", "introduction": "At present, observations do not provide an explicit picture of how the chromospheric network magnetic field is structured. On one hand, we have increasing observational evidence of something which can be interpreted loosely as a canopy structure: e.g., fibrils in the Ca II infrared triplet lines; \\citealt{Vecchio+others2007}, large-scale canopy structures in combined Zeeman and Hanle studies: \\cite{Bianda+others1998, Stenflo+others2002}, canopy-like expansion seen in magnetograms near the limb; \\cite{Jones+Giovanelli1983}). On the other hand, \\add{we know that the Sun is more complicated than implied by simple models \\citep{WedemeyerBoehm09}. An important question is to what extent simple models of flux tubes are able to reproduce the center-to-limb appearance of the network magnetic field structures.} The appearance of network magnetic flux concentrations in circular polarization maps changes from unipolar at disk center to bipolar near the limb. This is consistent with network magnetic fields expanding and fanning out with height as proposed by \\cite{Gabriel1976}. In the chromosphere, the fanning is clearly present; e.g., \\cite{Jones+Giovanelli1983} found low-lying, 200-800 km, magnetic canopies in magnetograms taken near the solar limb. More recently \\cite{Kontar+Hannah+MacKinnon2008} has used hard X-ray observations from RHESSI to estimate the expansion and found that the magnetic field expanded noticeably at a height of $\\approx$ 900 km. Expansion of magnetic field with height has been studied in photospheric structures mostly using magnetic flux tube models. Indeed, flux tube models predict a rapid expansion of the field with height. \\cite{Solanki+others1999} shows that magnetic structures as different in size and flux as small flux tubes and sunspots have similar relative expansion rates, which agree with the thin flux tube approximation. A study of the characteristics of magnetic flux structures in radiative magnetohydrodynamic (MHD) simulations revealed the expansion properties to be similar with the thin flux tube and sheet approximations (\\citealt{Yelles+others2009}). The expansion is seen in observations: e.g., a thin flux tube model can simultaneously reproduce the observed Zeeman splittings of Mg 12.32 $\\mu$m, Fe 525.0 nm and Fe 1.56 $\\mu$m lines, which span the upper to the lower photosphere in formation height \\citep{Bruls+Solanki1995}. Bruls \\& Solanki also showed that a flux tube model can explain the Mg 12.32 $\\mu$m line profile shapes observed by Zirin \\& Popp (1989). Additional evidence of expansion is that a canopy resulting from the expansion of a flux tube can best explain the observed photospheric asymmetric Stokes $V$ profiles with weak zero-crossing shifts \\citep{Grossmann-Doerth+others1988}. \\nocite{Zirin+Popp1989} In this paper we use circular polarization maps from the Solar Optical Telescope (SOT, \\citealt{Tsuneta+others2008}) on the Hinode satellite to study the \\add{average} expansion properties by characterizing how the appearance of network flux concentrations changes from the solar disk center to the limb. The center-to-limb approach lets us examine the expansion at different viewing angles and at different heights due to the shift in the height \\add{range where} spectral lines \\add{are formed} as a function of $\\mu$ ($\\mu=cos(\\theta)$, where $\\theta$ is the viewing angle). To further expand the coverage we use SOT observations from the spectropolarimeter (SP) and the narrowband filter imager (NFI) NaD$_{1}$ channel. We combine the observations with modeling the expansion of magnetic flux with height \\add{at various $\\mu$-values} by using the thin flux tube and sheet approximations and more realistic 3-dimensional magneto-convection simulations. ", "conclusions": "The SOT observations show clearly the expansion of magnetic flux concentrations with height as a change from unipolar radial cuts at disk center to bipolar near the limb. Using the ratio of the radial cuts' center- to the limb-side peak as a proxy for expansion indicates that the magnetic field has fanned out more at the formation height of the NaD$_{1}$ NFI signal than at the formation height of the Fe I 630 nm SP signal. The NaD$_{1}$ ratios are smaller than the Fe I and the switch over from unipolar to bipolar cuts takes place closer to disk center, i.e., at $\\mu$=0.5 in NaD$_{1}$ rather than at $\\mu \\approx$0.3 in Fe I. Furthermore, unlike the SP data the NFI data exhibit exhibit only very few unipolar features close to the solar limb. The apparent expansion between the SP and NFI data and the lack of unipolar features close to the limb in NFI data can be caused by several factors: the NaD$_{1}$ signal is formed higher and the features that would appear unipolar in SP look bipolar in NFI due to increased expansion of the field with height, as expected from flux tube models. Alternatively, the unipolar features are not strong enough or do not reach high enough to be visible in the NFI data. The line formation of the NaD$_{1}$ and Fe I 630 nm lines are somewhat different, e.g., the effective $g$ value of the NaD$_{1}$ line is smaller than for either Fe line making it less sensitive to low magnetic field strengths or fluxes. The difference may be accentuated by other factors such as the NFI data being filtergram while the Fe I signals are deduced from spectrograph data. The different noise levels and pixel sizes of the instruments may also play a role. To fully compare the two one should use simultaneous NFI and SP observations at various disk positions. Expansion based on the observed ratios and ratios derived from the thin flux sheet and tube approximations are in good agreement. The synthetic ratios are of similar magnitude and the switch over from unipolar to bipolar takes place at similar $\\mu$-values, i.e., around 0.5. Since the sheets expand faster, the thin sheet model results in bipolar features closer to disk center, first measurable ratios appear at $\\mu\\approx$0.8, and at smaller $r_{0}$: 200 km for the sheet and 400 km for the tube models (assuming the observations are perpendicular to the elongated direction of the sheet). Unlike in the observations, in the thin sheet/tube models the ratios between the limb-to-center side Stokes $V$ amplitudes and areas of the Fe lines are smaller than the NaD$_{1}$ ratios. While the NaD$_{1}$ line core is formed higher than the Fe line cores, the line flanks, from which the bulk of the Stokes $V$ signal is from, are formed at roughly the same height in the sheet and tube models (based on response functions computed in LTE). The difference between the synthetic and observed ratios is probably at least partly due to the radiative transfer being done in LTE and a sampling bias due to the differences in the SP and NFI data. Also, factors such as model magnetic field strength and microturbulence may influence the ratios. If the disk center-side of the flux concentration has a strong enough field it may become Zeeman saturated before the limb-side, leading to a relatively reduced center side amplitude and a decreased ratio. Since the Fe lines have a larger Zeeman sensitivity this effect would first take place in the Fe lines. The observations clearly show the expansion of the field which is found to be similar to the expansion derived from using the thin flux sheet and tube approximations. This result is in agreement with \\cite{Solanki+others1999}. In other aspects the magnetic network in the Sun is probably not well described by a simple flux tube or sheet scenario. In reality the network is dynamic and likely composed of individual flux concentrations which merge at some height. Observations of chromospheric lines, such as H-$\\alpha$ (e.g., \\citealt{DePontieu+others2007}), the Ca K (e.g., \\citealt{Zirin1974}, \\citealt{Pietarila+others2009}) and Ca II infrared triplet lines around 850 nm (e.g., \\citealt{Vecchio+others2007}), exhibit fibril structures which probably outline magnetic field lines that are more heated than others, i.e., the fibrillar canopy is thermally inhomogeneous. The magnetic field in contrast is likely quite smooth as indicated by the homogeneous appearance of both the uni- and bipolar patches in the SOT data. The choice of upper BC for the magnetic field in the MHD-simulations has a significant effect on the expansion properties of the field. The BC affects not only the upper portion of the simulation domain but effects are also seen at the $\\tau$=1 level, namely the existence of horizontal and vertical field also outside the initial magnetic strip. In the potential field BC simulations horizontal magnetic field is present everywhere and a canopy-like structure is formed at the top of the domain whereas in the vertical field BC simulation the field remains confined in the strip. \\add{We plan to address in a later paper the time evolution of the magnetic flux in the potential field BC simulation to study in detail the transport of magnetic field from the top of the domain, i.e., the canopy, to the bottom of the photosphere.} \\add{Note that with the term canopy we are considering the canopy-like expansion of the field, not the actual chromospheric canopy.} The choice of BC affects strongly the synthetic radial cuts: no apparent bipolar features are seen in radial cuts from the vertical BC simulation. In contrast, a bipolar feature begins to appear at $\\mu=0.34$ in the potential BC case. The switch over from unipolar to bipolar takes place closer to the limb and the resulting ratios are larger than those seen in the observations or in the zeroth order thin flux tube/sheet models, indicating that the field does not fan out enough. A stronger expansion in the potential field BC simulation could be achieved by, e.g., having two strips of opposite magnetic polarities in the initial setup or increasing the simulation domain size horizontally. The former would result in a canopy structure connecting the opposite polarities. However, based on the NFI ratios, the magnetic field in the mostly unipolar polar regions expands at a similar rate as in the equatorial limb regions, where mixed polarity fields are found. The latter alternative, larger simulation domain, would result in stronger fanning out because of the periodic BCs in the horizontal direction. Currently the domain size is 24 Mm which is comparable to the typical size of internetwork cells (30 Mm, \\citealt{Beckers1968}). Therefore a larger domain would not necessarily be more realistic. Placing the upper boundary higher might also result in stronger expansion \\add{since the field would have more volume to expand in}. However, the current treatment of radiative transfer in LTE would become increasingly incorrect. Finally, the field may not yet have reached a potential state at the upper boundary, so that even this boundary condition may not be appropriate. We plan to extend the current study by using spectropolarimetric observations of photospheric and chromospheric lines to study in detail the network magnetic fields. \\add{ \\subsection{Conclusions} The circular polarization signal of magnetic flux concentrations is known to change from unipolar at disk center to bipolar near the limb, consistent with the magnetic field fanning out with height. \\cite{Solanki+others1999} showed with observations then available that the relative expansion rates of features of various sizes, were consistent with rates derived for simple thin flux tubes. The data from the Hinode satellite, with their high spatial resolution, low noise and freedom from atmospheric seeing effects, provide a new opportunity to address the question with better statistics and with an emphasis on the variation from disk center to the limb. We again found the observations to be in good agreement with expansion properties derived from the thin flux sheet and tube approximations, confirming the previous results with the new data and a more extensive analysis. Since the signal we are interested in forms higher in the atmosphere near the limb, we found that realistic numerical simulations, which extend only to 700 km above $log(\\tau_{500nm})=0$, are less useful in modeling these observations, with the results strongly depending on the choice of boundary conditions for the magnetic field. }" }, "1005/1005.2709_arXiv.txt": { "abstract": "We present wide-field imaging of the 2007 outburst of Comet 17P/Holmes obtained serendipitously by SuperWASP-North on 17 nights over a 42-night period beginning on the night (2007 October 22-23) immediately prior to the outburst. Photometry of 17P's unresolved coma in SuperWASP data taken on the first night of the outburst is consistent with exponential brightening, suggesting that the rapid increase in the scattering cross-section of the coma could be largely due to the progressive fragmentation of ejected material produced on a very short timescale at the time of the initial outburst, with fragmentation timescales decreasing from $t_{frag}\\sim2\\times10^3$~s to $t_{frag}\\sim1\\times10^3$~s over our observing period. Analysis of the expansion of 17P's coma reveals a velocity gradient suggesting that the outer coma was dominated by material ejected in an instantaneous, explosive manner. We find an expansion velocity at the edge of the dust coma of $v_{exp}=0.55\\pm0.02$~km~s$^{-1}$ and a likely outburst date of $t_0=2007~{\\rm October}~23.3\\pm0.3$, consistent with our finding that the comet remained below SuperWASP's detection limit of $m_V\\sim15$~mag until at least 2007 October 23.3. Modelling of 17P's gas coma indicates that its outer edge, which was observed to extend past the outer dust coma, is best explained with a single pulse of gas production, consistent with our conclusions concerning the production of the outer dust coma. ", "introduction": "Discovered on 1892 November 9 by Edwin Holmes, Comet 17P/Holmes (hereafter, 17P) is a dynamically ordinary Jupiter-family comet, with a semimajor axis of $a=3.617$~AU, an eccentricity of $e=0.432$, an inclination of $i=19.11^{\\circ}$, an orbital period of $P_{orb}=6.8$~years, and a Tisserand parameter (with respect to Jupiter) of $T_J=2.859$. On 2007 October 24.1, it was discovered by J. A. Henr\\'iquez Santana to be undergoing a substantial outburst. Starting from an initially observed magnitude of $m_V\\sim8.4$ \\citep[compared to a pre-outburst brightness of $m_V\\sim17$ measured on 2007 October 23.1;][]{cas07}, it brightened by approximately 0.5 mag~hr$^{-1}$ over the course of 6 hours \\citep{buz07}. The comet reached $m_V\\sim2.0$ by 2007 October 25.1 \\citep{spo07,elh10}, representing a million-fold increase in brightness over only 2 days. This outburst of 17P mirrored a similar outburst that led to the comet's initial discovery in 1892, and continued to be monitored by both amateur and professional astronomers as it progressed. Spectroscopic analyses indicated that the coma was largely dominated by dust grains \\citep{sch07}, but evidence of cometary volatiles and sublimation byproducts, including OH, H$_2$O, NH, CN, HCN, HNC, C$_2$, C$_3$, C$_2$H$_2$, C$_2$H$_6$, CS, CH$_3$CN, and CH$_3$OH, were also detected during the outburst \\citep{dra07,fit07,kob07,sch07,wag07,boc08,del08,yan09}. Negative polarisation was found for the dust coma by \\citet{jos10}, consistent with measurements made for other comets. A delay in the appearance of CN, C$_2$, and [OI] emission lines between 2007 October 24.58 and October 25.46 led \\citet{kob07} to speculate that their observations could be at least partially explained by the delayed sublimation of icy grains ejected at the time of the outburst. This hypothesis is supported by \\citet{del08} who found evidence of a distributed source of sublimation products, consistent with the ejection and subsequent delayed sublimation of icy grains in the coma. The total mass loss due to this outburst has been estimated to be $\\sim10^{12}$ kg or a few percent of the comet's total mass \\citep{sek08a,mon08}. ", "conclusions": "\\subsection{Primary Findings} Wide-field imaging of the 2007 outburst of 17P/Holmes was serendipitously obtained by the SuperWASP-North facility on 17 nights over a 42-night period beginning on the night (2007 October 22-23) immediately prior to the outburst. We report the following key findings: \\begin{itemize} \\item{The comet was not detected in data from the night before the outburst's discovery on the night of 2007 October 22-23, indicating that it remained below the SuperWASP detection threshold of $m_V\\sim15$~mag (consistent with pre-outburst reports placing the comet's brightness at $m_V\\sim17$~mag) until at least 2007 October 23.3. } \\item{The unresolved coma (as seen by SuperWASP) was likely optically thin during our observations on the first night of the outburst. The comet's lightcurve during these observations is consistent with an exponential function, suggesting that its rapid brightening could have been driven by the progressive fragmentation of ejected material produced on a very short timescale at the time of the initial outburst. Our best-fit functions to the data imply an initial fragmentation timescale of $t_{frag}\\sim2\\times10^3$~s, decreasing to $t_{frag}\\sim1\\times10^3$~s, and a near-instantaneous leap in brightness from a magnitude of $m_V\\sim17$~mag to $m_V\\sim11$~mag at the moment of the initial outburst. } \\item{Analysis of coma surface brightness profiles reveals a velocity gradient consistent with the outer coma being dominated by material ejected in an explosive manner (i.e., at a single instant), rather than over an extended period of time. Near the outer edge of the visible coma, where the coma likely transitions from being dust-dominated to gas-dominated, this velocity gradient corresponds to an expansion velocity of $0.55\\pm0.02$~km~s$^{-1}$, consistent with previous reported measurements. From this analysis, we find a most likely outburst date of $t_0=2007~{\\rm October}~23.3\\pm0.3$. This finding of explosively ejected material dominating the outer coma does not rule out the possibility of subsequent sustained mass ejection supplying the inner coma at later times. } \\item{We measure the rate of motion (relative to the nucleus) of the secondary brightness peak in 17P's coma, which we refer to as the ``false nucleus'', and find a projected drift velocity of $v_{drift}=120\\pm5$~m~s$^{-1}$, consistent with previously reported measurements. } \\item{Dust modelling shows that 17P's nucleus is a plausible primary emission source of outer coma material, and that a secondary source such as a separated nucleus fragment is not required to explain the motion of the coma relative to the nucleus. We show instead that the drifting of the coma relative to the nucleus can be explained as the consequence of radiation pressure alone. } \\item{Modelling of 17P's gas coma indicates that the morphology of the portion of the observed coma profile hypothesised to be gas extending past the outer dust coma is best explained by a single, instantaneous outburst of gas production, rather than extended gas production that persists over several days. This result is consistent with our conclusion that the outer dust coma is likely dominated by material ejected instantaneously, and not over an extended period of time. We also note that C$_2$ is likely to be the dominant observed component of the gas coma, and find that its decay time is consistent with SuperWASP observations of the disappearance of the observed coma profile region attributed to gas. } \\end{itemize} \\subsection{Anatomy of an Outburst} To date, no consensus has emerged to explain the physical origin of 17P's spectacular outbursts in 1892 and 2007. A primary difficulty lies in reconciling the occurrence of such an apparently catastrophic event at least twice in the recent past with a lack of comparable outbursts in almost all other comets, although \\citet{sek08b} has noted that 1P/Halley exhibited a similarly powerful outburst in 1836. \\citet{whi84} suggested that a grazing encounter and eventual impact by a small satellite with the nucleus could have been responsible for the 1892 outburst, but such an explanation was rendered highly implausible with a second episode of outburst activity in 2007. Furthermore, any model formulated to explain 17P's behaviour must also account for the lack of outburst activity in the 115 years between 1892 and 2007 \\citep{sek09b}. \\citet{sek08a,sek08b} has proposed a scenario in which an inwardly-diffusing thermal wave gradually penetrated a large, weakly-cemented pancake-shaped layer of the nucleus \\citep[analogous to those observed for 9P/Tempel 1; {\\it cf}.][]{tho07} over numerous orbits until it finally reached a large reservoir of amorphous water ice at the layer's base. Upon being heated to the necessary transition temperature, this amorphous ice layer underwent an exothermic transformation to crystalline ice, leading to rapid sublimation and causing the pancake-shaped layer to separate from the nucleus and almost immediately explosively disintegrate. This two-part outburst scenario is invoked by Sekanina to explain the appearance of the false nucleus as the central source of 17P's dust coma, as well as its apparent motion relative to the true nucleus. We note, however, that our dust modelling (\\S\\ref{dustmodel}) shows that it is not necessary for the false nucleus to be the source of emission to explain the position of the outer dust coma relative to the true nucleus. Thus a model where a large surface fragment first violently separates from the nucleus (with a projected velocity exceeding 100~m~s$^{-1}$) and then explosively disintegrates, requiring two non-simultaneous catastrophic events, may be needlessly complicated. The false nucleus could still be a concentration of material related to a large fragment ejected in the initial outburst that then disintegrated, but may be better characterised as a byproduct, rather than a key component, of the outburst. \\citet{alt09} propose an alternate hypothesis in which a particularly close perihelion passage by 17P on 2007 May 4 substantially raised the sublimation rate of subsurface water ice relative to previous perihelion passages. This increased production of subsurface water vapour, coupled with an ``airtight'' layer of surface regolith, ultimately led to an explosive disintegration of 17P's dust mantle 172 days later, causing the observed outburst. While it is true that 17P's perihelion passage just 2.05~AU from the Sun was its closest approach to the Sun in over a century (according to JPL's Horizons ephemeris generator at http://ssd.jpl.nasa.gov/horizons.cgi), we note that 17P's perihelion distance immediately prior to its 1892 outburst was a more modest 2.14~AU from the Sun, a distance comparable to perihelion passage distances in 1864, 1871, 1878, 1885, 1899, 1906, 1972, 1979, 1986, 1993, and 2000, all of which were reached without any recorded reports of behaviour comparable to the comet's 1892 or 2007 outbursts. Most recently, \\citet{rea10} propose a scenario in which trapped gases released by the crystallisation of amorphous ice, as well as sublimation of other ices driven by the exothermic crystallisation process, cause a buildup of gas pressure in a subsurface cavity on the comet. The outburst of 17P then occurred when the pressure built up within this cavity exceeded the strength of the surrounding material (found to be relatively high --- $10-100$~kPa --- based on the nucleus's survival of the outburst), causing the cavity to rupture violently and energetically, creating the observed explosive ejection of nuclear material. We note that while \\citet{kos10} report that the crystallisation of amorphous water ice is unlikely to have caused 17P's outburst, they reach this conclusion based on a simple model in which the ice being crystallised is present in a sub-surface layer just below a dust mantle. The hypothesis presented by Reach {\\it et al.} is based on a different physical scenario where ice crystallisation occurs in a subsurface void in which gas pressure is able to build up, and as such, we believe that it remains plausible and, of the hypotheses proposed thus far, is the most consistent with our own findings. We note, however, that it then raises questions as to why 17P is the only comet known thus far with the apparently unique combination of high-tensile-strength nuclear material, subsurface cavities, and amorphous ice capable of driving two of the largest cometary outbursts observed in modern astronomy. \\citet{gai07} reported that observations they made between 2007 October 24 to 2007 November 4 with the Pic du Midi 1~m telescope show multiple dust streams with well-defined origin points, four of which were measured to recede from 17P's nucleus at roughly constant velocities (as projected on the sky) ranging from 50 to 100~m~s$^{-1}$, implying the presence of unresolved, steadily disintegrating nucleus fragments in the coma. These fragments were also calculated to have separated from the nucleus between 2007 October 23.7 and 2007 October 24.8. Additionally, \\citet{ste09} observed at least sixteen 10-m- to 100-m-scale fragments receding from the nucleus that showed evidence of ongoing sublimation and disintegration. Coupled with our finding of dust fragmentation in the first hours after outburst, these observations point to a scenario of continued fragmentation of cometary material following the outburst. We certainly still have far from a complete picture of 17P's 2007 outburst. We suggest that detailed dust modelling and analysis of images of the comet's inner coma using data with higher spatial resolution than SuperWASP would be useful for clarifying the temporal and kinematic nature of the outburst, in particular whether significant dust production continued from the nucleus after the initial outburst event and, if so, whether ejection velocities were comparable to those in the initial outburst event. More detailed discussions of spectroscopy would also help constrain the spatial and temporal nature of the contribution of gaseous species to the 17P's appearance, such as whether sublimation of nucleus-bound ices or ice particles in the coma was significant, and whether there was any appreciable delay in the sublimation of icy material in the coma that could be linked to hierarchical fragmentation of macroscopic nucleus particles." }, "1005/1005.5259_arXiv.txt": { "abstract": "This paper considers magnetic field generation by a fluid flow in a system referred to as the Archontis dynamo: a steady nonlinear magnetohydrodynamic (MHD) state is driven by a prescribed body force. The field and flow become almost equal and dissipation is concentrated in cigar-like structures centred on straight-line separatrices. Numerical scaling laws for energy and dissipation are given that extend previous calculations to smaller diffusivities. The symmetries of the dynamo are set out, together with their implications for the structure of field and flow along the separatrices. The scaling of the cigar-like dissipative regions, as the square root of the diffusivities, is explained by approximations near the separatrices. Rigorous results on the existence and smoothness of solutions to the steady, forced MHD equations are given. ", "introduction": "Much is known about fast dynamo action: the rapid growth of magnetic fields at high magnetic Reynolds number in fluid flows with chaotic streamlines, but the mechanisms for the dynamical saturation of such fields remain poorly understood. In many cases when the growing field equilibrates by modifying the fluid motion, the effect is to switch off the chaotic stretching in the flow, as measured for example by a reduction in the finite-time Liapunov exponents \\cite[e.g.,][]{CaHuKi96,ZiPoPo98}. What is left is a fluid threaded by a magnetic field which resists stretching and so suppresses overturning fluid motions, but supports elastic wave-like motions, essentially Alfv\\'en waves with coupled field and flow \\cite[e.g.,][]{CoHuPr10}. The final state of many simulations shows apparently chaotic behaviour in space and time, suggestive of an attractor of moderate or high dimension, although because of the three-dimensionality of MHD systems little can be done to explore its properties, for example the fractal dimension or spectrum of Liapunov exponents. Although this appears to be the outcome of many simulations, as far as they can be run, there are some intriguing examples where a further phase of evolution takes place: the magnetic field and flow align, depleting the nonlinear terms, and both fields evolve to a steady (or very slowly evolving) state. The key point is that in unforced, ideal magnetohydrodynamics (see equations (\\ref{eqNS0}--\\ref{eqdivfree0}) below with $\\nu=\\eta=0$ and $\\fv=0$) any state with $\\uv = \\pm \\bv$ is an exact steady solution. The remarkable fact that simulations of forced, non-ideal MHD turbulence could evolve to something very close to such a state was first observed by \\cite{Ar00} in his thesis, and published in \\cite{DoAr04} (hence referred to as DA), and \\cite{ArDoNo07}. These simulations use a compressible code with a Kolmogorov forcing function, (\\ref{eqSvdef}) below, first used as the form of a flow for simulations of fast, kinematic dynamo action by \\cite{GaPr92}. Subsequently \\cite{CaGa06a} undertook incompressible simulations of the same system as Archontis, and pushed up the fluid and magnetic Reynolds numbers; our work is linked closely to this paper, which we refer to as CG in what follows. What these authors found was that, starting with a forced fluid flow and a seed magnetic field, the growing magnetic field initially equilibrates in rough equipartition with the velocity field, in a messy, chaotic time-dependent state. However during this state, there is a slow but persistent exponential growth in the average alignment of the $\\uv$ and $\\bv$ vectors, as measured by the cross-helicity. This process of alignment continues until there takes place a sudden increase in the fluid and magnetic energies, and both fields tend to a steady state of almost perfect alignment, discrepancies being controlled by the weak dissipation and the forcing. In fact since any solution $\\uv=\\pm\\bv$ is a neutrally stable solution of the ideal problem \\citep{FrVi95}, the solution that is selected must depend delicately on balances involving these subdominant diffusive and forcing effects. We note that some alignment of field and flow has been noted in many other MHD flows, for example see \\cite{DoMaVe80}, \\cite{PoMeFr86}, \\cite{MaCaBo06} and references therein, but of a less dramatic nature. This observation of dynamo saturation in a steady state with such a high degree of alignment was a new phenomenon: CG refer to the saturated state as the `Archontis dynamo', though we prefer the term `Archontis saturation mechanism'. CG observed this aligned state as a solution branch over a wide range of magnetic and fluid Reynolds numbers (taking the magnetic Prandtl number to be unity in much of their work). Further developments include the development of bursts of rapid time dependence after some time in the steady state, in the study \\cite{ArDoNo07}. However this appears only to occur in the compressible case, as it has not been seen by CG nor in our simulations; we will therefore not discuss this further. \\cite{CaGa06b} also find slow time-dependent evolution of the saturated state for the Kolmogorov forcing with magnetic Prandtl number $\\Pra = \\nu/\\eta$ not equal to unity, and for more general spatially periodic steady forcings. In all cases though, the field and flow settle into a state of very close alignment, even if they then evolve on a slow time scale. The focus of the present paper is to understand more about the structure of the steady saturated state for the Kolmogorov forcing and unit magnetic Prandtl number $\\Pra$, with a particular focus on the regions where dissipation occurs and on rigorous results on existence and smoothness. DA and CG find a complex geometrical picture for the field and flow and identify these regions of high dissipation: they are localised along straight-line separatrices that join a family of stagnation points; similar structures are found in the 1:1:1 ABC flow \\citep{DoFrGrHeMeSo86}. These are found to have a width scaling as $\\sqrt{\\eps}$ where $\\eps$ is a dimensionless measure of the diffusivity, and one of our aims is to understand this power law. We set up the governing equations in \\S\\ref{secgov} and extend the solution branch to yet smaller values of the diffusivity $\\eps$ by means of large scale simulations in \\S\\ref{secnum}. In \\S\\ref{secsym} we then classify the symmetries of the Kolmogorov forcing, which are preserved by the nonlinear, saturated field and flow. These symmetries are the reason for the presence of the non-generic straight line separatrices that join stagnation points in the flow and field, and they constrain the local flow: it is in these regions that dissipation is strongest. We plot the local structure of fields along the separatrix from $(0,0,0)$ to $(\\pi,\\pi,\\pi)$ in \\S\\ref{secflowfield}. We determine the effects of diffusion by setting up PDEs for the advection of field as it enters the dissipative regions in \\S\\ref{secPDE} and use these to justify the order $\\sqrt{\\eps}$ scaling for the cigar widths found in CG. We then proceed with a formal mathematical investigation of the existence of steady-state solutions to the MHD problem at hand and bounds for them in various function spaces in \\S\\S \\ref{maths}--\\ref{smooth}. The reader should note that these sections use functional analysis and so have a different flavour from the earlier ones. Finally \\S\\ref{secdisc} offers concluding discussion. ", "conclusions": "\\label{secdisc} We have presented investigations into the structure of the magnetic field and flow in the equilibrated regime of the Archontis dynamo. Because of the highly three-dimensional nature of the system, application of the available analytical tools yields only rough results of limited value, and we lack any kind of complete solution. What we have done is first to extend the range of diffusivities $\\eps$ over which the saturation mechanism operates to give the steady state with nearly aligned fields. We have also classified the symmetries of these flows and measured the field structure on the separatrices, home of the cigar-like dissipative regions. Then, using basic analytical tools, we have investigated the scaling of diffusive terms near the separatrices. Here at leading order the field $\\lamv=\\eps^{-1}\\Lamv_-$ that enters from the body of the flow is transported along characteristics of $\\Lamv = \\Lamv_+$. Where these characteristics come together in the compressive flow at the stagnation points, where trajectories spiral in, large gradients in $\\lamv$ are generated, and diffusive terms enter the problem on scales of $\\eps^{1/2}$ as found by CG. In more general flows we may expect a similar behaviour, with regions of heightened dissipation localised at points where $\\Lamv=0$ and along the unstable manifolds of such points. Of course in the Archontis example the unstable manifolds link the stagnation points and so the topology here is very simple and the dissipative regions very small, of order $O(\\eps)$ in volume: in other cases they may wander through the three-dimensional space, giving a picture of much greater complexity, as could be occurring in examples in \\cite{CaGa06b}. Again wider regions of dissipation, perhaps dense in the space, could occur if examples exist where $\\Lamv$ has no stagnation points; unfortunately the form of $\\Lamv$ is not under our control except where strongly constrained by symmetries. In order to cope with unknown levels of geometrical complexity, an approach based on functional analysis is appropriate, and this is the final part of the paper, in which the existence and smoothness properties of steady solutions are established. An analogy of the naively truncated equations (namely (\\ref{eqLamv}, \\ref{eqlamv}) with $\\eps=0$) with the Euler equation, which is the subject of intense research, is instructive. A method for the investigation of the evolutionary Euler and Navier--Stokes equations consists of the introduction into the equations of new regularising terms, such as $\\eps(-\\nabla^2)^\\alpha\\uv$ or $\\eps(-\\nabla^2)^\\beta(\\partial\\uv/\\partial t)$. It has been known for decades that for $\\alpha>5/4$ solutions to the regularised equations are infinitely differentiable at any $t>0$; for $\\beta\\geq1 /2$ and $\\beta > 5/6$ one can prove analyticity, at any $t>0$, of solutions to the Navier--Stokes and Euler equations, respectively \\citep{Zh10}. For any $\\alpha$ or $\\beta$ below the respective thresholds, the problem is as difficult as the one for the original equation. When the limit $\\eps\\to0$ is considered, the results are so far inconclusive. One can only show that there exist sequences $\\eps_k\\to0$ such that solutions for these $\\eps_k$ converge to a weak solution to the non-regularised equation, and either the limit weak solution is unique for all such sequences, or there exists a continuum of weak solutions. Whether for $\\eps\\to0$ singularities develop in derivatives of the regularised solutions, and how strong they are if they develop, remains unknown. The difficulties arise in the general theory, because the bounds for solutions are singular in $\\eps$ as $\\eps\\to0$. Here the analogy with the Archontis dynamo problem crystallises: in the Archontis problem the diffusive terms can be regarded as a regularisation of the naively truncated diffusionless problem, and we need to find out what happens when the regularisation parameter $\\eps$ tends to zero. (In the diffusionless, i.e.\\ non-regularised, case it is unclear whether weak steady solutions exist.) We note that the analogy may work both ways: the asymptotic analysis near the separatrix in the Archontis dynamo (which we present in \\S\\S \\ref{secflowfield} and \\ref{secPDE}) may contain clues to what happens in solutions to the regularised Euler (or even Navier--Stokes) equations in the limit $\\eps\\to0$. Unfortunately, the clues are well hidden, because the regularising term in the Archontis problem is of a different structure, and a very specific symmetric steady solution to the general system of MHD equations is considered. Besides further attempts to carry out an asymptotic analysis of equations \\rf{eqLamv} and \\rf{eqlamv} and their evolutionary versions, a number of other directions could be pursued in the future, for example investigating time-dependent modifications to the steady Kolmogorov forcing used here, or studying the evolution of superposed large-scale fields and corresponding non-helical transport effects, as in the recent work of \\cite{SuBr09}." }, "1005/1005.3155_arXiv.txt": { "abstract": "We investigate the far-infrared--radio correlation (FRC) of stellar-mass-selected galaxies in the Extended Chandra Deep Field South using far-infrared imaging from \\textit{Spitzer} and radio imaging from the {Very Large Array} and {Giant Metre-Wave Radio Telescope}. We stack in redshift bins to probe galaxies below the noise and confusion limits. Radio fluxes are $K$--corrected using observed flux ratios, leading to tentative evidence for an evolution in spectral index. We compare spectral energy distribution (SED) templates of local galaxies for $K$--correcting FIR fluxes, and show that the data are best fit by a quiescent spiral template (M51) rather than a warm starburst (M82) or ULIRG (Arp220), implying a predominance of cold dust in massive galaxies at high redshift. In contrast we measure total infrared luminosities that are consistent with high star-formation rates. We observe that the FRC index ($q$) does not evolve significantly over $z=0-2$ when computed from $K$-corrected 24 or 160-\\mum\\ photometry, but that using 70-\\mum\\ fluxes leads to an apparent decline in $q$ beyond $z\\sim1$. This suggests some change in the SED at high redshift, either a steepening of the spectrum at rest-frame $\\sim25-35$\\mum\\ or a deficiency at $\\sim70$\\mum\\ leading to a drop in the total infrared/radio ratios. We compare our results to other work in the literature and find synergies with recent findings {on the high-redshift FRC, high specific star-formation rates of massive galaxies and the cold dust temperatures in these galaxies}. ", "introduction": "One of the most exciting research areas in observational astronomy at this time is the field of far-infrared astronomy. Our understanding of extragalactic sources in the far-infrared (FIR) and sub-millimetre (sub-mm) to millimetre regimes has improved exceptionally over the past decade, thanks to such instruments as ESA's \\textit{Infrared Space Observatory} (\\textit{ISO}, launched 1995; \\citealp{kessler_infrared_1996}) and NASA's \\textit{Spitzer Space Telescope} (launched 2003; \\citealp{werner_spitzer_2004}), alongside ground-based instruments such as the Sub-millimetre Common-User Bolometer Array (SCUBA; \\citealp{holland_scuba:common-user_1999}) and the Max Planck Millimeter Bolometer Array (MAMBO; \\citealp{kreysa_bolometer_1998}), both commissioned in the late 1990's. One of the most important FIR instruments to date is the Multiband Imaging Photometer for \\textit{Spitzer} \\citep[MIPS;][]{rieke_multiband_2004}, which, with \\textit{Spitzer}'s 0.85m mirror provides imaging with diffraction limited resolutions of 6, 18 and 40 arcsec in three broad bands centred at 24, 70 and 160\\mum\\ respectively. Recent advances in the resolution of the cosmic infrared background (CIB) by stacking into images from \\textit{Spitzer} {\\citep[see e.g.][]{dole_cosmic_2006,dye_scuba/spitzer_2007,chary_new_2010}} and BLAST \\citep[The Balloon--borne Large Aperture Sub-mm Telescope;][]{marsden_blast:_2009} have enabled an improved understanding of the history of star formation and galaxy formation and evolution, and this will be further improved by ongoing work \\citep[such as][]{berta_dissecting_2010} with ESA's new \\textit{Herschel Space Observatory} \\citep{pilbratt_herschel_2010}. Most of the stellar mass in the local universe is concentrated in the most massive galaxies ($M_\\star \\gtrsim 10^{11}\\text{M}_\\odot$; \\citealp{kauffmann_stellar_2003}) and observations show that these have been in place since $z\\sim 1$ \\citep{dickinson_evolution_2003, bundy_mass_2005, prez-gonzlez_stellar_2008, taylor_rise_2009, collins_early_2009}. The significant increase in density of luminous ($L_{8-1000\\mum} \\gtrsim 10^{11} \\text{L}_\\odot$) and ultra-luminous ($L_{8-1000\\mum} \\gtrsim 10^{12} \\text{L}_\\odot$) infrared galaxies (LIRGS and ULIRGs) from the local universe to $z\\sim 2-3$ \\citep[e.g.][]{daddi_passively_2005, daddi_population_2005, caputi_role_2006} is thought to reveal the formation stages of these latter-day giants, which apparently formed in a remarkably short time between $1 \\lesssim z \\lesssim 3$, in an antithesis to the paradigm of hierarchical structure formation \\citep[e.g.][]{de_lucia_formation_2006}. \\label{sec:sources} In the FIR--sub-mm regime the dominant source of the extragalactic background light (after the cosmic microwave background) is thermal continuum emission from interstellar dust, which is mainly composed of polycyclic aromatic hydrocarbons (PAHs), graphites and silicates, typically less than $\\sim 0.25\\mum$\\ in size \\citep{draine_infrared_2007, draine_dust_2007}. The FIR emission in star-forming galaxies is thought to arise both from cold dust in the large-scale `cirrus' component of the interstellar medium (ISM), and from warmer dust in and around star-forming regions \\citep[e.g.][]{de_jong_iras_1984,helou_iras_1986}. If there is a sufficient level of dust-enshrouded star formation then the FIR emission is dominated by this `warm' dust component, which has characteristic temperatures of around 30--50K \\citep{dunne_scuba_2001, sajina_1-1000m_2006, dye_scuba/spitzer_2007, pascale_blast:far-infrared_2009}. The dust is heated to these temperatures by the ultraviolet (UV) radiation field from hot O and B type stars, which are present only in regions of ongoing or recent star formation (`recent' meaning within the lifetime of these short-lived stars, $\\lesssim10^{8}\\text{yr}$; \\citealp{kennicutt_star_1998}). For this reason, the total IR luminosity ($L_\\text{TIR} = L_{8-1000\\mum}$) can be used as a tracer of star-formation rates (SFRs) in galaxies, often using one or a number of FIR fluxes (such as the \\textit{Spitzer} MIPS bands) or sub-mm fluxes to estimate $L_\\text{TIR}$ \\citep{kennicutt_star_1998}. One issue with using the FIR as a SFR tracer is the contribution from cold dust in the ISM, which is heated by older stars in the disk of the galaxy, and is therefore unrelated to star formation \\citep{calzetti_calibration_2010}. Contamination of samples by galaxies hosting active galactic nuclei (AGN) is another problem, as AGN also emit UV radiation which can heat dust in the torus. AGN-heated dust is generally hotter than dust heated in star-forming regions, so the thermal spectrum peaks at a shorter wavelength, and mid-infrared (MIR) fluxes (including \\textit{Spitzer}'s 24-\\mum\\ band, as well as shorter wavelengths) would be boosted. Mid-infrared fluxes are also affected more uncertainly by the emission features of PAH molecules, which are ubiquitous in star-forming galaxies \\citep[e.g.][]{leger_identification_1984, roche_atlas_1991, lutz_nature_1998, allamandola_modeling_1999}, and the 10-\\mum\\ silicate absorption trough, so the use of 24-\\mum\\ fluxes as SFR indicators at high redshifts is subject to some contention \\citep[see e.g.][]{dale_infrared_2005, calzetti_calibration_2007, daddi_multiwavelength_2007, papovich_spitzer_2007, young_mid-_2009, rieke_determining_2009}. \\label{sec:FIRRC} Another part of a galaxy's SED that can be used as a SFR indicator is the radio luminosity. Non-thermal radio continuum emission from star-forming galaxies originates from type II supernova remnants (SNRs), the endpoints of the same massive short-lived stars that heat the dust via their UV radiation. This connection with dust heating is important because it leads to the well-known (but not fully understood) FIR--Radio correlation \\citep[FRC;][etc]{van_der_kruit_high-resolution_1973, rickard_far-infrared_1984, helou_thermal_1985, condon_radio_1992}. The FRC is linear, remarkably tight and holds for a wide range of galaxy types over at least five orders of magnitude in luminosity \\citep{yun_radio_2001}. It can be explained in terms of ongoing star~formation producing hot massive ($M>8\\text{M}_\\odot$) stars: while the FIR flux is emitted from dust heated by these stars, the radio emission arises from synchrotron radiation by cosmic ray (CR) electrons accelerated in the SNRs of the dying stars. The non-thermal radio emission is smeared out through the galaxy as the relativistic CR electrons travel through the galaxy over lifetimes of $\\tau \\sim 10^{8}$ years, during which they emit synchrotron radiation via interactions with the galactic magnetic field {(as described by \\citealt{condon_radio_1992}, and shown observationally by \\citealt{murphy_effect_2006})}. A shallower thermal component is also present in the radio spectrum due to bremsstrahlung radiation from electrons in H\\,{\\sc ii} regions, but this becomes dominant only at high frequencies ($\\gtrsim 30$~GHz) and at 1.4~GHz only comprises $\\sim 10\\%$ of the radio flux \\citep{condon_radio_1992}. It is difficult however to explain the linearity and tightness of the correlation between thermal FIR luminosity and non-thermal radio luminosity. `Minimum energy' estimates of magnetic fields in galaxies \\citep{burbidge_synchrotron_1956}\\footnote{ Inferring magnetic field strengths of galaxies is problematic and depends on many unknowns. The minimum energy argument circumvents some of these problems by specifying the minimum of the total energy density (a function of field strength), which occurs at the point where the energy density in particles is in approximate equipartition with that in the magnetic field \\citep{longair_high_1994}. Although there is no physical requirement for this equipartition to occur, it does provide an order-of-magnitude estimate for the magnetic field energy density, and is physically motivated if one considers that the magnetic field is tangled by turbulent motions in the interstellar plasma, and that these motions efficiently accelerate cosmic ray particles (see \\citealt{longair_high_1994} and \\citealt{thompson_magnetic_2006}). Equipartition has been observationally confirmed locally in the Milky Way \\citep{strong_diffuse_2000} and in normal spiral galaxies \\citep{vallee_galactic_1995,beck_magnetic_2000}, although its validity in starbursts and ULIRGs is less certain \\citep{thompson_magnetic_2006}.} imply a large variation between normal galaxies and extreme starbursts like Arp220. To explain a constant FIR/radio ratio between such disparate systems, complex physical solutions need to be provided, for example invoking strong fine-tuning to regulate electron escape and cooling timescales, or short cooling timescales with magnetic fields $\\sim10$ times stronger than the `minimum energy' argument suggests \\citep{thompson_magnetic_2006}. \\Citet{voelk_correlation_1989} first suggested a `calorimeter' model whereby both UV light from massive stars and CR electrons from SNRs are (a) proportional to the supernova rate, and (b) efficiently absorbed and reprocessed within the galaxy, so that the respective energy outputs in FIR re-radiation and radio synchrotron would both be tied to the supernova rate. This theory requires a correlation between the average energy density of the radiation field and the galaxy magnetic field energy density. \\Citet{voelk_correlation_1989} argues this is plausible if the origin of the magnetic field is a turbulent dynamo effect, since the turbulence would be largely caused by the activity of massive stars, and hence correlated with the supernova rate. Alternative non-calorimetric models include those of \\citet{helou_physical_1993}, using a correlation between disk scale height and the escape scale length for CR electrons; and \\citet{niklas_new_1997}, in which the FRC is driven by correlations with the overall gas density and equipartition of magnetic field and CR energy. \\Citet{bell_estimating_2003} argues for a `conspiracy' to diminish both the FIR and radio emission originating from star formation in low luminosity galaxies when compared with luminous $\\sim L_\\star$ galaxies -- without this the relationship would not remain linear over the full luminosity range. Similarly, the calorimeter model of \\citet{lacki_physics_2009-1} invokes conspiracies in low and high gas surface density regions to maintain the relationship. The physical origin of the FRC therefore is still an open question. A full review of the theories is beyond the scope of this study, but a more detailed discussion of the literature can be found in \\citet{vlahakis_far-infrared-radio_2007}, and a more in-depth treatment is provided {in the numerical work of} \\citet{lacki_physics_2009-1}. An ongoing strand of research at the current time is the investigation of the FRC at high redshifts and low fluxes, in particular whether there is any evolution \\citep[e.g.][]{ appleton_far-_2004, frayer_spitzer_2006, ibar_exploringinfrared/radio_2008, garn_relationship_2009, seymour_investigatingfar-ir/radio_2009, ivison_blast:far-infrared/radio_2009,sargent_vla-cosmos_2010,sargent_no_evolution_2010}. Measurements of any evolution (or lack thereof) would improve the accuracy of FIR-/radio-estimated SFRs at high redshift, and could shed light on the mechanism governing the FRC, as well as highlighting differences in the physical and chemical properties of star-forming galaxies at high and low redshift \\citep{seymour_investigatingfar-ir/radio_2009}. In the current work we investigate the FRC over a range of redshifts, for a sample that is not limited by FIR or radio flux. Using \\textit{Spitzer} FIR data and radio data from the {Very Large Array} ({VLA}) and {Giant Metre-Wave Radio Telescope} ({GMRT}), we quantify the FIR--Radio Correlation as a function of redshift in massive galaxies selected from a near-infrared (NIR) survey of the Extended Chandra Deep Field South (ECDFS). We use Equation~\\ref{eqn:q_ir} to define the `$q$' index, which quantifies the FRC as the logarithmic ratio between a monochromatic FIR flux ($S_{\\nu,\\text{IR}}$, e.g. at 24, 70 or 160\\mum), and 1.4-GHz radio flux ($S_{\\nu,\\text{1.4~GHz}}$). \\begin{equation} \\label{eqn:q_ir} q_\\text{IR} = \\log_{10}\\left(\\frac{S_{\\nu,\\text{IR}}}{S_{\\nu,\\text{1.4~GHz}}}\\right) \\end{equation} We also investigate the effects of using different FIR bands to quantify the FRC, and the effects of assumptions about the SEDs of the galaxies in the sample. We employ a `stacking' methodology to recover sufficient signal-to-noise ratios on faint objects to obtain measurements of the average properties of the sample. The data are described in Section~\\ref{sec:data}, while the binning and stacking methodologies are described in Section~\\ref{sec:method}. The analysis of SEDs and application of $K$--corrections is covered in Section~\\ref{sec:kcorrs}, and the results are analysed and discussed in Section~\\ref{sec:discussion}. A concordance cosmology of $\\Omega_m=0.27$, $\\Omega_\\Lambda=0.73$, $H_0=71$\\ kms$^{-1}$Mpc$^{-1}$ is assumed throughout. ", "conclusions": "We have studied the FRC as a function of redshift for NIR-selected massive galaxies in the ECDFS, a sample which is unbiased by star-formation activity. We used a stacking analysis to evaluate the ratios of median FIR/radio fluxes of all galaxies in the sample, divided into redshift bins. This technique traces the typical objects in the population of massive galaxies from low redshift back to their formation epoch. A thorough analysis of clustering of the sample was used to correct for the differential effects of confusion in the three FIR bands. $K$--corrections were derived in the radio and FIR using ratios of observed fluxes, ensuring as much as possible a self-consistent analysis. A mass-limited sub-sample was also stacked to confirm the robustness of the results to Malmquist bias. The results for $q_{24}$, $q_{70}$ and $q_{160}$ show a slight decline in the \\textit{observed} relations, not dissimilar to the results of previous studies, which can be largely accounted for by the FIR $K$--correction using an M51 template. After $K$--correction $q_{70}$ is the only monochromatic index to show signs of evolution, suggesting that the 70-\\mum\\ $K$--correction may be less effective as a result of a steep slope in the SED from $\\sim 25-35\\mum$ (corresponding to $z\\sim1-2$) compared with M51. Observed MIPS colours at all redshifts are more consistent with the M51 template compared with hotter starburst galaxy templates, indicating that the typical IR SEDs of stellar-mass-selected galaxies at redshifts up to $\\sim 0.8$ (at least) appear to be dominated by cold dust. At higher redshifts it is not possible to constrain the dust temperature with MIPS colours, although it is still clear that M51 is the closest template. In contrast to this, both radio and total IR luminosities rise significantly with increasing redshift, as do derived SFRs. Specific SFRs similarly rise steeply, in agreement with {results in the literature \\citep{cowie_new_1996,madau_star_1998,brinchmann_mass_2000, bauer_specific_2005,feulner_connection_2005,prez-gonzlez_stellar_2008, dunne_star_2009,damen_evolution_2009,pannella_star_2009,oliver_specific_2010}.} The stacked radio data reveal tentative evidence for an evolution in radio spectral index across the redshift range, an unexpected result that implies some change in the radio loss processes in our sample towards higher redshifts. The most likely explanation seems to be a shift towards greater inverse-Compton losses of the CR electrons at $z>1$, {supporting} the predictions of \\citet{lacki_physics_2009}. Overall our results show evidence that the FRC, measured from 24-\\mum\\ fluxes or 160-\\mum\\ fluxes closer to the FIR peak, remains roughly constant up to $z\\sim2$, corresponding to 10 Gyr of cosmic time. This is similar to the conclusions of recent studies including \\citet{ibar_exploringinfrared/radio_2008, garn_relationship_2009, younger_millimetre_2009, ivison_blast:far-infrared/radio_2009} and \\citet{sargent_vla-cosmos_2010}. The issue is clouded however by measurements at 70\\mum, which appear to show a declining $q$ index with redshift, and when combined into a total IR luminosity, likewise show a slight decline (at low significance). This most likely implies a steeper spectral slope at wavelengths around $25-35$\\mum\\ (compared with the M51 template), leading to insufficient 70-\\mum\\ $K$--corrections. But a true evolution in the ratios of 70-\\mum/radio luminosity and of TIR/radio luminosity is plausible, considering the apparent increase in electron-calorimetry behaviour at $z>1$, and considering the fact that rest-frame 24, 70 and 160-\\mum\\ fluxes can arise from different components of the dust in a galaxy. It is also consistent with the results of \\citet{seymour_investigatingfar-ir/radio_2009} for $q_{70}$ and \\citeauthor{ivison_blast:far-infrared/radio_2009} (2010a/b) using BLAST/Herschel and \\textit{Spitzer} observations to measure $q_\\text{TIR}$. Constraining the FIR SED is one of the greatest problems in understanding the FRC and the dust emission in general from high redshift star-forming galaxies. Upcoming surveys with \\textit{Herschel}, such as the \\textit{Herschel} Multi-tiered Extragalactic Survey \\citep[HerMES;][]{oliver_herschel_2010} and the \\textit{Herschel} Astrophysical Terahertz Large Area Survey \\citep[H-ATLAS;][]{eales_herschel_2010} are anticipated to revolutionise our understanding of these topics by providing deep and wide observations spanning the peak of FIR emission across the history of cosmic star formation." }, "1005/1005.4891_arXiv.txt": { "abstract": "We present the gamma-ray data of the extraordinary flaring activity above 100 MeV from the flat spectrum radio quasar 3C~454.3 detected by AGILE during the month of December 2009. \\cp, that has been among the most active blazars of the FSRQ type since 2007, was detected in the gamma-ray range with a progressively rising flux since November 10, 2009. The gamma-ray flux reached a value comparable with that of the Vela pulsar on December 2, 2009. Remarkably, between {\\es December 2 and 3, 2009} the source more than doubled its gamma-ray emission and became the brightest gamma-ray source in the sky with a {\\es peak flux} of $F_{\\gamma,p} = (2000 \\pm 400) \\times 10^{-8} \\rm ph \\, cm^{-2} \\, s^{-1}$ for a 1-day integration above 100 MeV. The gamma-ray intensity decreased in the following days with the source flux remaining at large values near $F_{\\gamma} \\simeq (1000 \\pm 200) \\times 10^{-8} \\rm ph \\, cm^{-2} \\, s^{-1}$ for more than a week. This exceptional gamma-ray flare dissipated among the largest ever detected intrinsic radiated power in gamma-rays above 100 MeV {\\esc ($L_{\\gamma, source, peak} \\simeq 3 \\times 10^{46} \\; \\rm erg \\, s^{-1} $}, for a relativistic Doppler factor of $\\delta \\simeq 30$). The total isotropic irradiated energy of the month-long episode {\\es in the range 100 MeV -- 3 GeV} is $E_{\\gamma,iso} \\simeq 10^{56} \\, \\rm erg$. We report the intensity and spectral evolution of the gamma-ray emission across the flaring episode. We briefly discuss the important theoretical implications of our detection. ", "introduction": "Blazars (a special class of Active Galactic Nuclei with the relativistic jet pointing towards the Earth) show variability across their emitted spectrum on timescales of days, months, years. Rarely, intense gamma-ray flares are detected from blazars with fluxes reaching values near that of the Vela pulsar (i.e., the brightest steady gamma-ray source in the sky with a flux of $F_{\\gamma,Vela} \\simeq 900 \\times 10^{-8} \\rm ph \\, cm^{-2} \\, s^{-1}$ above 100 MeV). Even more rarely, a blazar gamma-ray ``super-flare'' reaches intensities substantially larger than $F_{\\gamma,Vela}$, {\\es as in the case of the June, 1995 flare from PKS 1622-29 (\\cite{1997ApJ...476..692M}). On these rare occasions, the gamma-ray sky is remarkably dominated by a single transient source. In this \\textit{Paper}, we report the observations by the AGILE satellite of the most recent gamma-ray super-flare from the blazar \\c during the period mid-November/mid-December, 2009. During a 1-month period, this source repeatedly reached a flux near $F_{\\gamma,Vela}$ for about 2 weeks, and then produced a very intense super-flare on {\\es December 2-3}, 2009, with $F_{\\gamma} > 2 F_{\\gamma,Vela} $. This flare turns out to be even more intense than that detected from PKS 1622-29 (\\cite{1997ApJ...476..692M}), and then qualifies as the most intense gamma-ray flare ever observed from a cosmic source at energies above 100 MeV. \\begin{figure*} % \\begin{center} \\includegraphics[width=12.cm, angle=270]{3c454_flare_agile_paper.eps} \\caption{Gamma-ray emission above 100 MeV from 3C~454.3 as monitored by AGILE. \\textit{(Upper panel:)} the 2 and 1/2 year flux lightcurve covering the period July, 2007 - December, 2009. The black data points are obtained with AGILE in the pointing mode (see \\cite{Vercellone2010:3C454}); the blue data points were collected in spinning mode. \\textit{(Lower panel:)} gamma-ray lightcurve obtained for the period November 07, 2009 and January 9, 2010 (spinning mode). All flux values are obtained with the AGILE standard maximum likelihood analysis, using the {\\es FM3.119} calibrated filter, with standard event selection that takes into account the SAA passage and the Earth albedo photons. The temporal bin for the blue data points is 2 days, except for the 9-days interval centered around the peak, for which a 1-day integration was performed. {\\escc Data obtained with a maximum off-axis angle $\\theta_{m}=60$ degrees.}} \\label{LC} \\end{center} \\end{figure*} The blazar \\c (PKS~2251$+$158; $z=0.859$) has been extensively studied over the last two decades. A wealth of multifrequency observations were obtained especially after the EGRET detections above 100 MeV during the 90's in the range $F_{\\gamma} = (40-140) \\times 10^{-8} \\rm ph \\, cm^{-2} \\, s^{-1}$ (\\cite{Hartman1992:3C454iauc}; \\cite{Aller1997:3C454_EGRET}). The source entered an active phase in 2000, and was very active in 2005-2006. {\\esc During the May-June, 2005 period the source showed} the strongest optical flare ever recorded in May, 2005, {\\esc and reached the} optical magnitude $R \\simeq 12$ (\\cite{vil06, Fuhrmann2006:3c454:opt}). X-ray (\\citealt{Giommi2006:3C454_Swift}) and hard X-ray observations (\\cite{Pian2006:3C454_Integral}) {\\esc covering this active phase in 2005} detected large fluxes between 10 and 100 mCrab. \\c has been subsequently monitored very extensively at all wavelengths \\citep{vil06,vil07, 2009A&A...504L...9V,rai07,rai08a}. Starting with the July, 2007 AGILE detection above 100 MeV (\\cite{Vercellone2008:3C454_ApJ}), \\c has been very active in gamma-rays, and certainly it can be referred as the most active blazar during the last 2 and 1/2 years. A series of papers describe the AGILE gamma-ray observations in 2007-2009 showing repeated flares usually in coincidence with periods of intense optical and X-ray enhanced activity (\\cite{Vercellone2008:3C454_ApJ}, \\cite{Vercellone2008:3c454:ApJ_P1}, \\cite{Vercellone2010:3C454}, \\cite{2009ApJ...707.1115D}); {\\esc see also the 1AGL catalog (\\cite{2009A&A...506.1563P}}). The multi-year optical evolution of \\c has been presented in \\cite{2008A&A...491..755R}. Also \\textit{Fermi} detected several gamma-ray flaring episodes above 100 MeV (\\cite{2009ATel.2200....1H,2008ATel.1628....1T}), and determined an average spectrum for the August-September, 2009 period in the range 200 MeV -- 10 GeV showing a distinct {\\es break in the power-law spectrum at energies of 2 -- 3 GeV} (\\cite{Abdo2009:3C454}, {\\es \\cite{2010ApJ...710.1271A}}). ", "conclusions": "\\c reveals itself as the most prolific gamma-ray blazar during the last 3 years, and is dominating the {\\esc gamma-ray} sky above 100 MeV since mid-2007. During the period of December, 2009 the source showed a dramatic activity reaching and maintaining for several weeks a flux above 100 MeV comparable or larger than the brightest persistent gamma-ray source such as the Vela pulsar. During the period Dec. 2-3, 2009 \\c produced a super-flare that turns out to be the brightest blazar emission episode above 100 MeV ever detected. The AGILE satellite followed {\\esc in a continuous way} the daily evolution of the flaring activity of \\cp. Even though a comprehensive picture of the physical mechanism at work can be obtained only from a multifrequency collection of simultaneous data, restricting ourselves to the gamma-ray range in any case provides very important information on the physics of the source. For a detailed theoretical modelling and a broad spectral evolution of the \\c exceptional activity see \\cite{2010arXiv1005.3263P}." }, "1005/1005.3649_arXiv.txt": { "abstract": "We present radio images of a sample of six Wide-Angle Tail (WAT) radio sources identified in the ATLAS 1.4\\,GHz radio survey, and new spectroscopic redshifts for four of these sources. These WATs are in the redshift range of 0.1469$-$0.3762, and we find evidence of galaxy overdensities in the vicinity of four of the WATs from either spectroscopic or photometric redshifts. We also present follow-up spectroscopic observations of the area surrounding the largest WAT, S1189, which is at a redshift of $\\sim$0.22. The spectroscopic observations, taken using the AAOmega spectrograph on the AAT, show an overdensity of galaxies at this redshift. The galaxies are spread over an unusually large area of $\\sim$12 Mpc with a velocity spread of $\\sim$4500 km s$^{-1}$. This large-scale structure includes a highly asymmetric FRI radio galaxy and also appears to host a radio relic. It may represent an unrelaxed system with different sub-structures interacting or merging with one another. We discuss the implications of these observations for future large-scale radio surveys. ", "introduction": "\\label{intro} Wide-Angle Tail (WAT) galaxies are radio galaxies whose radio jets appear to bend in a common direction. They are generally detected in dynamical, non-relaxed clusters of galaxies \\citep[e.g.][]{Burns90} and may be used as probes or tracers for clusters \\citep{Blanton00, Blanton01}. Clusters of galaxies are the largest gravitationally bound structures in the Universe and are powerful testbeds of cosmological models \\citep[e.g.][] {Borgani04, Sahlen09, Kravtsov09}. Clusters also host diffuse radio emission in the form of radio haloes and relics \\citep{Giovannini00, Feretti05, Ferrari08, Giovannini09}. The bent nature of WATs has commonly been attributed to strong intra-cluster winds caused by dynamical interactions such as cluster-cluster mergers \\citep{Burns98}. WATs are preferentially found in enhanced X-ray regions \\citep{Pinkney00} and are usually associated with dominant cluster galaxies \\citep{Owen76}. \\citet{Mao09a} found the tailed radio galaxies, including WATs, to be located in the densest regions of clusters in the local Universe, consistent with earlier studies \\citep[e.g.][] {Burns90, Blanton00, Blanton01}. Thus WATs represent valuable tracers of high density regions in the intracluster medium (ICM), and this approach has been used in a number of recent studies \\citep[e.g.][] {Blanton00, Blanton03, Smolcic07, Giacintucci09, Kantharia09, Oklopcic10}. Here we present the radio properties of six WATs that we have identified in ATLAS, the Australia Telescope Large Area Survey, carried out with the Australia Telescope Compact Array (ATCA) at 1.4\\,GHz \\citep{Norris06,Middelberg08}. ATLAS \\footnote{http://www.atnf.csiro.au/research/deep/index.html} will image seven square degrees of sky over two fields to an rms sensitivity of 10 $\\mu$Jy beam$^{-1}$. The ATLAS fields have been observed with a number of different ATCA configurations, and the typical resolution of the observations is $\\sim$10 arcsec. The two ATLAS fields, Chandra Deep Field South (CDFS) and European Large Area ISO Survey-South 1 (ELAIS-S1), were chosen to coincide with the \\textit{Spitzer} Wide-Area InfraRed Extragalactic (SWIRE) survey program \\citep{Lonsdale03} so that corresponding optical and infrared photometric data are available. In addition to the radio properties we present new spectroscopic redshifts for four of the WATs and follow-up spectroscopic observations of galaxies in the vicinity of the largest WAT in order to probe its surrounding structure. This WAT was first identified as radio source S1189 by \\citet{Middelberg08}, and is associated with the SWIRE source SWIRE4\\_J003427.54-430222.5 \\citep{Lonsdale03}. In this paper we present a summary of the data in Section 2, while the WATs in ATLAS are presented in Section 3. Section 4 presents the results of spectroscopic observations of S1189 and its surrounding region, and discusses the large-scale structure in its vicinity. In Section 5 we discuss cosmological inverse-Compton quenching and the implications for deep wide radio surveys and ATLAS. This paper uses H$_0$ = 71 km s$^{-1}$ Mpc$^{-1}$, $\\Omega$$_M$ = 0.27 and $\\Omega$$_\\Lambda$ = 0.73 and the web-based calculator of \\citet{Wright06} to estimate the physical parameters. Vega magnitudes are used throughout. ", "conclusions": "We have identified a sample of six Wide-Angle Tail (WAT) radio sources. We present new spectroscopic redshifts for four of these sources, and find that these WATs lie in the redshift range 0.1469$-$0.3762. We have examined the fields using both spectroscopic and photometric redshifts of galaxies in the vicinity of the WATs and find evidence of an overdensity of galaxies in four of these WATs. From a more detailed study of the field around S1189 we find an overdensity of galaxies which is spread over $\\sim$12 Mpc and has a velocity spread of $\\sim$4500 km s$^{-1}$, and a velocity dispersion of $\\sim$870 km s$^{-1}$. This large-scale structure hosts a putative cD galaxy with, at best, weak radio emission, a radio relic which has a size of $\\sim$274 kpc, and an asymmetric FRI radio galaxy with an extent of $\\sim$559 kpc. The peak brightness at the extremities of the outer lobes of the FRI source differ by a factor of $\\sim$4, possibly due to differences in the environment on opposite sides. The minor axis of the relic is not directed towards either the host galaxy of the WAT or the putative cD galaxy. This large-scale structure may represent an unrelaxed system with different sub-structures interacting or merging with one another. Therefore, deep X-ray observations of the field would be very valuable to further understand this interesting large-scale structure. WATs are known to occur in clusters of galaxies, and could in principle be useful tracers of clusters at moderate and high redshifts. IC cooling of electrons by interaction with CMBR increases rapidly with z. However, this does not imply a sharp drop in the number of WATs at high z. Deep and wide-field surveys, such as the Evolutionary Map of the Universe (EMU) \\citep{Norris09}, should provide additional information and insights on the range of structures at moderate and high redshifts. We expect these to be invaluable probes of large-scale structure." }, "1005/1005.3822_arXiv.txt": { "abstract": "{{We use deep, five band (100--500 $\\mu$m) data from the {\\it Herschel} Lensing Survey (HLS) to fully constrain the obscured star formation rate, SFR$_{\\rm FIR}$, of galaxies in the Bullet cluster ($z$ = 0.296), and a smaller background system ($z$ = 0.35) in the same field. {\\it Herschel} detects 23 Bullet cluster members with a total SFR$_{\\rm FIR}$ = 144 $\\pm$ 14 M$_{\\sun}$ yr$^{-1}$. On average, the background system contains brighter {far-infrared} (FIR) galaxies, with $\\sim$50\\% higher SFR$_{\\rm FIR}$ (21 galaxies; 207 $\\pm$ 9 M$_{\\sun}$ yr$^{-1}$). SFRs extrapolated from 24 $\\mu$m flux via recent templates (SFR$_{24{\\mu}m}$) agree well with SFR$_{\\rm FIR}$ for $\\sim$60\\% of the cluster galaxies. In the remaining $\\sim$40\\%, SFR$_{24{\\mu}m}$ underestimates SFR$_{\\rm FIR}$ due to a significant excess in observed $S_{100}/S_{24}$ (rest frame $S_{75}/S_{18}$) compared to templates of the same FIR luminosity.}} ", "introduction": "In the last decade many studies have attempted to quantify the star formation rate (SFR) within cluster galaxies. Ultraviolet and optical observations have successfully identified trends between unobscured star formation and local environment, suggesting that star formation in cluster core galaxies is generally more quenched \\citep[e.g.][]{kod04-1103,por07-1409}. However, star formation can be obscured by dust, which re-emits stellar light in the far-infrared (FIR), peaking at a rest frame $\\lambda_0$ $\\sim$ 100 $\\mu$m. Mid-infrared surveys \\citep[e.g.][]{met05-425,gea06-661,fad08-9} have explored obscured star formation by estimating total FIR luminosity from template spectra. These templates are often based on small numbers of well constrained local galaxies, e.g. \\citet{rie09-556}. The PACS \\citep{pog10} and SPIRE \\citep{gri10} instruments, onboard the ESA {\\it Herschel} Space Observatory \\citep{pil10}, enable unprecedented multi-band coverage of the FIR. The {\\it Herschel} Lensing Survey ({HLS}; PI: E Egami) consists of 5-band observations (100--500 $\\mu$m) of 40 nearby clusters ($z$ $\\sim$ 0.2--0.4). Nominally devised to exploit the gravitational lensing effect of massive clusters to observe high redshift galaxies (see \\citealt{ega10} for details on survey design), a useful by-product is deep FIR observations of the clusters themselves. At these redshifts, {\\it Herschel} photometry spans the peak of the dust component, allowing an accurate constraint of far infrared luminosity, {$L_{\\rm FIR}$}, and hence obscured SFR. During the {\\it Herschel} Science Demonstration Phase, {HLS} observed the Bullet cluster (1E0657--56; $z$ = 0.296). The reason for this choice was two-fold. First, previous studies report bright {submillimeter} galaxies in the background \\citep[e.g. ][]{rex09-348}, {with HLS analysis presented in \\citet{rex10}}. Second, the Bullet cluster is a recent collision of two clusters \\citep{mar02-27}, offering a unique laboratory for the study of star formation within a dynamic environment. The sub-cluster has conveniently fallen through the main cluster perpendicular to the line of sight ($< 8^\\circ$ from the sky plane; \\citealt{mar04-819}). Analysis of X-ray emission shows that a supersonic bow shock precedes the hot gas, while the weak lensing mass profile indicates that this X-ray bright component lags behind the sub-cluster galaxies due to ram pressure \\citep{mar02-27,bar02-816}. A recent mid-infrared study by \\citet{chu09-963} concluded that ram pressure from the merger event had no significant impact on the star formation rates of nearby galaxies. We can re-evaluate these previous studies by using {\\it Herschel} data to constrain {$L_{\\rm FIR}$} directly. In this letter, we present an exploration of obscured star formation in this cluster environment. \\begin{figure} \\centering \\includegraphics[height=80mm,angle=270,clip]{figs/zhist.eps} \\caption{Distribution of spectroscopic redshifts (0.27 $<$ $z$ $<$ 0.37) for galaxies within the Bullet cluster field (outline). {\\it Herschel} detected galaxies are also shown (filled). In addition to the Bullet cluster ($z$ = 0.296), there is a background {system} at $z$ = 0.350. Dotted lines show our membership limits of 3000 km s$^{-1}$ and 2000 km s$^{-1}$ respectively.} \\label{fig:zhist} \\end{figure} ", "conclusions": "Using deep {\\it Herschel} observations (100--500 $\\mu$m) to fully constrain the FIR component, we derive obscured SFRs for galaxies in the Bullet cluster ($z$ = 0.296), and a background system ($z$ = 0.35) in the same field. {{\\it Herschel} detects 23 Bullet cluster members, with a total SFR$_{\\rm FIR}$ = 144 $\\pm$ 14 M$_{\\sun}$ yr$^{-1}$, while the background system contains 21 detections but $\\sim$50\\% higher SFR (207 $\\pm$ 9 M$_{\\sun}$ yr$^{-1}$). The relative distributions of SFR$_{\\rm FIR}$ and optical flux suggest a difference in dust retention between the two systems. For $\\sim$60\\% of galaxies, SFR$_{\\rm FIR}$ agrees well with estimated SFRs from 24 $\\mu$m flux via recent templates. However, the remaining galaxies display a significant excess at 100 $\\mu$m ($\\lambda_{\\rm 0}$ $\\approx$ 75 $\\mu$m) compared to templates, which causes an under-prediction in SFR$_{24{\\mu}m}$}. We note that such an excess is not found in the high redshift, field sample \\citep{rex10}. Future studies will exploit the full range of 5-band {\\it Herschel} cluster observations available in {HLS}, to form a more complete understanding of the environmental effect on obscured star formation rates, and explore the origin and dependencies of the 100 $\\mu$m excess." }, "1005/1005.3539_arXiv.txt": { "abstract": "The origin of the atmosphere of the largest moon of Saturn, Titan, is poorly understood and its chemistry is rather complicated. Ground-based millimeter/sub-millimeter heterodyne spectroscopy resolves line shapes sufficiently to determine information in Titan's atmospheric composition (on vertical profiles and isotopic ratios). We test the capabilities of the Swedish Heterodyne Facility Instrument (SHFI), Receiver APEX-1, together with the Atacama Pathfinder EXperiment APEX 12-m telescope for Titan's atmospheric observations. In particular we present sub-millimeter observations of the CO(2-1) and HCN(3-2) lines of the Titan stratosphere with APEX, and with SHFI taken during the Science Verification (SV) instrument phase on March and June 2008. With the help of appropriate radiative transfer calculations we investigate the possibility to constrain the chemical concentrations and optimize the performance of the APEX-1 instrument for inferring vertical profiles of molecular components of the atmosphere of Titan. \\\\ This study attempts to contribute to constrain radiative transfer and retrieval algorithms for planetary atmospheres, and to give hints to the current and future ground and space-based data acquisition leading to a more thorough understanding of the chemical composition of Titan. ", "introduction": "Titan's atmosphere exhibits a complex photochemistry. The origin of carbon monoxide (CO) is not well understood (whether photochemical or primordial). Hydrogen Cyanide (HCN), the most abundant nitrile in Titan, is a key intermediate in production of more complex hydrocarbons and organic molecules. HCN in the Titan atmosphere has been discovered by the infrared observations of Voyager\\,1\\textcolor{black}{\\cite{h81}}, and CO has been detected by the ground-based near-infrared observations\\textcolor{black}{\\cite{l83}}. Following these detections, the molecular concentrations of CO and HCN in Titan's atmosphere have been determined from infrared, millimeter, and sub-millimeter observations as well as from modeling (see Tables\\,2 and 3 in section\\,4 for details). CO seems to be uniformly mixed in Titan's atmosphere up to high altitudes. HCN abundances, however, display a steeper profile with ambiguous enrichment values (Table\\,3). Already several observational data confirm that Titan's atmospheric composition is indeed seasonally\\textcolor{black}{\\cite{roe04,teanby08}}, and spatially dependent\\textcolor{black}{\\cite{teanby09}}. Full behavior of the seasonal characteristics of the spatial distribution is not constrained yet due to the limitation in the temporal coverage of the previous observations, and to the use of different instruments. New sub-millimeter observations are required not only to provide new abundance constraints and shed more light on the rate of these seasonal variations, for example, but for support, complement and cross-calibration the ESA's Herschel Space Observatory mission. One of the goals of the Key Program of Herschel, entitled \\textit{Water and Related Chemistry in the Solar System}\\textcolor{black}{\\cite{ha09}}, is to understand water inventory in the Titan's atmosphere as well as distributions of other hydrocarbons and nitriles. During the guaranteed time, line surveys on Titan at the frequency range of 500 and 5000\\,GHz are going to be carried out using two low-to-medium resolution spectrometers (Photodetector Array Camera and Spectrometer (PACS) and Spectral and Photometric Imaging Receiver (SPIRE)). Our model calculations of the synthetic spectra of Titan expected to be observed with Herschel show that several CO and HCN lines will be detected with PACS and SPIRE\\textcolor{black}{\\cite{re09}}. Because of their low-to-medium spectral resolutions, SPIRE and PACS are not capable of measuring the shapes of CO and HCN lines, which prevents us from determining precisely the vertical profile of CO and HCN mixing ratios. Therefore, high spectral resolution observations of CO and HCN with ground-based sub-millimeter telescopes would significantly improve the accuracy of retrieving the CO and HCN profiles. The Atacama Pathfinder EXperiment (APEX) 12-m telescope is operational since 2005\\textcolor{black}{\\cite{g06}}, and has been already used to the planetary science (monitoring mesospheric winds of Venus\\textcolor{black}{\\cite{l08}}), among numerous galactic and extra-galactic objects. In spring 2008, the new APEX Swedish Heterodyne Facility Instrument (SHFI) has been commissioned on the APEX\\,12-m telescope\\textcolor{black}{\\cite{vas08,nys09}}, which consists of four single-pixel receivers (APEX-1, APEX-2, APEX-3, and APEX-T2) mounted in a single cryostat located in the Nasmyth\\,A cabin of APEX. In this report we present a summary of our observations of CO and HCN performed with APEX-1 during the Science Verification (SV) phase with the aims to demonstrate the capabilities of APEX and SHFI for Titan's atmospheric observations, and to investigate the possible retrieval of CO and HCN abundances. We also examine the possible detection of the millimeter/sub-millimeter rotational transition of C$_{2}$H$_{2}$ (already observed in the mid-infrared spectra by the Composite Infrared Spectrometer, CIRS, on Cassini\\textcolor{black}{\\cite{vina07}}), and the search for one new compound, HC$_{5}$N(83-82) which the possible presence has been suggested by experimental results simulating Titan's atmosphere \\textcolor{black}{\\cite{van95,coll95,co06}}. The observations are described in detail in section\\,2 while the radiative transfer modeling and the data analysis are discussed in sections\\,3 and 4, respectively. A brief discussion is given in section\\,5 and a summary in section\\,6. ", "conclusions": "How does one explain the behavior of the retrieved HCN vertical profile obtained here? the relative poor signal-to-noise of the observations obtained here results in a poor fit to the HCN line wings. Furthermore, comparing the results obtained with previous results derived from disk-averaged information, interferometer measurements, and space-based observations summarized in Table\\,3 must be considered cautiously. In other words, direct comparison of the results from different instruments requires us to assume that all instrument-related offsets have been accounted for. Certainly, Titan's atmospheric temperature and composition are not in a steady state. Does the HNC abundance retrieved here perhaps indicate that the composition of Titan's atmosphere has changed during the last years? Fig.\\,7 in Teanby et al.\\,2009\\textcolor{black}{\\cite{teanby09}} shows that the HCN abundance measured by Cassini (at southern latitudes) has not significantly changed between 2006 and 2008. Then it seems to be that possible detections of time variations (if any) between different observations could be maybe due to observing more of the northern hemisphere than of the southern one as Titan's season changes. As we derived disk-averaged information we need to keep in mind, however, that our vertical distributions are mostly representative of the equatorial region of Titan since the measured flux density spectra are more heavily weighted toward equatorial latitudes. In case that vertical distributions would be representative of the northern latitudes, the profile analysis would become rather complicated. There is evidence of a vortex, which acts to separate enriched from unenriched air, at $25-55\\,^{\\circ}$\\,N encircling the North Pole \\textcolor{black}{\\cite{teanby09,fla,ach08}}. Further higher signal-to-noise HCN observations at different times are necessary in order to open the possibility to detect seasonal and even spatial variations in HCN, and confirm or not an enrichment layer in Titan's stratosphere." }, "1005/1005.3013_arXiv.txt": { "abstract": "We demonstrate a new model which uses an ADD type braneworld scenario to produce a multi-state theory of dark matter. Compactification of the extra dimensions onto a sphere leads to the association of a single complex scalar in the bulk with multiple Kaluza-Klein towers in an effective four-dimensional theory. A mutually interacting multi-state theory of dark matter arises naturally within which the dark matter states are identified with the lightest Kaluza-Klein particles of fixed magnetic quantum number. These states are protected from decay by a combination of a global $U(1)$ symmetry and the continuous rotational symmetry about the polar axis of the spherical geometry. We briefly discuss the relic abundance calculation and investigate the spin-independent elastic scattering off nucleons of the lightest and next-to-lightest dark matter states. \\\\ \\\\ \\\\ ", "introduction": "The standard model (SM) of particle physics, while describing the results of collider experiments with unprecedented precision, lacks a suitable candidate for dark matter. This is a problem that deserves urgent attention, as cosmological observations have both measured the energy density of atoms and all other SM particles to be $\\Omega_{\\rm SM} h^2 \\simeq 0.02$, and shown that a new type of \\mbox{\\emph{non-SM}} dark matter contributes $\\Omega_{d} h^2 \\simeq 0.11$ to the density of the Universe \\cite{Dunkley:2008ie}. Corroborating evidence from astrophysics and structure formation suggests that dark matter, whatever it is, appears like non-relativistic cold dark matter (CDM) particles interacting at most weakly with SM particles (see, e.g., Refs.~\\cite{Jungman:1995df,Bergstrom:2000pn,Bertone:2004pz,Hooper:2007qk} or, e.g., Ref.~\\cite{D'Amico:2009df} for a pedagogical review). Most of the theoretical focus to date has been on minimal dark matter theories containing a single new stable elementary particle. However, while this is a reasonable conservative assumption, this need not be the case. Furthermore, anomalies such as recent cosmic ray results \\cite{:2008zzr,Adriani:2008zr,Abdo:2009zk} and long-observed annual modulation by the DAMA experiment \\cite{Bernabei:2008yi} have sparked theoretical interest in a non-minimal dark sector. It is worth investigating the possibility that the dark sector might not be so simple and may, in reality, be better described by a multi-particle set up (see, e.g., Refs.~\\cite{Zurek:2008qg,Profumo:2009tb}). Extra dimensional models have been studied extensively as a possibility for new physics in both particle physics and cosmology. Interest in these models has been fueled by, amongst other things, their ability to accommodate Higgsless electroweak symmetry breaking \\cite{Csaki:2003zu,Csaki:2003sh} and provide potential solutions to the hierarchy problem \\cite{ArkaniHamed:1998rs,ArkaniHamed:1998nn,Antoniadis:1998ig,Randall:1999ee,Davoudiasl:1999tf,Huber:2000ie}. In particular, the universal extra dimensional (UED) model has been a subject of intense investigation recently because it can easily provide a dark matter candidate \\cite{Servant:2002aq,Servant:2002hb,Bertone:2002ms}. This model, originally proposed in Ref.~\\cite{Appelquist:2000nn}, is so named because all particles are allowed to propagate in the bulk and have universal access to all compact dimensions \\cite{Appelquist:2000nn}. This is in contrast to both \\mbox{Arkani-Hamed--Dimopoulos--Dvali} (ADD) type models where all standard model fields are restricted to a brane \\cite{ArkaniHamed:1998rs} or models in which only some of the standard model fields can access the compact dimensions \\cite{Muck:2001yv,Muck:2003kx}. A key issue when constructing a particle theory of dark matter is its stability on cosmological timescales. In the case of the UED models a vestigial discrete translation invariance, known as Kaluza-Klein (KK) parity, ensures stability of the lightest KK particle (LKP) to all orders in perturbation theory. In this paper we discuss a theory containing many scalar dark matter states that originate from a single complex scalar in the factorized spacetime, $M_4 \\times S^2$. This spherical compactification, with associated spherical-harmonic eigenfunctions labeled by quantum numbers $\\ell$ and $m$, naturally organizes the associated four-dimensional KK states, $\\tchi_\\ell^m$, into towers of definite $m$, with $\\tchi_{|m|}^m$ being the lightest state in each tower. These lightest KK particles are nominally stable due to the residual rotational symmetry about the polar axis of the extra dimensional geometry and an imposed global $U(1)$ symmetry.\\footnote{Stable on cosmologically-interesting timescales. While gravitational decay may occur we imagine here an effective theory with stabilized extra-dimensions and very massive shape moduli.} This set of stable KK particles (SKPs) then comprises the set of dark matter states in the theory. Since this type of compactification does not allow for a chiral zero-mode fermion in the four-dimensional (4D) effective spectrum \\cite{Camporesi:1995fb, Abrikosov:2001nj}, we restrict ourselves to a non-universal extra dimensional scenario. For simplicity, we also restrict the entire standard model field content, not just the fermions, to the brane on $S^2$. As the position of the standard model brane on $S^2$ is completely arbitrary we can choose our coordinates such that it resides at the north pole.\\footnote{We imagine here that the ultraviolet physics fixing the brane at the pole renders branon fluctuations very massive.} While the location of the brane breaks some of the spherical rotational symmetry, rotations about the polar axis are preserved. Although we discuss here the simplest spherical case, we expect similar results will hold for any manifold which preserves a continuous isometry of the extra dimensional ground state metric. In Section~\\ref{sec:formalism} we describe the formalism of a complex scalar field on the $M_4 \\times S^2$ spacetime while Section~\\ref{sec:instability} discusses the instability of the excited KK particles and the stability of the SKPs within our effective framework. In Section~\\ref{sec:abundance} we briefly describe the relic abundances of the corresponding SKPs and in section~\\ref{sec:scat} we investigate the spin-independent nucleon scattering related to relevant direct-detection experiments. We make some final statements and discuss our conclusions in Section~\\ref{sec:conc}. ", "conclusions": "\\label{sec:conc} We have demonstrated a novel particle model which uses an ADD type braneworld scenario to produce a multi-state theory of dark matter. Compactification of the extra dimensions onto a sphere leads to the association of a single complex scalar bulk field with multiple KK towers in the 4D effective theory. The lightest KK state in each tower was then naturally stabilized by the combination of an imposed $U(1)$ symmetry and a subset of the continuous spherical symmetry of the extra dimensions(though much of the formalism we discuss here would remain true even if these states were meta-stable on cosmological timescales). We have also shown that viable regions of the parameter space exist that can readily produce the observed dark-matter relic abundance~\\cite{Winslow&Sigurdson}. This model remains unconstrained after comparison of the spin-independent $\\tchi_0^0$-proton cross section with recent direct detection data. The most important aspect of this paper is that the multi-state nature of the dark matter is a direct and natural consequence of the continuous compactification geometry." }, "1005/1005.3103_arXiv.txt": { "abstract": "We present a strong lens analysis of SDSS~J1004+4112, a unique quasar lens produced by a massive cluster of galaxies at $z=0.68$, using a newly developed software for gravitational lensing. We find that our parametric mass model well reproduces all observations including the positions of quasar images as well as those of multiply imaged galaxies with measured spectroscopic redshifts, time delays between quasar images, and the positions of faint central images. The predicted large total magnification of $\\mu \\sim 70$ suggests that the lens system is indeed a useful site for studying the fine structure of a distant quasar and its host galaxy. The dark halo component is found to be unimodal centered on the brightest cluster galaxy and the {\\it Chandra} X-ray surface brightness profile. In addition, the orientation of the halo component is quite consistent with those of the brightest cluster galaxy and member galaxy distribution, implying that the lensing cluster is a relaxed system. The radial profile of the best-fit mass model is in good agreement with a mass profile inferred from the X-ray observation. While the inner radial slope of the dark halo component is consistent with being $-1$, a clear dependence of the predicted A--D time delay on the slope indicates that an additional time delay measurement will improve constraints on the mass model. ", "introduction": "SDSS~J1004+4112 is a unique quasar lens system \\citep{inada03,oguri04,inada05,inada08}. A quasar at $z=1.734$ is multiply imaged into five images, with their maximum image separation of $14\\farcs7$, produced by a massive cluster of galaxies at $z=0.68$. It is one of only two known examples of cluster-scale quasar lenses, the other being the triple lens SDSS~J1029+2623 with the maximum image separation of $22\\farcs5$ \\citep{inada06,oguri08}. In addition to the quasar images, SDSS~J1004+4112 contains spectroscopically confirmed multiply imaged galaxies at $z\\sim 3$ \\citep{sharon05}. Both the quasar images and the lensing cluster have been detected in {\\it Chandra} X-ray observations \\citep{ota06}. Moreover, time delays between some of the quasar images have also been measured from long-term optical monitoring observations \\citep{fohlmeister07,fohlmeister08}. \\begin{table*} \\caption{Previous mass modeling of SDSS~J1004+4112\\label{table:model}} \\begin{center} \\begin{tabular}{lccc} \\hline\\hline Reference & model\\footnotemark[$*$] & constraints (quasar)\\footnotemark[$\\dagger$] & constraints (galaxy)\\footnotemark[$\\dagger$]\\\\ \\hline \\citet{inada03} & SIE+pert & pos+flux (4) & $\\cdots$ \\\\ \\citet{oguri04} & NFW+SIE+pert & pos+flux (4) & $\\cdots$ \\\\ \\citet{williams04}& non-parametric& pos (4) & $\\cdots$ \\\\ \\citet{sharon05} & SPL+gals & pos (5) & pos (5) \\\\ \\citet{kawano06} & gNFW+SIE+pert & pos+flux (4) & $\\cdots$ \\\\ \\citet{saha06} & non-parametric& pos (5) & pos (8) \\\\ \\citet{fohlmeister07} & NFW+deV+gals+pert & pos+flux (5) & $\\cdots$ \\\\ \\citet{saha07} & non-parametric& pos (5), $\\Delta t$ (2) & pos (8) \\\\ \\citet{inada08} & NFW+SIE+gals+pert & pos+flux (5), $\\Delta t$ (2) & $\\cdots$\\\\ \\citet{liesenborgs09}& non-parametric & pos (5), $\\Delta t$ (3) & pos (7) \\\\ \\hline This work & gNFW+Jaffe+gals+pert & pos+flux (5), $\\Delta t$ (3) & pos (27)\\\\ \\hline \\multicolumn{4}{@{}l@{}}{\\hbox to 0pt{\\parbox{160mm}{\\footnotesize \\footnotemark[$*$] Name of models: SIE = singular isothermal ellipsoid, pert = external perturbation (e.g., external shear), NFW = Navarro, Frenk \\& White (NFW) profile, SPL = softened power-law model, gals = perturbations from member galaxies, typically modeled by truncated isothermal profiles, gNFW = generalized NFW profile, dev = de Vaucouleurs profile, Jaffe = pseudo-Jaffe profile. \\par\\noindent \\footnotemark[$\\dagger$] ``pos'' indicates constraints from image positions, ``flux'' from fluxes of images, and $\\Delta t$ from time delays. Numbers in parentheses show the number of images used as constraints. }\\hss}} \\end{tabular} \\end{center} \\end{table*} Such a wealth of observational data available enable detailed investigations of the central mass distribution of the lensing cluster. Indeed, there have been several attempts to model the mass distribution of SDSS~J1004+4112, using either parametric or non-parametric method. \\citet{oguri04} adopted two-component (halo plus central galaxy) model to successfully reproduce the positions of four quasar images, but even the parities and temporal ordering of the quasar images could not be determined because of model degeneracies. \\citet{kawano06} extended mass modeling along this line, and explored how time delay measurements can distinguish different mass profiles. \\citet{fohlmeister07} pointed out that it is important to include perturbations from member galaxies to reproduce the observed time delay between quasar images A and B. On the other hand, \\citet{williams04} and \\citet{saha07} performed non-parametric mass modeling to show possible substructures in the lensing cluster. From the non-parametric mass modeling, \\citet{saha06} and \\citet{liesenborgs09} concluded that the radial mass profile is consistent with that predicted in $N$-body simulation \\citep{navarro97}. We summarize previous mass modeling in Table~\\ref{table:model}. In this paper, we revisit strong lens modeling of SDSS~J1004+4112 adopting a parametric mass model. We include many observational constrains currently available, including central images of lensed quasars and galaxies, flux rations, and time delay between quasar images (see Table~\\ref{table:model}). In particular this paper represents first parametric mass modeling that includes {\\it both} the quasar time delays and the positions of multiply imaged galaxies as constraints. We compare our best-fit mass model with the {\\it Chandra} X-ray observation of this system \\citep{ota06}. This paper is organized as follow. We describe our mass model in \\S\\ref{sec:model}. We show our results in \\S\\ref{sec:result}, and give conclusion in \\S\\ref{sec:model}. A new software for gravitational lensing, which is used for the mass modeling, is presented in Appendix. Throughout the paper we adopt the matter density of $\\Omega_M=0.26$ and the cosmological constant of $\\Omega_\\Lambda=0.74$, but regard the dimensionless Hubble constant $h$ as a parameter. With this choice of cosmological parameters, a physical transverse distance of $1h^{-1}$~kpc at the redshift of the lensing cluster ($z=0.68$) corresponds to $0.20$~arcsec. We denote a angular diameter distance from observer to lens as $D_{\\rm l}$, from observer to source as $D_{\\rm s}$, and from lens to source as $D_{\\rm ls}$. ", "conclusions": "\\label{sec:summary} We have revisited parametric mass modeling of SDSS~J1004+4112, a unique quasar--cluster lens system, using a newly developed mass modeling software. We include several new constraints, including positions of spectroscopically confirmed multiply imaged galaxies, time delays between some quasar images, and faint central images. Our model comprising of a halo component modeled by the generalized NFW profile, member galaxies including the brightest cluster galaxy G1, and perturbation terms, well successfully reproduced all observations including time delays. Unlike earlier claims based on parametric mass modeling, we have found that the center and orientation of the dark halo component is in good agreement with those of member galaxies and {\\it Chandra} X-ray observation, implying that the cluster is highly relaxed. The radial profile from strong lensing is also in excellent agreement with the mass profile inferred from the X-ray observation. Our mass modeling prefers the dark halo component with the inner slope close to $\\alpha=1$, being consistent with so-called NFW density profile. Additional measurement of the time delay between quasar image A and D will be useful to constrain the mass model further. The predicted total magnification of $\\mu_{\\rm tot}=67$ for the NFW profile is quite large compared with those for typical galaxy-scale lenses, because of the shallower density profiles for clusters. This makes the lens system a good site for studying the fine structure of the quasar through microlensing (\\cite{richards04}; \\cite{green06}; \\cite{lamer06}; \\cite{pooley07}) or for studying host galaxies of distant quasars \\citep{ross09}. We note that our result based on parametric mass modeling is broadly consistent with recent non-parametric mass modeling by e.g., \\citet{liesenborgs09}. Our result suggests that the core of the lensing cluster at $z=0.68$ is highly evolved. Recently, \\citet{limousin10} showed that the cluster MACSJ1423.8+2404 at $z=0.54$ is similarly highly relaxed based on the comparison of mass, light, and gas distributions. Therefore our result may point to the fact that relaxed clusters are quite common already at $z\\sim 0.6$. It is clear that the lensing cluster SDSS~J1004+4112 is currently one of the best-studied high-redshift clusters whose inner density profile is very tightly constrained by strong lensing. Additional constraints on this cluster with weak lensing, Sunyaev-Zel'dovich effect, and spectroscopic identifications of member galaxies will be important to advance our understanding of high-redshift clusters. \\bigskip I would like to thank the referee, Marceau Limousin, for useful comments and suggestion. \\appendix" }, "1005/1005.1937_arXiv.txt": { "abstract": "{ The intermediate-mass star-forming core UYSO~1 has previously been found to exhibit intriguing features. While deeply embedded and previously only identified by means of its (sub-)millimeter emission, it drives two powerful, dynamically young, molecular outflows. Although the process of star formation has obviously started, the chemical composition is still pristine. We present {\\it Herschel} PACS and SPIRE continuum data of this presumably very young region. The now complete coverage of the spectral energy peak allows us to precisely constrain the elevated temperature of 26\\,--\\,28 K for the main bulge of gas associated with UYSO1, which is located at the interface between the hot H {\\sc ii} region Sh 2-297 and the cold dark nebula LDN 1657A. Furthermore, the data identify cooler compact far-infrared sources of just a few solar masses, hidden in this neighbouring dark cloud. } ", "introduction": "Star formation occurs predominantly in structured environments in the case of intermediate and high-mass star formation. In particular, one often finds different stages of evolution located within close proximity. This poses a challenge for working on the earliest phases of star formation, which can only be revealed with long-wavelength observations. To observe these deeply embedded objects at the spectral peak of their emission, airborne or satellite missions are necessary, which in the past delivered only a very modest spatial resolution ($>$20$''$). We use ESA's new far-IR and sub-millimeter satellite {\\it Herschel} \\citep{A&ASpecialIssue-Herschel} --- with its unprecedented spatial resolution --- to scrutinise a sample of very young low- and high-mass star-forming cores within our programme {``Earliest Phases of Star Formation''} (EPoS, PI: O.~Krause). One of these targets is a core located in the \\object{Canis Majoris OB1/R1} region \\citep[e.g.,][]{2009A&A...506..711G} of the outer Galaxy at a distance of $\\sim 1$~kpc. Originally detected as a distinct submillimetre source in the vicinity of \\object{IRAS 07029-1215} \\citep{2004ApJ...602..843F}, it was coyly named ``UYSO~1'' (unidentified young stellar object 1) since it was only detected at 450 and 850 $\\mu$m and seemed to have no counterpart at other wavelengths. In the optical, the entire region is dominated by the strong H{\\sc ii} region \\object{Sh 2-297}, excited by the early B star \\object{HD 53623}. The mid-IR emission is dominated by warm dust and PAH emission from the associated southern photon-dominated region (PDR). All these sources, however, do not coincide with UYSO~1, located more than 1$'$ to the north-west. Further to the west, one finds the optical dark cloud \\object{LDN 1657A}, about which little is known. Members of our group carried out a multiwavelength study of \\object{UYSO 1} \\citep{2009A&A...493..547F}, which revealed more puzzles. The source drives a dynamically young, but strong CO molecular outflow. The imprints of two crossed jets are visible in shocked H$_2$ emission in the near-IR, and their vertex is very close to the position of UYSO~1. Millimeter continuum interferometry resolved the source into two peaks separated by 4\\farcs2. {\\it Spitzer}/MIPS data indicated that at 24 and 70 $\\mu$m, UYSO~1 is still too deeply embedded to be directly detected, a finding we revise in the present paper for 70 $\\mu$m with {\\it Herschel}. In addition to the core being chemically pristine and showing no signs of more evolved chemistry, this object is very interesting for the investigation of the early evolution of (intermediate-mass) protostars. ", "conclusions": "We have presented new {\\it Herschel} PACS and SPIRE scan map data for the star-forming complex containing the intermediate-mass core UYSO~1. They show the dust emission structures associated with this region in unprecedented detail. The high spatial resolution of the PACS 70 $\\mu$m map facilitates the differentiation of the true emission peak position from emission arising from the surrounding PDR material. The PACS data show that the 70 $\\mu$m emission is closely associated with the known location of the central millimetre peak(s) in UYSO~1. This revises an earlier finding of large offsets (based on {\\it Spitzer}/MIPS 70 $\\mu$m data). Hence, the measured {\\it Herschel} photometry can be used with more confidence than for previous data to estimate the SED of UYSO~1. Since the core was not clearly detected at 24 $\\mu$m (which otherwise would indicate a distinct component of warmer dust), a one-temperature modified black-body component with T = 28 K fits the data between 70\\,--\\,500 $\\mu$m very well. While clearly not in the hot-core/hot-corino regime, this temperature is higher than the typical values of $<$17--20 K found in other very young regions of star formation. The known outflow activity can be one explanation. However, the influence of the nearby PDR has also to be taken into account \\citep{2009A&A...493..547F}. The Herschel data reveal the serendipitous discovery of five compact very red sources, situated to the north-west of UYSO~1, within the neighbouring dark cloud LDN 1657A. These objects have lower temperatures than UYSO~1. A particularly intriguing object is LDN 1657A-4, which has a very red SED and exhibits a 24 $\\mu$m shadow. We note that two different methods for the column density estimation provided very consistent results. For both objects, the \\citet{1994A&A...291..943O} grain opacities seem more suitable than the use of ISM grains." }, "1005/1005.4474.txt": { "abstract": "%Studies of cosmic-ray accelerations and interactions in supernova remnants (SNRs) are critical step to solve the mystery of the cosmic-ray origins. %Observation of enhanced $\\pi^0$ gamma rays in the GeV--TeV band from SNRs interacting with molecular clouds is one of the promising approaches to confirm hadronic nature of cosmic rays accelerated in SNRs. We present detailed analysis of the two gamma-ray sources, \\FGLnorth\\ and \\FGLsouth, that have been found toward the supernova remnant~(SNR) W28 with the Large Area Telescope~(LAT) on board the \\emph{Fermi} Gamma-ray Space Telescope. \\FGLnorth\\ is found to be an extended source within the boundary of SNR W28, and to extensively overlap with the TeV gamma-ray source HESS~J1801$-$233, which is associated with a dense molecular cloud interacting with the supernova remnant. The gamma-ray spectrum measured with LAT from 0.2--100 GeV can be described by a broken power-law function with a break of $\\sim$~1~GeV, and photon indices of 2.09~$\\pm$~0.08~(stat)~$\\pm$~0.28~(sys) below the break and 2.74~$\\pm$~0.06~(stat)~$\\pm$~0.09~(sys) above the break. Given the clear association between HESS~J1801$-$233 and the shocked molecular cloud and a smoothly connected spectrum in the GeV--TeV band, we consider the origin of the gamma-ray emission in both GeV and TeV ranges to be the interaction between particles accelerated in the SNR and the molecular cloud. The decay of neutral pions produced in interactions between accelerated hadrons and dense molecular gas provide a reasonable explanation for the broadband gamma-ray spectrum. % \\FGLsouth\\, located outside the southern boundary of SNR W28, is consistent with a point source. \\FGLsouth\\, located outside the southern boundary of SNR W28, cannot be resolved. An upper limit on the size of the gamma-ray emission was estimated to be $\\sim$~16$'$ using events above $\\sim$~2~GeV under the assumption of a circular shape with uniform surface brightness. It appears to coincide with the TeV source HESS~J1800$-$240B, which is considered to be associated with a dense molecular cloud that contains the ultra compact \\hii\\ region W28A2 (G5.89$-$0.39). We found no significant gamma-ray emission in the LAT energy band at the positions of TeV sources HESS~J1800$-$230A and HESS~J1800$-$230C. %The LAT data for HESS~J1800$-$230A combined with the TeV data points %indicate a spectral break around 100~GeV. The LAT data for HESS~J1800$-$230A combined with the TeV data points indicate a spectral break between 10~GeV and 100~GeV. %Again, the gamma rays are naturally explained by $\\pi^0$ decays. %The W28A2 is a possible energy source although it requires extremely dense gas due to the energetics of cosmic rays. %Cosmic rays escaped from W28 are another plausible explanation given the energetics. %The SNR W28 is a prime target for observations of GeV gamma rays due to the detection of TeV gamma rays and its interaction with molecular clouds along its northeastern boundary and other clouds situated nearby. %We report observations of two GeV gamma-ray sources around W28 with the \\emph{Fermi} Gamma-ray Space Telescope~(FGST) - Large Area Telescope~(LAT) in the energy band between 200~MeV and 100~GeV. %We accumulated high energy gamma-ray events (200~MeV $<$ E $<$ 100~GeV) for $\\sim$~11 months. %Spatial extension and source position, and spectra were investigated by carrying out maximum likelihood fits to the data set. %One is spatially extended, and is positionally coincident with TeV gamma-ray source HESS~J1801$-$233 and also the molecular clouds interacting with this SNR at the northeast boundary of the SNR. %Decays of $\\pi^0$s produced by the interaction of hadrons accelerated in the SNR naturally explain the origin of gamma-ray emission in this source since electron models faces difficulties with high electron/proton ratio and energetics of cosmic rays constrained by the gamma-ray flux. %W28 is the most plausible energy source due to observational evidence of interaction with the molecular clouds, such as OH masers. %We found no significant gamma-ray emission in the LAT energy band at the positions of TeV sources, HESS~J1800$-$230A and HESS~J1800$-$230C, and obtained upper limits. %The upper limits for HESS~J1800$-$230A combined with the TeV data points indicate a spectral break around 100~GeV. %The energy-dependent escape of cosmic rays from the SNR and/or propagation effect of cosmic rays may produce such a break. %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 (~2 kpc) and extending up to ~4 % 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": "Diffusive shock acceleration operating at supernova shock waves can distribute particles to very high energies with a power-law form having number index about 2 \\citep[e.g.,][]{blandford87}. It is generally expected that if a dense molecular cloud is overtaken by a supernova blast wave, the shocked molecular cloud can be illuminated by relativistic particles accelerated at supernova shocks \\citep{Aharonian94}. If the accelerated particles are comprised mostly of protons, say $>100$ times more abundant than electrons like the observed Galactic cosmic rays, decays of neutral pions produced in inelastic collisions of the accelerated protons with dense gas are expected to be a dominant radiation component in the gamma-ray spectrum of the cosmic-ray-illuminated molecular cloud. Although an earlier attempt to detect TeV gamma-ray emission from supernova remnants (SNRs) that have evidence for molecular cloud interactions with the Whipple telescope failed \\citep{Whipple98}, two archetypical SNRs interacting with molecular clouds, IC~443 \\citep{Albert07, Acciari09} and W28 \\citep{Aharonian08}, have been detected with the current generation of imaging atmospheric Cherenkov telescopes. However, the identities of the particles responsible for the TeV sources remains elusive. The advent of the Large Area Telescope (LAT) onboard the \\emph{Fermi} Gamma-ray Space Telescope has brought a new opportunity to study the gamma-ray emission from SNRs at GeV energies. LAT observations of middle-aged SNRs interacting with molecular clouds, W51C~\\citep{LAT-W51C}, W44~\\citep{LAT-W44}, and IC~443~\\citep{LAT-IC443}, have revealed bright extended gamma-ray sources coincident with the SNRs. The gamma-ray luminosity reaches $\\sim 10^{36}\\ \\rm erg\\ s^{-1}$, which immediately rules out an inverse-Compton origin of the GeV gamma rays since it requires total electron energy comparable to or larger than the typical kinetic energy released by a supernova explosion, $\\sim 10^{51}$~erg. The gamma-ray spectra of the three remnants exhibit remarkable spectral breaks at an energy of several GeV, making these SNRs much less luminous at TeV energies. This characteristic demonstrates the importance of observations at GeV energies. %In this paper, detailed analysis of LAT data in a field that contains %an archetypical SNR-cloud system, W28, will be presented. W28 is a mixed-morphology SNR, characterized by center-filled thermal X-ray emission and shell-like radio morphology. In addition X-ray observations show limb-brightened shells in the northeast and southwest \\citep{Rho02}. The shell-like radio emission is prominent in the northeastern region with slightly fainter emission at the northern boundaries \\citep{Dubner00}. Interactions of the SNR with molecular clouds~\\citep{Wootten81} along its northern and northeastern boundaries are traced by the high concentration of 1720~MHz OH masers \\citep{Frail94, Claussen97,Claussen99}, and high density ($\\nh > 10^3$~cm$^{-3}$) shocked gas \\citep{Arikawa99, Reach05}. The overall shape of W28 is elliptical with a size of $50'\\times 45'$. W28 is located within a complex, star-forming region along the Galactic plane toward the large \\hii\\ regions (M8 and M20) and young clusters~(e.g., NGC~6530)~\\citep{goudis76}. The observations of molecular lines place SNR W28 at a distance of $\\sim 2$ kpc~\\citep{Velazquez02}. Estimates for its age vary between $3.5$ and $15\\times 10^4$~yrs \\citep{Kaspi93}. W28 is considered to be an evolved remnant in the radiative stage of SNR evolution \\citep{Lozinskaya92}, which is supported by optical observations \\citep{Lozinskaya74}. Measurements with Energetic Gamma-Ray Experiment Telescope (EGRET) onboard the \\emph{Compton Gamma-ray Observatory} found a gamma-ray source, 3EG~J1800-2338~\\citep{Hartman99} in the W28 field. However, its association to SNR W28 remained unclear mainly due to large source location uncertainties from EGRET. A gamma-ray source is listed in the W28 field in the AGILE (Astro-rivelatore Gamma a Immagini LEggero) one year catalog~\\citep{AGILEcatalog}. However, detailed analysis of this field is not published by AGILE yet. H.E.S.S. observations of the W28 field have revealed four TeV gamma-ray sources positionally coincident with molecular clouds \\citep{Aharonian08}: HESS~J1801$-$233, located along the northeastern boundary of W28, and a complex of sources, HESS~J1800$-$240A, B and C, located $\\sim 30\\arcmin$ south of SNR W28. HESS~J1801$-$233 coincides with a molecular cloud interacting with SNR W28, providing one of the best examples of a cosmic-ray-illuminated cloud. Understanding the origins of TeV emission in HESS~J1800$-$240ABC is of particular interest; they may be due to cosmic rays that have diffused from W28. In this paper, we report \\emph{Fermi} LAT observations of gamma-ray sources in the W28 field in the GeV domain. First, we give a brief description of the observation and gamma-ray selection in Section~\\ref{sec:obs}. The analysis procedures and results are explained in Section~\\ref{sec:ana}, where the spatial extension and spectra of the LAT sources in the W28 field are described. Discussion is given in Section~\\ref{sec:discuss}, followed by conclusions in Section~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} We have investigated two LAT sources in the W28 field. \\SourceN\\ (\\FGLnorth) which is located at the northeast boundary of the SNR W28 is positionally coincident with shocked molecular clouds, and is spatially extended. The spectrum has a break around 1.0~GeV and smoothly connects to the TeV spectrum, suggesting a physical relationship. % and a spectral curvature due to decays of $\\pi^0$s. } Decay of $\\pi^0$s produced by the interaction of an SNR with molecular clouds naturally explains the gamma rays from \\SourceN\\ based on the spatial correlation between GeV gamma rays and molecular clouds and the energetics of cosmic rays. Electron bremsstrahlung can not be ruled out completely although it requires a low density and low magnetic field in contradiction with the association with the molecular clouds. W28 is the most plausible energy source due to the observational evidence of interaction with the molecular clouds. The soft spectrum of the gamma rays may be explained by the time evolution of non-thermal emission from molecular clouds illuminated by cosmic rays from a nearby SNR due to energy-dependent diffusion of cosmic rays. %The \\SourceS\\ (\\FGLsouth) is consistent with the emission from a point source and spatially coincides with the TeV source HESS~J1800$-$240B, molecular clouds, and the ultracompact \\hii\\ region W28A2~(G5.89$-$0.39). The \\SourceS\\ (\\FGLsouth) was found to have no significant extension and spatially coincides with the TeV source HESS~J1800$-$240B, molecular clouds, and the ultracompact \\hii\\ region W28A2~(G5.89$-$0.39). %Again, leptonic models require high electron/proton ratio and the $\\pi^0$ decay is natural explanation of the origin of the gamma rays. The compact \\hii\\ region W28A2 and the SNR W28 are possible energy sources, but the W28A2 hypothesis requires extremely dense gas. %While no significant LAT counterpart is found at the positions of %HESS~J1800$-$230A and HESS~J1800$-$230C, the LAT upper limits for %HESS~J1800$-$230A coupled with the H.E.S.S. data points imply a %spectral break around 100~GeV. While no significant LAT counterpart is found at the positions of HESS~J1800$-$230A and HESS~J1800$-$230C, the LAT upper limits for HESS~J1800$-$230A coupled with the H.E.S.S. data points imply a spectral break between 10~GeV and 100~GeV. %The energy-dependent escape from SNR and/or propagation of cosmic rays may be responsible for this break. %our observations with the HESS \u00a1\u00f1?\u00a1\u00f1?-ray telescopes %have revealed VHE \u00a1\u00f1?\u00a1\u00f1?-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 %W28, while HESS J1800?240 situated just beyond the southern %boundary ofW28 comprises three components.Our studies with %NANTEN 12CO(J = 1?0) data show molecular clouds spanning %a broad range in local standard of rest velocity VLSR = 5 %to ?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 ?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 ?4 kpc distance (VLSR > 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 ?10 to ?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 VLSR > %10 km s?1 are at distances d ? 4 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 (e.g. 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 notW 28 and %other energetic sources have disrupted molecular material at line %velocities >10 km s?1." }, "1005/1005.4981_arXiv.txt": { "abstract": "The streaming instability (SI) provides a promising mechanism for planetesimal formation because of its ability to concentrate solids into dense clumps. The degree of clumping strongly depends on the height-integrated solid to gas mass ratio $Z$ in protoplanetary disks (PPDs). In this letter, we show that the magnitude of the radial pressure gradient (RPG) which drives the SI (characterized by $q\\equiv\\eta v_K/c_s$, where $\\eta v_K$ is the reduction of Keplerian velocity due to the RPG and $c_s$ is the sound speed) also strongly affects clumping. We present local two-dimensional hybrid numerical simulations of aerodynamically coupled particles and gas in the midplane of PPDs. Magnetic fields and particle self-gravity are ignored. We explore three different RPG values appropriate for typical PPDs: $q=0.025, 0.05$ and $0.1$. For each $q$ value, we consider four different particle size distributions ranging from sub-millimeter to meter sizes and run simulations with solid abundance from $Z=0.01$ up to $Z=0.07$. We find that a small RPG strongly promotes particle clumping in that: 1) At fixed particle size distribution, the critical solid abundance $Z_{\\rm crit}$ above which particle clumping occurs monotonically increases with $q$; 2) At fixed $Z$, strong clumping can occur for smaller particles when $q$ is smaller. Therefore, we expect planetesimals to form preferentially in regions of PPDs with a small RPG. ", "introduction": " ", "conclusions": "" }, "1005/1005.1089_arXiv.txt": { "abstract": "{} {We exploit deep observations of the GOODS-N field taken with PACS, the \\textit{Photodetector Array Camera and Spectrometer}, on board of \\textit{Herschel}, as part of the \\textit{PACS Evolutionary Probe} guaranteed time (PEP), to study the link between star formation and stellar mass in galaxies to $z\\sim 2$.} {Starting from a stellar mass -- selected sample of $\\sim4500$ galaxies with mag$_{4.5 \\mu m}<23.0$ (AB), we identify $\\sim350$ objects with a PACS detection at 100 or 160 $\\mu$m and $\\sim1500$ with only {\\textit Spitzer} 24 $\\mu$m counterpart. Stellar masses and total IR luminosities ($L_{IR}$) are estimated by fitting the Spectral Energy Distributions (SEDs). } {Consistently with other \\textit{Herschel} results, we find that $L_{IR}$ based only on 24 $\\mu$m data is overestimated by a median factor $\\sim1.8$ at $z\\sim2$, whereas it is underestimated (with our approach) up to a factor $\\sim1.6$ at $0.510^{11}M_{\\odot}$ galaxies from $z=0$ to $z=2$, and seems to flatten at $z>1.5$ in this mass range. Moreover, the most massive galaxies have the lowest SSFR at any $z$, implying that they have formed their stars earlier and more rapidly than their low mass counterparts ($downsizing$). } {} ", "introduction": "\\label{intro} The link between galaxy stellar mass and star formation rate (SFR), and its cosmic evolution is crucial to shed light on the processes of galaxy formation. The specific SFR (SSFR = SFR/mass) plays an important role as it measures the star formation efficiency of a galaxy and the fraction of a galaxy mass can be converted into stars per unit time (see e.g. de Cunha et al. 2010). Several studies at $0$20 kpc \\Ha\\ tail in the cluster. \\citet{Kenney2008} detected 120 kpc \\Ha\\ filaments from NGC 4438, a spiral galaxy in the Virgo cluster. \\citet{Yoshida2008} reported a galaxy with detached \\Ha\\ clouds with young blue stars in the Coma cluster, which resembles two galaxies in z$\\sim$0.2 clusters reported by \\citet{Cortese2007}. These \\Ha\\ cloud findings imply that gas removal mechanisms to form such extended emission line regions are common in clusters of galaxies. However, it is not yet clear how common this phenomenon is. The Coma cluster is the nearest rich cluster of galaxies, and is the best target for investigating the \\Ha\\ clouds in the cluster environment with a deep and spatially resolved observation. We have performed a deep \\Ha\\ narrow band imaging survey of the Coma Cluster using the Suprime-Cam at the Subaru Telescope and have already reported some of the prominent clouds detected \\citep{Yagi2007,Yoshida2008}. Here, we report significant, extended \\Ha\\ clouds around further 12 galaxies, increasing the sample to a total of 14. This is the first paper in a series providing a statistical discussion of extended emission line regions in the central region of the Coma cluster. The nature of the clouds and physical interpretations will be presented in subsequent papers. We assume that the distance modulus of the Coma cluster is $(m-M)_0=35.05$ \\citep{Kavelaars2000} and ($h_0$,$\\Omega_M$,$\\Omega_\\lambda$)=(0.71,0.27,0.73) \\citep{Larson2010}. Under these assumptions, the angular diameter distance is 97.47 Mpc, and 1 arcsec corresponds to 0.473 kpc. ", "conclusions": "\\subsection{Possible evolution sequence of the parent galaxies} The findings presented in Sections \\ref{sec:result:pos} and \\ref{sec:result:color} suggest that the parent galaxies were recently captured by the Coma cluster potential and are now infalling toward the cluster center, while their disk gas is stripped off and seen as \\Ha\\ clouds. The classification based on apparent morphology may imply an evolutionary sequence of gas stripping: a star-forming disk galaxy (\\#1) evolves into star-forming + dead disk phase (\\#2) and then the gas clouds are detached (\\#3). Recent numerical simulations of ram-pressure stripping \\citep{Kronberger2008,Kapferer2009,Tonnesen2010} also showed the evolution of \\Ha\\ appearance in $\\sim$ 500 Myr. However, it is puzzling that \\#3 phase galaxies consist of less luminous blue galaxies and bright red giants. Moreover, GMP2374, the brightest \\#1 galaxy, shows a red color. Although part of the red color is due to possible extinction by internal dust, the spectrum of the whole galaxy in the GOLDMINE database \\citep{Gavazzi2003} indicates that it is dominated by intermediate and old stellar populations as inferred from its large D4000 and small CaHK ratio. Therefore it is difficult to assume that this galaxy will evolve into a post-starburst phase with a blue color. The parents may consist of several kinds of galaxies on different evolutionary paths. As a second parameter of various evolutionary paths, we estimated the masses of the parents. Higher mass galaxies should have larger potential wells and retain some gas for star formation, while lower mass galaxies might simply evolve from type \\#1 to \\#3 without passing through \\#2. The masses of the parents were estimated by three different methods and are given in Table 4. \\citet{Bell2003} gives stellar mass-to-light ratios (M/L) as a function of color. They adopted a \"diet\" Salpeter initial mass function (IMF), which has 30\\% lower mass than Salpeter IMF \\citep{Salpeter1955}. We adopted \\[ \\log_{10} (M/L)_{\\rm i-band} = 0.006 + 1.114(r-i) - 0.15 \\] and calculated the stellar mass from the SDSS $r$ and $i$ magnitudes. We neglected K-correction for both the $i$ magnitude and $r-i$ colors, which were smaller than 0.1 mag. The possible effects of internal dust were also neglected. We also estimated the masses of the parents from the model SED fitting, using kcorrect v4.1.4 \\citep{Blanton2007} with SDSS and public GALEX magnitudes and redshifts as in \\citet{Miller2009}. The recommended corrections on the kcorrect web page \\footnote{http://howdy.physics.nyu.edu/index.php/Kcorrect} were applied, which included the correction between the AB and SDSS magnitudes, and the minimum uncertainty of the SDSS magnitude. The kcorrect used \\citet{Chabrier2003} IMF and Padova 1994 isochrones \\citep[and references therein]{Bruzual2003}. Yet another mass estimation was available in the MPA-JHU SDSS catalog \\footnote{http://www.mpa-garching.mpg.de/SDSS/DR7/}. The DR7\\_v5.2 version included 11 parents of this study. We adopted the ``MEDIAN'' column as the estimated mass. The IMF by \\citet{Kroupa2001} was adopted for the catalog\\citep{Kauffmann2003}. Although there were variations among the mass estimations of the three different methods, we can see that the type \\#3 dwarfs had less than $2 \\times 10^9$ M$_{\\sun}$. On the other hand, the classification between \\#1 and \\#2 showed no clear correlation with mass. In addition, the lowest mass \\#3 parents ($<2 \\times 10^9$ M$_{\\sun}$) stopped star formation quickly, while most of the \\#1 and \\#2 parents, the masses of which were $10^9-10^{11}$ M$_{\\sun}$, continued star formation. The correlation of the detachment of the clouds, the activity of star formation, and the mass implied that the quick quench of the star formation in the \\#3 blue dwarfs was due to shallow potentials and/or small amounts of gas. This result regarding the dependency of the star formation quench on the mass was similar to those in previous studies \\citep{Kauffmann2004,Haines2007,Wolf2009,Mahajan2010}, although this is the first time that galaxies of low mass ($<10^9$ M$_{\\sun}$) have been investigated in this way. \\subsection{Comparison with galaxies with UV asymmetries} Recently \\citet{Smith2010} investigated UV and \\Ha\\ images of the Coma cluster, and listed the galaxies with gaseous stripping events (GSE), which were detected as UV asymmetries and had colors with $NUV-i<4$. Their observed field overlapped with our field, and there were six galaxies in common (GMP2559, GMP2910, GMP3016, GMP3816, GMP4060, and GMP4232). It is striking that the UV asymmetric galaxies in our fields are only the six. This suggests that GSE galaxies have extended \\Ha\\ clouds with high probability. The other eight parents were not listed as GSE galaxies, although all of the parents in this study satisfied the selection criteria of \\citet{Smith2010} ($i<18$, and $0.01<$z$<0.043$). As the NUV-optical color is a function of the star formation strength and the age since the last star formation \\citep[e.g.][]{Kaviraj2007}, one possibility is that the non-GSE parents are older. This can explain the red \\#3 galaxies, as age quickly makes the $g-r$ color redder. The other five star-forming galaxies and a post-starburst galaxy have young stars, and the age cannot explain the non-detection of UV asymmetry. It is possible that the UV asymmetric feature is faint because the star formation is/was weaker than other cases. It is also possible that the UV emitting stars have a symmetric distribution, while the \\Ha\\ is asymmetric. The last case suggests that the distributions of the young stars and gas differ. The sign of the different distribution was shown in our previous studies \\citep{Yagi2007,Yoshida2008}. This also suggests that the source of ionization is not always the UV photons from young stars. A detailed discussion of the source of ionization will be presented in our next paper (Yoshida et al. in preparation). If we assume that all galaxies from which gas was removed experienced an \\Ha\\ parent phase, we can speculate on the evolution of the parents. With regard to infalling galaxies, \\citet{Graham2003} investigated two disk dwarf galaxies in the Coma cluster, which still show spiral arm patterns. The galaxies, GMP3292 and GMP3629, are thought to be transforming from spiral to passive galaxies. \\citet{Graham2003} also noted that these galaxies have large peculiar velocities relative to the Coma center. However, these two galaxies are, much redder than parents in this study ($r=16.5, g-r=0.68$ for GMP3292 and $r=18.3, g-r=0.64$ for GMP3629), but are of comparable mass (4.1$\\times 10^9$ and 8$\\times 10^8$ M$_{sun}$), with values estimated in the same way as for the parents. They might be recently evolved shapes from the parents, as they still retain the arm pattern." }, "1005/1005.4665_arXiv.txt": { "abstract": "The {\\it Chandra\\/} Source Catalog (CSC) is a general purpose virtual X-ray astrophysics facility that provides access to a carefully selected set of generally useful quantities for individual X-ray sources, and is designed to satisfy the needs of a broad-based group of scientists, including those who may be less familiar with astronomical data analysis in the X-ray regime. The first release of the CSC includes information about 94,676 distinct X-ray sources detected in a subset of public ACIS imaging observations from roughly the first eight years of the {\\it Chandra\\/} mission. This release of the catalog includes point and compact sources with observed spatial extents $\\lesssim 30''$. The catalog (1)~provides access to the best estimates of the X-ray source properties for detected sources, with good scientific fidelity, and directly supports scientific analysis using the individual source data; (2)~facilitates analysis of a wide range of statistical properties for classes of X-ray sources; and (3)~provides efficient access to calibrated observational data and ancillary data products for individual X-ray sources, so that users can perform detailed further analysis using existing tools. The catalog includes real X-ray sources detected with flux estimates that are at least 3 times their estimated $1\\,\\sigma$ uncertainties in at least one energy band, while maintaining the number of spurious sources at a level of $\\lesssim 1$ false source per field for a $100\\rm\\,ks$ observation. For each detected source, the CSC provides commonly tabulated quantities, including source position, extent, multi-band fluxes, hardness ratios, and variability statistics, derived from the observations in which the source is detected. In addition to these traditional catalog elements, for each X-ray source the CSC includes an extensive set of file-based data products that can be manipulated interactively, including source images, event lists, light curves, and spectra from each observation in which a source is detected. ", "introduction": "Ever since {\\it Uhuru\\/} \\citep{gia71}, X-ray astronomy missions have had a tradition of publishing catalogs of detected X-ray sources, and these catalogs have provided the fundamental datasets used by numerous studies aimed at characterizing the properties of the X-ray sky. While source catalogs are the primary data products from X-ray sky surveys \\citep[e.g.,][]{gia72, for78, elv92, vog93, vog99}, the {\\it Einstein\\/} IPC catalog \\citep{har90} demonstrated the utility of catalogs of {\\it serendipitous\\/} sources identified in the fields of {\\it pointed-observation\\/} X-ray missions. More recent serendipitous source catalogs \\citep[e.g.,][]{gio90, whi94, ued05, wat08} have further expanded the list of sources with X-ray data available for further analysis by the astronomical community. Source catalogs typically include a uniform reduction of the mission data. This provides a significant advantage for the general scientific community because it removes the need for end-users, who may be unfamiliar with the complexities of the particular mission and its instruments, to perform detailed reductions for each observation and detected source. When compared to all previous and current X-ray missions, the {\\it Chandra\\/} X-ray Observatory \\citep[e.g.,][]{wei00, wei02} breaks the resolution barrier with a sub-arcsecond on-axis point spread function (PSF). Launched in 1999, {\\it Chandra\\/} continues to provide a unique high spatial resolution view of the X-ray sky in the energy range from $0.1$ to $10\\,\\rm keV$, over a $\\sim\\!60$--$250$ square arcminute field of view. The combination of excellent spatial resolution, a reasonable field of view, and low instrumental background translate into a high detectable-source density, with low confusion and good astrometry. {\\it Chandra\\/} includes two instruments that record images of the X-ray sky. The Advanced CCD Imaging Spectrometer \\citep[ACIS;][]{bau98, gar03} instrument incorporates ten $1024\\times1024$ pixel CCD detectors (any six of which can be active at one time) with an effective pixel size of $\\sim\\!0.5''$ on the sky, an energy resolution of order $110\\,\\rm eV$ at the Al-K edge ($1.49\\,\\rm keV$), and a typical time resolution of $\\sim\\!3.2\\,\\rm s$. The High Resolution Camera \\citep[HRC;][]{mur00} instrument consists of a pair of large format micro-channel plate detectors with a pixel size $\\sim\\!0.13''$ on the sky and a time resolution of $\\sim\\!15.6\\ \\mu\\rm s$, but with minimal energy resolution. The wealth of information that can be extracted from identified serendipitous sources included in {\\it Chandra\\/} observations is a powerful and valuable resource for astronomy. \\begin{figure} \\epsscale{1.0} \\plotone{f1} \\caption{\\label{fig:sources} Distribution of CSC release 1.0 master sources on the sky, in Galactic coordinates.} \\end{figure} The aim of the \\dataset [ADS/Sa.CXO#CSC] {{\\it Chandra\\/} Source Catalog} (CSC) is to disseminate this wealth of information by characterizing the X-ray sky as seen by {\\it Chandra\\/}. While numerous other catalogs of X-ray sources detected by {\\it Chandra\\/} may be found in the literature \\citep[e.g.,][]{zez06, bra08, rom08, luo08, mun09, elv09}, the region of the sky or set of observations that comprise these catalogs is restricted, and they are typically aimed at maximizing specific scientific goals. In contrast, the CSC is intended to be an all-inclusive, uniformly processed dataset that can be utilized to address a wide range of scientific questions. The CSC is intended ultimately to comprise a definitive catalog of X-ray sources detected by {\\it Chandra\\/}, and is being made available to the astronomical community in a series of increments with increasing capability over the next several years. The first release of the CSC was published in 2009 March. This release includes information about 135,914 source detections, corresponding to 94,676 distinct X-ray sources on the sky, extracted from a {\\it subset\\/} of public imaging observations obtained using the ACIS instrument during the first eight years of the {\\it Chandra\\/} mission. The distribution of release~1 sources on the sky is presented in Figure~1. We expect that the CSC will be a highly valuable tool for many diverse scientific investigations. However, the catalog is constructed from pointed observations obtained using the {\\it Chandra\\/} X-ray Observatory, and is neither all-sky nor uniform in depth. The first release of the catalog includes only point and compact sources, with observed extents $\\lesssim 30''$. Because of the difficulties inherent in detecting highly extended sources and point and compact sources that lie close to them, and quantifying in a consistent and robust way the properties of such sources, we have chosen to exclude entire fields (or in some cases, individual ACIS CCDs) containing such sources from the first release of the CSC, as described in \\S~3.1. Therefore, the catalog does not include sources near some of the most famous {\\it Chandra\\/} targets, and there may be selection effects that restrict the source content of the catalog and which therefore may limit scientific studies that require unbiased source samples. The minimum flux significance threshold for a source to be included in the first release of the CSC is set conservatively, and corresponds typically to $\\sim\\!10$ detected source photons (on-axis) in the broad energy band integrated over the total exposure time. This conservative threshold was chosen to maintain the spurious source rate at an acceptable level over the wide variety of {\\it Chandra\\/} observations that are included in this release of the catalog. We expect to relax this criterion in future releases based on experience gained constructing the current release. A number of other {\\it Chandra\\/} catalogs do include sources with fewer net counts than the CSC\\null. Such fainter thresholds are attainable typically either because of specific attributes of the observations included in those catalogs, or because of the assumptions made when constructing the catalog. As an example of the former category, the XBootes survey catalog \\citep{ken05} includes sources that are roughly a factor of two fainter than the CSC flux significance threshold. That survey is constructed from short ($5\\rm\\,ks$) observations obtained in an area with low line-of-sight absorption. This results in a negligible background level that substantially simplifies source detection and enables identification of sources with very few counts. Some {\\it Chandra\\/} catalogs derived from observations with the range of exposures comparable to those that comprise the CSC \\citep[e.g.,][]{elv09, lai09, mun09} also include fainter sources. However, in these cases the additional source fractions are in general not large, typically adding $\\lesssim 10\\%$ more sources below the CSC threshold, as described in detail in \\S~3.7.1. For other {\\it Chandra\\/} catalogs, visual review and validation at the source level is a planned part of the processing thread \\citep[e.g.,][]{kim07, mun09}. In some cases \\citep[e.g.,][]{bro07}, visual review may be used to adjust processing parameters for individual sources. Such manual steps are time-consuming, but enable lower significance levels to be achieved while maintaining an acceptable spurious source rate. In contrast, the CSC catalog construction process requires that the processing pipelines run on a wide range of observations with a minimum of manual intervention. The scope of the CSC is simply too large to require manual handling at the source level. We do not manually inspect individual source detections, nor do we adjust source detection or processing parameters based on manual evaluation. Instead, the CSC uses a largely automated quality assurance approach, as described in \\S~3.14. The sky coverage of the first catalog release (Fig.~2) totals $\\sim\\!320$ square degrees, with coverage of $\\sim\\!310$ square degrees brighter than a $0.5$--$7.0\\rm\\,keV$ flux limit of $1.0\\times10^{-13}\\rm\\,erg\\,cm^{-2}\\,s^{-1}$, decreasing to $\\sim\\!135$ square degrees brighter than $1.0\\times10^{-14}\\rm\\,erg\\,cm^{-2}\\,s^{-1}$, and $\\sim\\!6$ square degrees brighter than $1.0\\times10^{-15}\\rm\\,erg\\,cm^{-2}\\,s^{-1}$. These numbers will continue to grow as the {\\it Chandra\\/} mission continues, with a 15 year prediction of the eventual sky coverage of the CSC of order 500 square degrees, or a little over 1\\% of the sky. \\begin{figure} \\epsscale{1.0} \\plotone{f2} \\caption{\\label{fig:skycov} Sky coverage of release 1.0 of the CSC, in the ACIS broad energy band. The ordinate value is the total sky area included in the CSC that is sensitive to point sources with fluxes at least as large as the corresponding value on the abscissa.} \\end{figure} In this paper we describe in detail the content and construction of release~1 of the CSC\\null. However, where appropriate we also discuss in addition the steps required to process HRC instrument data used to construct release~1.1 of the catalog, since the differences in the algorithms are small. Release~1.1 of the catalog is scheduled for spring 2010. This paper is organized into 5 sections, including the introduction. In \\S~2, we present a description of the catalog. This includes the catalog design goals, an outline of the general characteristics of {\\it Chandra\\/} data that are relevant to the catalog design, the organization of the data within the catalog, approaches to data access, and an outline of the data content of the catalog. Section~3, which comprises the bulk of the paper, describes in detail the methods used to extract the various source properties that are included in the catalog, with particular detail provided when the algorithms are new or have been adapted for use with {\\it Chandra\\/} data. A brief description of the principal statistical properties of the catalog sources is presented in \\S~4; this topic is treated comprehensively by \\citetfap. Conclusions are presented in \\S~5. Finally, Appendix~A contains details of the algorithm used to match source detection from multiple overlapping observations, as well as the mathematical derivation of the multivariate optimal weighting formalism used for combining source position and positional uncertainty estimates from multiple observations. ", "conclusions": "The {\\it Chandra\\/} Source Catalog is a general purpose virtual X-ray astrophysics facility that provides access to a carefully crafted set of scientifically useful quantities for individual X-ray sources observed by the {\\it Chandra\\/} X-ray Observatory. The first release of the catalog was published to the astronomical community in March 2009, and includes source properties for 94,676 point and compact X-ray sources detected in a subset of public ACIS imaging observations from roughly the first eight years of the {\\it Chandra\\/} mission. This release of the catalog includes sources with observed spatial extents are $\\lesssim 30''$, and whose flux estimates are at least 3 times their estimated $1\\,\\sigma$ uncertainties. Observations that include substantially extended sources are not included in the first release of the catalog. For each X-ray source, the catalog tabulates about 60 distinct measured and derived source properties, generally with associated lower and upper confidence limits, in several energy bands. These properties are generally derived from all of the observations in which a source is detected. However, in the first catalog release, multiple observations are not {\\it combined\\/} prior to source detection, so the depth of the catalog is limited by the duration of the longest single exposure of a field. The catalog further tabulates roughly 120 observation-specific properties for each observation of a source, again with associated lower and upper confidence limits, and in several energy bands. Tabulated source properties include source position, spatial extent, multi-band aperture fluxes computed in several different ways, X-ray hardness ratios and spectral model fits, and intra- and inter-observation variability measures. In addition to these ``traditional'' catalog elements, for each source detection the catalog includes an extensive set of FITS format file-based data products that can be manipulated interactively by the user, including source images, event lists, light curves, and spectra from each observation in which a source is detected. \\begin{figure*} \\epsscale{1.0} \\plotone{f26} \\caption{\\label{fig:phot} Comparison of input ($\\rm F_0$) and measured ($\\rm F$) ACIS broad ({\\it b\\/}) band fluxes for simulated sources with power-law spectra and off-axis angles $\\theta\\leq 10'$ ({\\it left\\/}) and $\\theta > 10'$ ({\\it right\\/}). For each bin, the horizontal line indicates the median measured flux value. The boxes include 90\\% of the measurements in each bin, and the vertical lines indicate the extreme values. Bins colored red include fewer than 100 measurements; bins colored blue include 100--400 measurements; bins colored black include more than 400 measurements. The green line has a slope of 1.\\vglue 5pt} \\end{figure*} Looking towards the future, release 1.1 of the catalog, scheduled for spring 2010, will include data from public HRC-I imaging observations and newly public ACIS imaging observations, but will otherwise retain the same limitations as release 1. In release 2, we plan to co-add multiple observations of the same field that use the same or similar instrument configurations, and that have similar spacecraft pointings (within $\\sim\\!30''$) prior to source detection, to achieve fainter limiting sensitivities in many fields. We anticipate that new algorithms will allow this release to have a significantly fainter source detection threshold than release 1. This release should also provide limited improvements in the area of extended source handling (for example allowing for the inclusion of exposures containing moderately extended emission from galaxy cores up to $\\sim\\!60''$ spatial scale), as well as numerous algorithm enhancements that will refine field and source property calculations." }, "1005/1005.5455_arXiv.txt": { "abstract": "The PICASSO project is using superheated droplets of C$_4$F$_{10}$ for the direct detection of Dark Matter candidates in the {\\it spin-dependent} (SD) sector. The total setup includes 32 detectors installed in the SNOLAB underground laboratory in Sudbury (Ontario, Canada). With a concentrated effort in detector purification and with new discrimination tools now available for analysis, Picasso published competitive results in June 2009 \\cite{publi2009} and became the leading experiment in the SD sector of direct dark matter searches. The present level of sensitivity is at 0.16 pb on protons at 90\\% C.L. (M$_W$= 24GeV/c$^2$) following an analysis of two detectors only. The rest of the detectors are now in the process of being analyzed and the experimental search continues in order to further improve the limits or hopefully discover a signal of dark matter. The status of the experiment and the ongoing analysis will be presented. ", "introduction": "The astronomical and cosmological observations strongly suggest the presence of Dark Matter and show that only 1\\% of the matter of the Universe is luminous and 23\\% of all the matter should be of a new exotic kind \\cite{bennett} : Cold Dark Matter (CDM). The preferred candidate for particle physicists is the neutralino, the Lightest Supersymmetric Particle (LSP). It is a very Weakly Interacting Massive Particle (WIMP) and is a natural candidate in the Minimal SuperSymmetric Model (MSSM) \\cite{supersym}. PICASSO (Project In CAnada to Search for Supersymmetric Objects) is one of the many worldwide efforts to hunt for dark matter through the direct detection of neutralinos via their SD interactions with $^{19}$F nuclei \\cite{ellis91,divari,bednyako}. The goal of the PICASSO project at SNOLAB is to exploit the favorable properties of $^{19}$F by using the superheated droplet detection technique, which is based on the operation principle of the classic bubble chamber \\cite{glaser52,NC94}. ", "conclusions": "PICASSO recently developed a new generation of detectors using a new kind of gel without cesium chloride which is presently the main source of alpha-contamination. In this case, viscous glycerin and polyethylene glycol allow to suspend the C$_4$F$_{10}$ droplets without density matching. Many improvements have been brought to the various fabrication steps, in order to obtain the lowest possible alpha-background. The preliminary results are very encouraging and may lead to a background reduction of at least a factor 10. Presently, half of the set-up has been replaced by these new detectors and the full installation is expected to be completed before the end of the year. The analysis of the new generation of detectors is now in progress and studies of a complete discrimination between the particle induced recoils and the alpha-particles will allow PICASSO to improve substantially the level of its sensitivity in the spin-dependant sector.\\\\" }, "1005/1005.2619_arXiv.txt": { "abstract": "The Hypervelocity Star survey presents the currently largest sample of radial velocity measurements of halo stars out to 80 kpc. We apply spherical Jeans modeling to these data in order to derive the mass profile of the Galaxy. We restrict the analysis to distances larger than 25 kpc from the Galactic center, where the density profile of halo stars is well approximated by a single power law with logarithmic slope between $-3.5$ and $-4.5$. With this restriction, we also avoid the complication of modeling a flattened Galactic disk. In the range $25 < r < 80$ kpc, the radial velocity dispersion declines remarkably little; a robust measure of its logarithmic slope is between $-0.05$ and $-0.1$. The circular velocity profile also declines remarkably little with radius. The allowed range of $V_c\\, (80\\, \\mathrm{kpc})$ lies between 175 and 231 $\\kms$, with the most likely value 193 $\\kms$. Compared with the value at the solar location, the Galactic circular velocity declines by less than 20\\% over an order of magnitude in radius. Such a flat profile requires a massive and extended dark matter halo. The mass enclosed within 80 kpc is $6.9^{+3.0}_{-1.2} \\times 10^{11}\\ \\Msun$. Our sample of radial velocities is large enough that the biggest uncertainty in the mass is not statistical but systematic, dominated by the density slope and anisotropy of the tracer population. Further progress requires modeling observed datasets within realistic simulations of galaxy formation. ", "introduction": "\\label{sec:intro} Measuring the mass of the Galaxy is an astonishingly difficult task. Convenient tracers of disk rotation -- stars and gas clouds -- extend only to 20 kpc \\citep[e.g.][]{sofue_etal09}. At larger radii, statistical analysis of radial velocities must be used. Traditional tracers at distances up to 100 kpc include globular clusters and dwarf satellite galaxies \\citep[e.g.,][]{kochanek96, wilkinson_evans99, sakamoto_etal03}. Recently, \\citet{battaglia_etal05} derived the radial velocity dispersion profile to 120 kpc using a combined sample of globular clusters, satellite galaxies, and halo red giant stars. They found the dispersion falling from $\\sim 120\\ \\kms$ to $\\sim 50\\ \\kms$ between 20 and 120 kpc. In contrast, \\citet{xue_etal08} assembled a large sample of blue horizontal-branch (BHB) stars from the Sloan Digital Sky Survey Data Release 6 (SDSS DR6) and found a much flatter profile between 20 and 60 kpc. Here we use a new spectroscopic survey \\citep{brown_etal10a} aimed at finding hypervelocity stars (HVS) to set the most precise constraint on the Galactic mass within 80 kpc. \\citet{brown_etal10a} present a sample of 910 late B-stars and early A-stars in the Galactic halo. Their luminosity, and therefore distance, depends on whether these stars are BHB stars or main-sequence blue stragglers with similar effective temperature and surface gravity. The ambiguous nature of the stars is especially problematic at redder colors, $u-g > 0.6$, where the luminosity can differ by a factor of 5, as demonstrated in Fig. 4 in \\citet{brown_etal10a}. This bimodal distribution of distance does not have a well-defined average value and therefore requires statistical sampling. \\citet{brown_etal10a} create 100 Monte Carlo realizations of each star to sample the color and metallicity distributions and to derive the distributions of luminosity and distance. We use this full Monte Carlo catalog of distances in our analysis, while retaining the observed values of radial velocity and its uncertainty. ", "conclusions": "Using maximum-likelihood analysis of a new sample of radial velocities of distant halo stars, we infer that their radial velocity dispersion profile declines little with distance from the Galactic center in the range $25 < r < 80$ kpc: $\\sigma(r) = 111 \\, (r/40\\, \\mathrm{kpc})^{-0.08}\\, \\kms$. Spherical Jeans modeling indicates that the circular velocity profile $V_c(r)$ also falls only slightly over the same radial range and reaches between 175 and 231 $\\kms$ at 80 kpc. The corresponding enclosed mass $M(80)$ is between $5.7\\times 10^{11}\\ \\Msun$ and $1.0\\times 10^{12}\\ \\Msun$. A three-component model for the baryon and dark matter mass distribution gives the total virial mass of the Galaxy $M_{\\rm vir} = (1.6 \\pm 0.3)\\times 10^{12}\\ \\Msun$ at the virial radius $R_{\\rm vir} = 300$ kpc. Our inferred mass of the Galaxy is higher than that obtained by \\citet{battaglia_etal05} ($M_{\\rm vir} \\approx 0.8\\times 10^{12}\\ \\Msun$) and \\citet{xue_etal08} ($M_{\\rm vir} \\approx 1.0\\times 10^{12}\\ \\Msun$) based on the modeling of their radial velocity datasets. Our HVS sample contains more objects at $r > 40$ kpc than the Battaglia et al. and Xue et al. datasets. Thus we have a stronger constraint on the shallow slope of the velocity dispersion profile and we derive a correspondingly larger mass. Our inferred mass is consistent with the larger scale measurement by \\citet{li_white08} based on the Andromeda-Milky Way timing argument, $M_{\\rm vir} \\approx 2.4\\times 10^{12}\\ \\Msun$. The implied dynamical mass-to-light ratio of the Galaxy, $M_{\\rm vir}/L_V \\approx 50$ in solar units, is also consistent with galaxy-galaxy weak lensing measurements by \\citet{mandelbaum_etal06}, galaxy kinematics modeling by \\citet{more_etal10}, and halo abundance matching modeling by \\citet{moster_etal10}. Our sample of radial velocities is large enough that the biggest uncertainty in the mass estimate is not statistical but systematic. Within the framework of spherical Jeans modeling, the uncertainty is dominated by the density slope and anisotropy of the tracer population. These parameters could be better constrained by future all-sky surveys of halo BHB stars. Deeper surveys that target more distant stars at $r \\gtrsim 100$ kpc would be similarly dominated by uncertainty over the underlying distribution of the tracers. The validity of spherical Jeans modeling is also limited by the presence of structure in the distribution of halo stars. Galactic stellar halo contains remnants of disrupted satellite galaxies, some of which are still detectable as tidal streams. Stars at $\\sim 100$ kpc from the Galactic center may not have had enough dynamical times to reach dynamical equilibrium, further limiting the application of equilibrium modeling. A first step in the direction of circumventing these systematics was taken by \\citet{xue_etal08}, who modeled the motion of particles in realistic simulated halos. Extension of such analysis to many different halo realizations using large samples of observed velocities may reduce the uncertainty over the global mass distribution in the Galaxy." }, "1005/1005.5039_arXiv.txt": { "abstract": "Here we show that in the case when double peaked emission lines originate from outer parts of accretion disk, their variability could be caused by perturbations in the disk emissivity. In order to test this hypothesis, we introduced a model of disk perturbing region in the form of a single bright spot (or flare) by a modification of the power law disk emissivity in appropriate way. The disk emission was then analyzed using numerical simulations based on ray-tracing method in Kerr metric and the corresponding simulated line profiles were obtained. We applied this model to the observed H$\\beta$ line profiles of 3C 390.3 (observed in the period 1995-1999), and estimated the parameters of both, accretion disk and perturbing region. Our results show that two large amplitude outbursts of the H$\\beta$ line observed in 3C 390.3 could be explained by successive occurrences of two bright spots on approaching side of the disk. These bright spots are either moving, originating in the inner regions of the disk and spiralling outwards by crossing small distances during the period of several years, or stationary. In both cases, their widths increase with time, indicating that they most likely decay. ", "introduction": "The huge amount of Active Galactic Nucleus (AGN) energy is released through accretion onto super-massive black hole (BH), supposed to exist in the center of AGN. The emission of the accretion disk is not only in the continuum, but also in the emission lines (e.g. in Fe K$\\alpha$ line) and in low ionization lines, as e.g. in broad Balmer emission lines which are seen as double peaked (DP). DP Balmer lines are found in 20\\% of radio loud AGN at $z < 0.4$ \\citep{eh94,eh03} and 4\\% of the Sloan digital Sky Survey (SDSS) quasars at $z < 0.33$ \\citep{st03}. Broad, double-peaked emission lines of AGN provide dynamical evidence for presence of an accretion disk feeding a supermassive black hole in the center of AGN. But in some cases, the variability of these lines shows certain irregularities which could not be explained just by standard model of accretion disk. The DP line profiles are often used to extract the disk parameters \\citep[see e.g.][]{ch89,ch90,eh94,eh03,st08,er09}. In a series of papers Dumont \\& Collin-Souffrin \\citep[see][and references therein]{cs87,csd90,dcs90a,dcs90b,dcs90c} investigated the radial structure and emission of the outer regions of the optically thin accretion disks in AGN and calculated detailed grid of photoionisation models in order to predict the relative strengths of low-ionization lines emitted from the disk. They obtained integrated line intensities and line profiles emitted at each radius of the disk, for its different physical parameters. They also studied the influence of the external illumination on the structure of the disk, considering the point source model, where a compact source of non-thermal radiation located at a given height illuminates the disk and the diffusion model, where the radiation of a central source is scattered back towards the disk by a hot diffusing medium. One of the first methods for calculating the profiles of optical emission lines from a relativistic accretion disk was proposed by \\citet{ch89}. The limitation to this method is that the accretion disk structure required to explain the variability of the line profiles cannot be axi-symmetric, i.e. very often the red peak is higher than blue one and that cannot be explained by this model. It is not possible in a circular disk, in which the blue peak is always Doppler boosted to be stronger than the red peak. Therefore, \\citet{er95} adapted the circular accretion disk model to elliptical disks in order to fit the profiles of double-peaked emitters with a red peak stronger than the blue one. This model introduced eccentricity and phase angle parameters to the circular model described above, and the pericenter distance of the elliptical orbits \\citep[see][]{er95}. Spectroscopic monitoring of double-peaked emitters \\citep[see e.g.][]{sh01,ge07,sh09} has revealed that a ubiquitous property of the double-peaked broad emission lines is variability of their profile shapes on the timescales of months to years. DP line profiles are observed to vary on timescales of months to years, i.e. on timescales of the order of the dynamical time or longer \\citep[e.g.][]{vz91,zh91,ma93,ro98,se00,sh01,sb03,ge07}. This slow, systematic variability of the line profile is on the timescale of dynamical changes in an accretion disk, and has been shown to be unrelated to the shorter timescale variability seen in the overall flux in the line, due to reverberation of the variable ionizing continuum. Patterns in the variability of the broad Balmer lines are often a gradual change and reversal of the relative strengths of the blue-shifted and red-shifted peaks \\citep[see e.g.][]{ne97}. Periodic variability of the red and blue peak strengths has also been attributed to a precessing elliptical disk, a precessing single-armed spiral (as e.g. 3C 332, 3C 390.3: \\citet{gi99}; NGC 1097: \\citet{sb03}), and a precessing warp in the disk. For instance, \\citet{wu08} computed the profiles of Balmer emission lines from a relativistic, warped accretion disk in order to explore the certain asymmetries in the double-peaked emission line profiles which cannot be explained by a circular Keplerian disk. Elliptical disks and spiral waves have been detected in cataclysmic variables \\citep{shh97,bc00}, and a radiation induced warp has been detected in the large-scale disk of the AGN NGC 4258 \\citep{mbp96}. Spiral waves are a physically desirable model since they can be produced by instability in the vicinity of a black hole. They can play an important role in accretion disks, because they provide a mechanism for transporting angular momentum outward in the disk, allowing the gas to flow inwards, towards the central black hole. Long-term profile variability is thus a useful tool for extracting information about the structure and dynamics of the accretion disk most likely producing the double-peaked emission lines. In this paper, we present an investigation of the disk line variations due to instability in accretion disk. First we developed a model, assuming that instability in the accretion disk affects disk emissivity. This model and some simulations of expected line profile variability are presented in \\S 2. In \\S 3 we compare the model with observations taken from long-term monitoring of 3C 390.3 \\citep{sh01} in order to obtain parameters of perturbations. In \\S 4 we discuss our results in the light of possible physical mechanisms which could cause such perturbations, and finally, in \\S 5 we outline our conclusions. ", "conclusions": "We developed a model of the disk perturbing region in the form of a single bright spot (or flare) by a modification of the power law disk emissivity and used this model to simulate the disk line profiles. This model has been used to fit the observed H$\\beta$ line of 3C 390.3 observed from 1995 to 1999. From this investigation we can point out the following results: \\begin{enumerate} \\item The model which includes perturbation (bright spot) in the accretion disk can successfully explain difference in double peaked line profiles, as e.g. higher red peak even if we have the standard circular disk. The position of a bright spot has a stronger influence on one particular part of spectral line profiles (such as e.g. its core if the spot is in the central part of the disk, or \"red\" and \"blue\" wings if the spot is located on receding and approaching part, respectively). \\item Using the model for perturbing region we were able to successfully model and reproduce the observed variations of the H$\\beta$ line profile in the case of 3C 390.3, including the two large amplitude outbursts observed during the analyzed period. Therefore, the observed variations of the 3C 390.3 H$\\beta$ line could be caused by perturbations in the disk emissivity. \\item We found that two outbursts referred by \\citet{sh01} could be explained by successive occurrences of two different bright spots on approaching side of the disk which are either moving, originating in the inner regions of the disk and spiralling outwards, or stationary. Both bright spots decay by time until they completely disappear. \\item Our results support hypothesis that the perturbations in accretion disk emissivity are probably caused by fragments in the spiral arms of the disk. \\end{enumerate} The results presented above show that a circular disk with perturbations (bright spots) can be applied to explain different double peaked line profiles, and can be also used to trace perturbations (as well as their characteristics) from the broad double peaked line shapes." }, "1005/1005.2934_arXiv.txt": { "abstract": "In this paper the possibility of generating large scale curvature perturbations induced from the entropic perturbations during the waterfall phase transition of standard hybrid inflation model is studied. % We show that whether or not appreciable amounts of large scale curvature perturbations are produced during the waterfall phase transition depend crucially on the competition between the classical and the quantum mechanical back-reactions to terminate inflation. If one considers only the classical evolution of the system we show that the highly blue-tilted entropy perturbations induce highly blue-tilted large scale curvature perturbations during the waterfall phase transition which dominate over the original adiabatic curvature perturbations. However, we show that the quantum back-reactions of the waterfall field inhomogeneities produced during the phase transition dominate completely over the classical back-reactions. The cumulative quantum back-reactions of very small scales tachyonic modes terminate inflation very efficiently and shut off the curvature perturbations evolution during the waterfall phase transition. This indicates that the standard hybrid inflation model is safe under large scale curvature perturbations during the waterfall phase transition. ", "introduction": "Inflation proved to be very successful both theoretically \\cite{Guth:1980zm} and observationally \\cite{Komatsu:2010fb} as a theory of early universe. The simplest models of inflation consist of a scalar field which is minimally coupled to gravity. A period of acceleration expansion is obtained if the potential is flat enough to allow for the inflaton field to slowly roll towards its minimum. With sufficient tunings in the parameters of the model, one can achieve 60 number of e-foldings or more to solve the horizon and the flatness problems of the standard cosmology. Hybrid inflation \\cite{Linde:1993cn, Copeland:1994vg} is an interesting model of inflation containing two scalar fields, the inflaton field and the waterfall field. In Linde's original hybrid inflation, the energy density during inflation is dominated by the vacuum while the inflaton is slowly rolling. The waterfall field is very heavy compared to the Hubble expansion rate during inflation, $H$, and it quickly rolls to its instantaneous minimum. The potential has the property that once the inflaton field reaches a critical value, $\\phi=\\phi_c$, the waterfall field becomes tachyonic triggering an instability and inflation ends quickly thereafter and the systems settles down into its global minimum. Usually it is assumed that the waterfall field does not play any role in curvature perturbations during inflation and during phase transition. In this picture one basically borrows the technics and the results of single field inflationary models. That is, the super-horizon curvature perturbations, once they leave the Hubble radius, are frozen and remain unchanged until they re-enter the Hubble radius at a later time, such as at the time of CMB decouplings. Here we would like to examine this picture more closely. We would like to see if hybrid inflation is safe under large scale curvature perturbations during the waterfall phase transition. If one considers only the classical evolution of the system, we show that during the phase transition the highly-blue tilted entropy perturbations can induce large blue-tilted curvature perturbations on super-horizon scales which can completely dominate over the original adiabatic curvature perturbations. However, we show that the quantum back-reactions of the waterfall field inhomogeneities produced during the phase transition become important before the classical back-reactions become relevant. We demonstrate that the cumulative quantum back-reactions of the short-wavelength inhomogeneities are so strong that they uplift the tachyonic instability of the entropy perturbations and the curvature perturbations freezes. Ideas similar to this line of thought, studying the amplifications of large scale curvature perturbations during preheating, were studied in \\cite{Taruya:1997iv}-\\cite{Kohri:2009ac} and more recently in \\cite{Levasseur:2010rk}. The paper is organized as follows. In section \\ref{hybrid} we review the basics of hybrid inflation and obtain the background evolutions of the inflaton and the waterfall fields. In section \\ref{entropy} the entropy perturbations and in section \\ref{curvature} their effects on curvature perturbations are studied. In Section \\ref{back} the classical non-linear back-reactions as well as the quantum mechanical back-reactions are calculated and are compared to each other. Brief conclusions and discussions are followed in section \\ref{conclusions}. While our work was finished the work by Lyth \\cite{Lyth:2010ch} appeared which has overlaps with our results. See also \\cite{Fonseca:2010nk} which appeared shortly after our work. ", "conclusions": "\\label{conclusions} In this paper the possibility of producing large scale curvature perturbations induced from the entropy perturbations during the waterfall phase transition in hybrid inflation are studied. We have shown that whether or not appreciable amounts of large scale curvature perturbations are produced depend crucially on the competition between classical and quantum mechanical back-reactions to terminate inflation. If one considers only the classical back-reaction effects, one obtains a significant large scale curvature perturbations which completely dominate over the initial curvature perturbations. The induced large scale curvature perturbations would be highly blue-tilted with $n_{\\cal R} \\simeq 4$ as in \\cite{Gong:2008ni}. However, we have shown that the quantum-mechanical back-reactions of the waterfall field inhomogeneities produced during the phase transition dominate before the classical-back-reaction becomes important. In the Hartree approximation we found that the quantum back-reactions shuts off the classical tachyonic instability very efficiently terminating inflation as well as curvature perturbations evolutions quickly after phase transition. We have shown that the main contribution to quantum back-reactions comes from the cumulation of the very small scales inhomogeneities, modes which are tachyonic during the phase transition but remain sub-horizon during entire inflationary period. We also made the interesting observation that in standard hybrid inflation where the waterfall field rapidly freezes to $\\psi=0$ at the background level, one can use $\\psi \\rightarrow \\sqrt{ \\langle \\delta \\psi^2 \\rangle} $ as the effective classical trajectory. In summary, the quantum back-reactions have two crucial effects. First they determine the end of inflation given by Eq. (\\ref{qtoc}) and { Eq.} (\\ref{eoi}). Second, as mentioned above, they provide the effective classical trajectory via $\\psi \\rightarrow \\sqrt{ \\langle \\delta \\psi^2 \\rangle} $ . Although we have presented the analysis here only for the standard hybrid inflation, but we believe that this picture also holds for other models of inflation where there are sharp phase transitions at the end of inflation. This includes models of brane inflation where inflation ends abruptly due to tachyon formation once the distance between the brane and anti-brane reaches a critical value. However, it would be interesting to see what happens in models of inflation, such as in double inflation \\cite{Silk:1986vc} where there is a mild phase transition in fields evolution during early stages of inflation. Since the difference $N_c - N_*$ in Eq. (\\ref{final}) is not very large, an appreciable amount of curvature perturbations can be created even when the quantum mechanical back-reactions are taken into account. This in turn can produce features in curvature perturbations such as in models \\cite{Joy:2007na, Battefeld:2008py, Battefeld:2010rf}. It would be interesting to see the observational effects of the phase transitions during inflation as considered e.g. in \\cite{Parkinson:2004yx, Joy:2008qd}. \\vspace{1 cm} {\\bf" }, "1005/1005.0100_arXiv.txt": { "abstract": "The Large Area Telescope on the {\\it Fermi} $\\gamma$-ray Space Telescope provides unprecedented sensitivity for all-sky monitoring of $\\gamma$-ray activity. It has detected a few Galactic sources, including 2 $\\gamma$-ray binaries and a microquasar. In addition, it is an adequate telescope to detect other transient sources. The observatory scans the entire sky every three hours and allows a general search for flaring activity on daily timescales. This search is conducted automatically as part of the ground processing of the data and allows a fast response to transient events, typically less than a day. Most of the outbursts detected are spatially associated with known blazars, but in several cases during the first years of observations, $\\gamma$-ray flares occurring near the Galactic plane did not reveal any initially compelling counterparts. This prompted follow-up observations in X-ray, optical, and radio to attempt to identify the origin of the emission and probe the possible existence of a class of transient $\\gamma$-ray sources in the Galaxy. Here we report on these LAT events and the results of the multiwavelength counterpart searches. ", "introduction": "There has been one firmly established class of variable sources in the high-energy $\\gamma$-ray sky. The Energetic $\\Gamma$-Ray Experiment Telescope ({\\it EGRET}) on the Compton $\\Gamma$-Ray Observatory discovered a population of variable $\\gamma$-ray blazars above 100 MeV\\cite{hartman:1999}. However, {\\it EGRET} also left the legacy of a large fraction of unidentified sources in the 3EG catalog. Many of these were found at low Galactic latitudes and believed to be Galactic in nature. {\\it EGRET} established $\\gamma$-ray pulsars as a Galactic population and these were thought to contribute to the unidentified sources. Several studies found indications of variability in some of the sources along the Galactic Plane (GP)\\cite{torres:2001,nolan:2003}, a behavior not expected of the pulsars, which are steady on these timescales. Additionally, no blazar counterpart was identified in several cases. This suggested the possible existence of a new Galactic $\\gamma$-ray class. Several sources showed both strong variability and a convincing lack of a blazar within the {\\it EGRET} localization errors, like 3EG\\,J0241+6103, 3EG\\,J1824-1514 and GRO\\,J1834-04. One of these has emerged as a new type of $\\gamma$-ray source with the rediscovery of 3EG\\,J0241+6103 (COS-B 2CG\\,135+01), which was associated, although the position was uncertain, with LSI\\,$+61 \\adeg 303$, a \"prominent radio flaring star system'', high mass X-ray binary (HMXB) system at 2 kpc constituted of a B0\\,Ve star and a neutron star orbiting on an excentric orbit with a 26.5 days period. A daily/monthly variability was seen with {\\it EGRET}, but no periodicity detected\\cite{tavani:1998}. A TeV source has then been detected by MAGIC and then VERITAS, at this position, with a periodic signal modulated at the orbital period. {\\it Fermi} has detected a $\\gamma$-ray source, 0FGL\\,J0240.36113, at the position RA=40.076, DEC=61.233 with a 95\\% error radius of $1.8 \\amin$, consistent with the optical counterpart, and with a periodicity of $26.6 \\pm 0.5$\\,days, the emission peaking at the periastron, and a spectrum reminiscent of the pulsars, therefore definitely identifying in the MeV-GeV domain the first $\\gamma$-ray binary source with the HMXB system\\cite{abdo:2009ApJ701L123}. The second interesting source is the case of 3EG\\,J1824-1514, detected by {\\it EGRET}, but without modulation, as spatially coincident with LS\\,5039, an HMXB constituted of a likely neutron star orbiting around an O6.5 star. HESS detected a periodic signal modulated at the orbital period of 3.91 days. {\\it Fermi} firmly detected at more than $12 \\sigma$ a periodic source, modulated at $3.91 \\pm 0.05$ days, in a complicated region with a very intense Galactic diffuse emission. The most dynamic example from {\\it EGRET} is GRO\\,J1838-04 (3EG\\,J1837-0423), which produced an intense outburst in June 1995\\cite{tavani:1997}. The flux above 100\\,MeV in a 3.5 day period was found to be a factor of 7 brighter than in later observations of the region. Notably, no blazar counterpart is known within the 99\\% {\\it EGRET} error contour. The absence of a flat spectrum radio source at the levels typical of the {\\it EGRET} blazars made this a candidate for a different type of $\\gamma$-ray emitter. The proximity of such a unique outburst to the inner Galaxy led to speculation of a possible Galactic origin. The question of the progenitor of this activity remains as well as the broader question of the existence of similar sources of this type. As stated in Ref.~\\refcite{tavani:1997}, ``other unidentified {\\it EGRET} sources near the GP appear to be time variable with V$>1.5$'' in Ref.~\\refcite{mclaughlin:1996ApJ473.763}. Finally, {\\it Fermi} has recently detected a variable high energy source coinciding with the position of the X-ray binary and microquasar Cygnus X-3, modulated at its short orbital period of 4.8 hours. Cygnus X-3 is an HMXB system located at a distance of $\\sim 7$\\,kpc, with a compact object of nature still matter in the debate, orbiting a Wolf-Rayet star\\cite{abdo:2009Science326.1512}. ", "conclusions": "The task of identifying counterparts for unidentified transients remains challenging. Although greatly reduced from {\\it EGRET}, the LAT error circles remain large compared to the resolution of telescopes in other wavebands. Additional arguments based on temporal and spectral characteristics are required to support firm associations. Ultimately, identifications of the Galactic transients require observations of related variability between the $\\gamma$-ray source and a candidate Galactic counterpart at lower frequency. In the absence of a detection of significant activity of the potential radio and X-ray counterparts for the LAT transients, they remain unidentified. These sources continue to be monitored regularly for $\\gamma$-ray activity as a part of the {\\it Fermi} sky survey observations." }, "1005/1005.3209_arXiv.txt": { "abstract": "If new physics were capable to push the neutrino-nucleon inelastic cross section three orders of magnitude beyond the standard-model (SM) prediction, then ultra-high energy (UHE) neutrinos would have already been observed at neutrino observatories. We use such a constraint to reveal information on the scale of noncommutativity (NC) $\\Lambda_{\\rm NC}$ in noncommutative gauge field theories (NCGFT) where neutrinos possess a tree-level coupling to photons in a generation-independent manner. In the energy range of interest ($10^{10}$ to $10^{11}$ GeV) the $\\theta$-expansion ($|\\theta| \\sim 1/\\Lambda_{\\rm NC}^2$) and therefore the perturbative expansion in terms of $\\Lambda_{\\rm NC}$ retains no longer its meaningful character, forcing us to resort to those NC field-theoretical frameworks involving the full $\\theta$-resummation. Our numerical analysis of the contribution to the process coming from the photon exchange, pins impeccably down a lower bound on $\\Lambda_{\\rm NC}$ to be as high as around up to 900 (450) TeV, depending on the estimates for the cosmogenic neutrino flux. If, on the other hand, one considers a surprising recent result occurred in Pierre Auger Observatory (PAO) data, that UHE cosmic rays are mainly composed of highly-ionized Fe nuclei, then our bounds get weaker, due to the diminished cosmic neutrino flux. Nevertheless, we show that even for the very high fraction of heavy nuclei in primary UHE cosmic rays, our method may still yield remarkable bounds on $\\Lambda_{\\rm NC}$, typically always above 200 TeV. Albeit, in this case one encounters a maximal value for the Fe fraction from which any useful information on $\\Lambda_{\\rm NC}$ can be drawn, delimiting thus the applicability of our method. ", "introduction": " ", "conclusions": "" }, "1005/1005.3453_arXiv.txt": { "abstract": "We present EIS/Hinode \\& SUMER/SoHO observations of propagating disturbances detected in coronal lines in inter-plume and plume regions of a polar coronal hole. The observation was carried out on $13^{th}$ November 2007 as JOP196/HOP045 programme. The SUMER spectroscopic observation gives the information about the fluctuation in radiance and on both resolved (Doppler shift) and unresolved (Doppler width) line-of-sight velocities whereas EIS $40\\arcsec$ wide slot images detect fluctuations only in radiance but maximizes the probability of overlapping field of view between the two instruments. From distance-time radiance maps, we detect the presence of propagating waves in a polar inter-plume region with a period of 15~min to 20~min and a propagation speed increasing from ($130~\\pm~14$)~km~s$^{-1}$\\ just above the limb, to ($330~\\pm~140$)~km~s$^{-1}$\\ around $160\\arcsec$\\ above the limb. These waves can be traced to originate from a bright region of the on-disk part of the coronal hole where the propagation speed is in the range of ($25~\\pm~1.3$)~km~s$^{-1}$ to ($38~\\pm~4.5$)~km~s$^{-1}$, with the same periodicity. These on-disk bright regions can be visualized as the base of the coronal funnels. The adjacent plume region also shows the presence of propagating disturbance with the same range of periodicity but with propagation speeds in the range of ($135~\\pm~18$)~km~s$^{-1}$ to ($165~\\pm~43$)~km~s$^{-1}$\\ only. A comparison between the distance-time radiance map of both regions, indicate that the waves within the plumes are not observable (may be getting dissipated) far off-limb whereas this is not the case in the inter-plume region. A correlation analysis was also performed to find out the time delay between the oscillations at several heights in the off-limb region, finding results consistent with those from the analysis of the distance-time maps. To our knowledge, this result provides first spectroscopic evidence of acceleration of propagating disturbances in the polar region close to the Sun (within 1.2~R/R$_{\\odot}$), which provides clues to the understanding of the origin of these waves. We suggest that the waves are likely either Alfv\\'{e}nic or fast magnetoacoustic in the inter-plume and slow magnetoacoustic in plume regions. This may lead to the conclusion that inter-plumes are preferred channel for the acceleration of the fast solar wind.\\\\ ", "introduction": "Coronal holes are regions of cool and low density plasma that, as such, are `dark' at coronal temperatures \\citep{1972ApJ...176..511M}. During solar minima, coronal holes are generally confined to the Sun's polar regions, while at solar maxima they can also be found at lower latitudes, usually associated with remnant active regions, as so-called `equatorial' coronal holes. The predominantly unipolar magnetic field from coronal hole regions is thought to give rise to the fast solar wind \\citep[e.g.,][]{1973SoPh...29..505K}. During solar minimum, Ulysses observations clearly show that the solar wind exhibits two modes of outflow: the fast wind, associated with polar coronal holes, with outflow speeds of $\\approx800$~km~s$^{-1}$\\ and the slow wind with outflow speeds of $\\approx400$~km~s$^{-1}$\\ associated with equatorial regions \\citep{1997GeoRL..24.2885W,2000JGR...10510419M}. However, during solar maxima, low latitude coronal holes also show faster than average solar wind speed upto $\\approx600$~km~s$^{-1}$ \\citep{2003JGRA..108.1144Z}. Extreme-ultraviolet images of polar coronal holes reveal the presence of diffuse, spike-like or sheet-like structures called plumes \\citep{1975spre.conf..651B, 1977SoPh...53..397A}, which subtend an angle of roughly $2\\degree$ relative to Sun center at low altitude and expands super-radially with the coronal hole \\citep{1997SoPh..175..393D}. Regions between these structures are termed as inter-plumes. From VUV spectroscopy, plumes are known to be denser and cooler than the surrounding inter-plume regions \\citep[e.g.,][]{2006A&A...455..697W}, while spectral lines are observed to be broader in inter-plumes \\citep[i.e.,][]{2000SoPh..194...43B,2000ApJ...531L..79G,2003ApJ...588..566T}. However differences in mass, momentum and energy flux in plumes and in inter-plumes are still not known precisely. There are several theoretical models which describe the role of MHD waves in the acceleration of the fast solar wind in coronal holes \\citep[see review by][and references therein]{2005SSRv..120...67O,2009LRSP....6....3C} and inter-plumes are often believed to be the primary site for this acceleration. It is further conjectured that these waves originate from the on-disk bright network regions \\citep{1997ApJ...484L..75W,2000A&A...359L...1P,2000ApJ...531L..79G,2001A&A...380L..39B}. A number of studies \\citep{1997ApJ...491L.111O, 2000ApJ...529..592O, 2001A&A...380L..39B, 2005A&A...442.1087P} have reported detection of oscillations in the off-limb regions of polar coronal holes. All of these studies point to the presence of compressional waves, thought to be slow magneto-acoustic waves \\citep{1998ApJ...501L.217D, 2006A&A...452.1059O, 2007A&A...463..713O, 2009A&A...499L..29B}. On the other hand, evidence for Alfv\\'{e}n waves propagating into the corona had been reported by \\citet{1998A&A...339..208B,2009A&A...501L..15B,2008A&A...483..271D,2009ApJ...691..794L} by studying the line width variations with height in polar coronal holes. Recent reports of detections of low-frequency ($<$ 5 mHz), propagating transverse motions in the solar corona \\citep{2007Sci...317.1192T} (from coronagraphic observation) and chromosphere \\citep{2007Sci...318.1574D} and their relationship with chromospheric spicules observed at the solar limb \\citep{2007PASJ...59S.655D} with the Solar Optical Telescope aboard Hinode \\citep{2007SoPh..243....3K} have widened interest in the subject. Recently \\citet{2009Sci...323.1582J} have reported detection of torsional Alfv\\'{e}nic motions associated with a large on-disk bright-point group. These waves are believed to be a promising candidate for the heating of the corona and acceleration of the solar wind \\citep{1971ApJ...168..509B,2005ApJ...632L..49S}. Furthermore, it has been suggested that the fast solar wind streams originate from coronal hole funnels and are launched by reconnection at network boundaries, \\citep{2005Sci...308..519T}. Measurements of the outflow speed in the extended corona have been obtained with the Ultraviolet Coronagraph Spectrometer (UVCS) aboard SoHO \\citep[e.g.,][]{2000SoPh..197..115A,2003ApJ...588..566T,2004A&A...416..749A, 2007A&A...472..299T}. Some of these studies concluded that plumes have lower outflow speeds than inter-plume regions \\citep{1997AdSpR..20.2219N,2000ApJ...531L..79G,2000A&A...353..749W,2000A&A...359L...1P,2003ApJ...588..566T,2007ApJ...658..643R} and, hence, may not contribute significantly to the fast solar wind, whereas some other theoretical and observational studies find higher outflow speeds in plumes than in inter-plume regions for at least some altitudes above the photosphere \\citep{1999JGR...104.9947C,2003ApJ...589..623G,2005ApJ...635L.185G}. These contradictory reports led to the debate on whether plumes or inter-plumes are the preferred source regions for the acceleration of the fast solar wind. This topic is highly debated and still open for further confirmation. Recently, \\citet{2009A&A...499L..29B} reported the detection of propagating slow magnetoacoustic waves with periods between 10~min and 30~min and speed $\\approx75$~km~s$^{-1}$ to $125$~km~s$^{-1}$\\ above the limb of a polar coronal hole. In their study, the propagating disturbances which are due to radiance perturbations are seen from the limb region up to $\\approx100\\arcsec$ above the limb. There is no discernible acceleration or deceleration of any individual feature as it propagates. In that study, the oscillations were detected in the two spectral lines of Ne~{\\sc viii}~770~\\AA\\ and Fe~{\\sc xii}~195~\\AA\\ observed with the Solar Ultraviolet Measurements Of Emitted Radiation \\citep[SUMER,][]{1995SoPh..162..189W} aboard the Solar and Heliospheric Observatory (SoHO) and with the EUV imaging spectrometer \\citep[EIS,][]{2007SoPh..243...19C} aboard Hinode, respectively. In this paper, we combine again the capabilities of SUMER and EIS to observe the on-disk, limb and far off-limb region of the coronal hole, to search for the origin of waves close to the Sun and study their propagating nature. The plan of the paper is as follows: in section~\\ref{sec:obs}, the observations acquired for this study and the data reduction techniques are outlined. In section~\\ref{sec:result} results of the present study are presented with the distance-time radiance map analysis, power series analysis and time-delay analysis. A discussion of the observational results and a comparison with similar results are taken up in section~\\ref{sec:discussion} and finally conclusions are drawn in section~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} Analysis of Ne~{\\sc viii} and Fe~{\\sc xii} radiance x-t maps reveal the presence of outward propagating radiance disturbances in the off-limb and near off-limb region of inter-plume with periodicities of about 15~min to 20~min. From the SUMER Ne~{\\sc viii} line radiance x-t map, one can infer that the waves originate from a bright location (presumably the footpoint of a coronal funnel) and propagate towards the limb with a speed ($25~\\pm~1.3$)~km~s$^{-1}$. Around the limb the speed has increased to ($38~\\pm~4.5$)~km~s$^{-1}$, reaching ($130~\\pm~51$)~km~s$^{-1}$ off-limb. Further far off-limb, the speed of the propagation becomes ($330~\\pm~140$)~km~s$^{-1}$ as seen in the EIS Fe~{\\sc xii} line. Similar propagating disturbances are also seen in the plume region but with negligible acceleration, if any. The waves are not visible far off-limb, suggesting that they may be dissipated or, more simply, merge into the background. The waves as recorded in the inter-plume regions are either Alfv\\'{e}nic or fast magnetoacoustic in nature whereas the one seen in plumes are more likely slow magnetoacoustic type. \\citet{2005Sci...308..519T} have conjectured that the solar wind outflow is launched by reconnection at network boundaries between open flux lines and intra-network closed loops. The intra-network closed loops are pushed by supergranular convection towards the network triggering reconnection. This scenario is consistent with our identification of the origin of the propagating disturbances in inter-plume region with an on-disk bright region. These results support the view that the inter-plume regions are the preferred channel for the acceleration of the fast solar wind." }, "1005/1005.3981_arXiv.txt": { "abstract": "The short-lived radioisotope $^{60}$Fe requires production in a core collapse supernova or AGB star immediately before its incorporation into the earliest solar system solids. Shock waves from a somewhat distant supernova, or a relatively nearby AGB star, have the right speeds to simultaneously trigger the collapse of a dense molecular cloud core and to inject shock wave material into the resulting protostar. A new set of FLASH2.5 adaptive mesh refinement hydrodynamical models shows that the injection efficiency depends sensitively on the assumed shock thickness and density. Supernova shock waves appear to be thin enough to inject the amount of shock wave material necessary to match the short-lived radioisotope abundances measured for primitive meteorites. Planetary nebula shock waves from AGB stars, however, appear to be too thick to achieve the required injection efficiencies. These models imply that a supernova pulled the trigger that led to the formation of our solar system. ", "introduction": "Primitive meteorites contain daughter products of the decay of short-lived radioisotopes (SLRIs) such as $^{26}$Al, $^{41}$Ca, $^{53}$Mn, and $^{60}$Fe, distributed in different minerals in a way that indicates the parent isotopes were still alive at the time of their incorporation into the refractory inclusions and chondrules that record the earliest history of the solar system. The presence of $^{60}$Fe is particularly significant, as its production requires stellar nucleosynthesis (Tachibana \\& Huss 2003; Tachibana et al. 2006). Given half-lives on the order of $\\sim 10^6$ yr, the evidence for these radioisotopes suggests that the same stellar source that synthesized them may well have triggered the collapse of the presolar dense cloud core as well, while simultaneously injecting the freshly-synthesized radioisotopes (Cameron \\& Truran 1977; Boss 1995). Supernovae resulting from massive stars in the range of $\\sim 20 M_\\odot$ to $\\sim 60 M_\\odot$ or planetary nebulae derived from intermediate-mass ($\\sim 5 M_\\odot$) AGB stars have been proposed as possible sources of all or most of these radioisotopes (e.g., Huss et al. 2009). Shock-triggered collapse and injection into the presolar cloud (Cameron \\& Truran 1977) has been proposed and studied in detail (e.g., Boss 1995; Foster \\& Boss 1997; Vanhala \\& Boss 2002; Boss et al. 2008, 2010). Recent calculations have shown that simultaneous triggered gravitational collapse and injection of shock wave gas and dust into the collapsing cloud core is possible even when detailed heating and cooling processes in the shock-cloud interaction are included (Boss et al. 2008). Shock speeds in the range from 5 km/sec to 70 km/sec are capable of achieving simultaneous triggering and injection for a 2.2 $M_\\odot$ target cloud (Boss et al. 2010). However, these models led to considerably lower injection efficiencies than those previously estimated on the basis of models where the shock-cloud interaction was assumed to be isothermal (Boss 1995; Foster \\& Boss 1997; Vanhala \\& Boss 2002). When the injection efficiency ($f_i$) is defined to be the fraction of the incident shock wave material that is injected into the collapsing cloud core, values of $f_i \\sim 0.001$ result from the nonisothermal models (Boss et al. 2008, 2010), about 100 times lower than the values of $f_i$ found previously for strictly isothermal interactions. Considering that the shock fronts in these models contain 0.015 $M_\\odot$ of gas and dust, this means that the Boss et al. (2008, 2010) models produced nominal dilution factors $D \\sim 10^{-5}$, where $D$ is defined as the ratio of the amount of mass derived from the stellar source of the shock front that ends up in the protoplanetary disk to the amount of mass in the disk that did not derive from the stellar source. Such values appear to be much too low to explain the initial abundances inferred for typical SLRIs, which range from $\\sim 10^{-4}$ to $\\sim 3 \\times 10^{-3}$ for supernovae (Takigawa et al. 2008; Gaidos et al. 2009) and $\\sim 3 \\times 10^{-3}$ for an AGB star (Trigo-Rodr\\'iguez et al. 2009). Boss et al. (2010) found that varying the shock speed from 5 to 70 km/sec had relatively little effect on $f_i$, while doubling the density of the target cloud could decrease $f_i$ by a factor of 3. Here we explore the effects of changes in the assumed shock wave parameters, in order to learn if higher values of $f_i$ and therefore $D$ might thereby result. In addition, we seek to learn if these shock wave variations will indicate a preference for either a supernova or an AGB star wind for triggering the formation of the solar system. ", "conclusions": "A new set of models with varied shock densities and thicknesses has shown that injection efficiencies $f_i$ and dilution factors $D$ can be increased by large factors ($>$ 10 and $>$ 1000, respectively), large enough to maintain the viability of this SLRI injection mechanism. Observations of supernova remnants and planetary nebulae imply that while the former shock fronts are thin enough to be suitable for SLRI injection, the latter are not. These results lend support to previous studies that have favored a supernova over an AGB star for the source of the solar system's SLRIs. Huss et al. (2009) found that an intermediate-mass AGB star could explain the production of $^{26}$Al, $^{41}$Ca, $^{60}$Fe, but not that of $^{53}$Mn. Kastner \\& Myers (1994) pointed out that AGB stars are seldom found in the vicinity of star-forming regions, so the chances of SLRI injection from a planetary nebula wind into a dense cloud core are small. The culprit appears to have been a long-forgotten supernova remnant that swept through the galaxy $\\sim$ 4.56 Gyr ago." }, "1005/1005.3386_arXiv.txt": { "abstract": "The source HESS J1813-178 was detected in the survey of the inner Galaxy in TeV $\\gamma$-rays, and a composite supernova remnant (SNR) G12.8-0.0 was identified in the radio band to be associated with it. The pulsar wind nebula (PWN) embedded in the SNR is powered by an energetic pulsar PSR J1813-1749, which was recently discovered. Whether the TeV $\\gamma$-rays originate from the SNR shell or the PWN is uncertain now. We investigate theoretically the multiwavelength nonthermal radiation from the composite SNR G12.8-0.0. The emission from the particles accelerated in the SNR shell is calculated based on a semianalytical method to the nonlinear diffusive shock acceleration mechanism. In the model, the magnetic field is self-generated via resonant streaming instability, and the dynamical reaction of the field on the shock {\\bf is} taken into account. Based on a model which couples the dynamical and radiative evolution of a PWN in a non-radiative SNR, the dynamics and the multi-band emission of the PWN are investigated. The particles are injected with a spectrum of a relativistic Maxwellian plus a power law high-energy tail with an index of $-2.5$. Our results indicate that the radio emission from the shell can be well reproduced as synchrotron radiation of the electrons accelerated by the SNR shock; with an ISM number density of $1.4$ cm$^{-3}$ for the remnant, the $\\gamma$-ray emission from the SNR shell is insignificant, and the observed X-rays and very high energy (VHE) $\\gamma$-rays from the source are consistent with the emission produced by electrons/positrons injected in the PWN via synchrotron radiation and inverse Compton (IC) scattering, respectively; the resulting $\\gamma$-ray flux for the shell is comparable to the detected one only with a relatively larger density of about $2.8$ cm$^{-3}$. The VHE $\\gamma$-rays of HESS J1813-178 can be naturally explained to mainly originate from the nebula although the contribution of the SNR shell becomes significant with a denser ambient medium. ", "introduction": "\\label{sec:intro} The VHE source HESS J1813-178 was discovered with the High Energy Stereoscopic System (HESS) in a survey of the inner Galaxy in VHE $\\gamma$-rays \\citep[][]{A05}. The VHE $\\gamma$-ray image obtained with the HESS shows a pointlike source with an extension of $\\sim2.2'$, and the observed spectrum has a hard photon index $2.09\\pm0.08$ \\citep[][]{A06}. HESS J1813-178 was also detected in VHE $\\gamma$-rays with the Major Atmospheric Gamma Imaging Cerenkov (MAGIC) telescope. The differential flux given by MAGIC between 0.4 and 10 TeV can be well described by a power law with a index of $-2.1\\pm0.2$ \\citep[][]{Al06}, which is consistent with the result obtained with the HESS. Firstly, HESS J1813-178 was unidentified, and it was assumed to be a \"dark particle accelerator\" since no counterpart at lower frequencies was reported ever. However, the shell of the SNR G12.8-0.0 associated with HESS J1813-178 was discovered with a diameter of $\\sim2.5'$ in a new low-frequency Very Large Array (VLA) 90 cm survey \\citep[][]{Bet05}. The non-thermal radio flux densities of the shell-type SNR are $0.65\\pm0.10$ and $1.2\\pm0.08$ Jy at 20 and 90 cm, respectively \\citep[][]{Bet05}. A highly absorbed X-ray source AX J1813-178, for which the column density is $10^{23}$ cm$^{-2}$, detected with {\\it ASCA} is spatially coincident with the SNR. The X-ray emission extending to 10 keV with a sharp cutoff below 2 keV is primarily nonthermal, which can originate either from the SNR shell or form the pulsar wind nebula (PWN) inside the remnant \\citep[][]{Bet05}. Due to the high column density derived from the {\\it ASCA} data, a distance of $\\geq4$ kpc is derived for the source \\citep[][]{Bet05}. Moreover, a soft $\\gamma$-ray source, IGR J18135-1751, was discovered as the counterpart of HESS J1813-178 \\citep[][]{Uet05}. It is persistent with a 20--100 keV luminosity of $5\\times10^{34}$ erg s$^{-1}$ for a distance of 4 kpc. \\citet[][]{Uet05} argued that the observed properties of the source in the radio and X-ray bands can be explained with the assumption that the source is a pulsar wind nebula embedded in G12.8-0.0. Furthermore, high-angular resolution X-ray observation with {\\it XMM-Newton} shows that the X-ray emitting object appears as a compact core located in the center of the radio shell-type SNR G12.8-0.0 \\citep[][]{Fet07}. They argued that the source is a composite SNR since the central object shows morphological and spectral resemblance to a PWN. The observation with {\\it Chandra} on the SNR G12.8-0.0 indicates the X-ray source is a point surrounded by structured diffuse emission that fills the interior of the radio shell \\citep[][]{Het07}. The compact source has a spectrum characterized by a power law with an index of $\\sim1.3$, typical of young and energetic rotation-powered pulsars, and the morphology of the diffuse emission strongly resembles that of a pulsar wind nebula \\citep[][]{Het07}. Recently, an energetic pulsar PSR J1813-1749 with a period of $\\sim44.7$ ms, a characteristic age of 3.3 - 7.5 kyr, and a distance of $4.7$ kpc by assuming the association with an adjacent young stellar cluster, was discovered in a long, continuous {\\it XMM-Newton} X-ray timing observation \\citep[][]{GH09}. The pulsar was found to be associated with the SNR G12.8-0.0, and it powers the PWN \\citep[][]{GH09}. High-energy $\\gamma$-rays can be produced either from SNR shells in which particles are accelerated to relativistic through the first Fermi process \\citep[e.g.,][]{A05b,A07,BV06,FZ08,Fet09,MAB09}, or from PWNe powered by the pulsars inside them \\citep[e.g.,][]{Vet08, ZCF08,GSZ09}. Although the source HESS J1813-178 is pointlike in VHE $\\gamma$-rays, the possibility of the VHE $\\gamma$-rays originating from the shell of the remnant cannot be ruled out given the size of the SNR, the angular resolution of the HESS telescope, and the depth of the observations \\citep[][]{Al06}. In this paper, we study the multiband nonthermal emission from the shell of the SNR G12.8-0.0 and the nebula inside it. The emission from the particles accelerated by the SNR shock is investigated based on a semianalytical method to the nonlinear diffusive shock acceleration mechanism with a free escape boundary proposed by \\citet{Cea10fre}, in which the amplified magnetic field due to resonant streaming instability induced by cosmic rays and the dynamical feedback of this self-generated magnetic field on the shock are taken into account. On the other hand, the dynamics and the multi-band radiative properties of the PWN are investigated basically according to the model in \\citet[][]{GSZ09}, which can self-consistently describe the dynamical and radiative evolution of a pulsar wind nebula in a non-radiative supernova remnant. Recently, based on the long-term two dimensional particle-in-cell simulations, \\citet[][]{Sp08} found that the particle spectrum downstream of a relativistic shock consists of two components: a relativistic Maxwellian and a power law high-energy tail with an index of $-2.4\\pm0.1$. Different from \\citet{GSZ09}, in which a single power law injection spectrum for the electrons/positrons is employed to discuss the radiative properties during different phase of the PWN, in this paper we argue that the high-energy particles are injected with the spectrum of a relativistic Maxwellian plus a power-law high-energy tail during the evolution, and a kinetic equation is used to obtain the energy distribution of the particles. ", "conclusions": "\\label{sec:discussion} The origin of the VHE $\\gamma$-rays from HESS J1813-178 is investigated in this paper. Although the source is pointlike in VHE $\\gamma$-rays, a shell origin of the VHE $\\gamma$-rays cannot be ruled out because the PSF of the HESS telescope is very close to the size of the SNR's shell \\citep[e.g.,][]{Al06}. A SNR with a diameter of $2.5'$ was identified in the radio bands to be associated with the VHE $\\gamma$-ray source. Moreover, X-ray observations showed a PWN powered by an energetic pulsar is embedded in the SNR \\citep[][]{Bet05,Uet05,Fet07,Het07}, and the pulsar was recently discovered in a long, continuous {\\it XMM-Newton} X-ray timing observation \\citep[][]{GH09}. We apply a model which can self-consistently calculate the multiwavelength nonthermal emission both from the SNR shell and from the PWN embedded in the remnant. In the model, electrons/positrons with a spectrum of a relativistic Maxwellian plus a power-law high-energy tail are injected in the nebula during the PWN during its evolution inside the SNR; protons and electrons are accelerated by the SNR shock wave, and the spectrum of the accelerated particles are calculated with a semi-analytical non-linear model. Our results indicate that: with the parameters in Table\\ref{para}, (1) the observed emission of the shell in the radio bands can be well explained as synchrotron radiation of the electrons accelerated by the SNR shock wave, whereas the flux of p-p collisions of the accelerated protons is significantly smaller than the TeV flux observed with HESS; (2) the observed emission in the X-ray to hard X-ray band detected with {\\it XMM-Newton} and {\\it INTEGRAL}, respectively, can be explained as the synchrotron radiation of the electrons/positrons injected in the PWN; (3) the VHE $\\gamma$-rays in the TeV band are mainly produced via IC scattering of the electrons/positrons injected in the nebula on the CMB and interstellar IR photons. With a sub-energetic explosion ($E_{\\rm sn}\\sim 0.4\\times 10^{50}$ erg), an age of $\\sim 1200$ yr, and an ambient density of $1.4$ cm$^{-3}$, the radius of the composite SNR G12.8-0.0, 1.7 pc, as well as the multiband observed nonthermal fluxes for the remnant can be reproduced within this model described in this paper. The present value of the SNR shock's velocity is about 1000 km s$^{-1}$, and the VHE $\\gamma$-rays are predominately produced from the nebula inside the remnant via IC scattering. Of course, p-p interaction from the SNR shell can be enhanced with a denser medium around the remnant (Fig.\\ref{figs6}). Even if this condition were satisfied, the $\\gamma$-rays produced via IC scattering in the PWN remain significant, hence, in this case, both the SNR shell and the PWN would contribute to the production of the observed $\\gamma$-ray flux. In GeV $\\gamma$-rays, the spectral property of the resulting emission in this SNR-dominated scenario (Fig.\\ref{figs6}) differs significantly from that in the PWN-dominated case (Fig.\\ref{g128}). The VHE $\\gamma$-rays of HESS J1813-178 detected with HESS can be naturally explained as the IC scattering of the electrons/positrons injected into the PWN although the p-p collisions become important with a denser ambient medium for the remnant. Our study give more insights on the nature of the multiband nonthermal emission of the composite SNR G12.8-0.0, even though some assumptions are made in the model." }, "1005/1005.1660_arXiv.txt": { "abstract": "Turbulent, two-dimensional, hydrodynamic flows are characterized by the emergence of coherent, long-lived vortices without a need to invoke special initial conditions. Vortices have the ability to sequester particles, with typical radii $\\sim 1$ mm to $\\sim 10$ cm, that are slightly decoupled from the gas. A generic feature of discs with surface density and effective temperature profiles that are decreasing, power-law functions of radial distance is that four vortex zones exist for a fixed particle size. In particular, two of the zones form an annulus at intermediate radial distances within which small particles reside. Particle capture by vortices occurs on a dynamical time scale near and at the boundaries of this annulus. As the disc ages and the particles grow via coagulation, the size of the annulus shrinks. Older discs prefer to capture smaller particles because the gas surface density decreases with time, a phenomenon we term ``vortex aging''. More viscous, more dust-opaque and/or less massive discs can have vortices that age faster and trap a broader range of particle sizes throughout the lifetime of the disc. Thus, how efficiently a disc retains its mass in solids depends on the relative time scales between coagulation and vortex aging. If vortices form in protoplanetary discs, they are important in discs with typical masses and for particles that are likely to condense out of the protostellar nebula. Particle capture also occurs at distances relevant to planet formation. Future infrared, submillimetre and centimetre observations of grain opacity as a function of radial distance will test the hypothesis that vortices serve as nurseries for particle growth in protoplanetary discs. ", "introduction": "The formation of $\\sim$ km-sized planetesimals from sub-micron-sized dust grains likely involves more than one physical process \\citep{youdin10,cy10}. It is generally accepted that particle growth in dusty, circumstellar discs is hierarchical, eventually forming planetesimals that are the building blocks of planets \\citep{armitage07,armitage10}. There is reasonable understanding of building small particles, but forming planetesimals from these particles remains mired in controversy. The initial stage of growth probably proceeds through the nucleation of sub-micron-sized dust grains from the primordial nebula, which then form the monomers of fractal dust aggregates up to $\\sim 1$ mm to $\\sim 10$ cm sizes in $\\gtrsim 10^3$ yrs, beyond which growth is stalled by collisional bouncing and fragmentation \\citep{bw08,zsom10}. In this regime, the particle dynamics and coagulation are described by Brownian motion and van der Waals forces. One of the best astrophysical pieces of evidence for grain growth to these sizes is the detection of 3.5 cm dust emission from the classical T Tauri star TW Hya (age $\\sim 5$--10 Myr), located 56 pc away, which has a face-on circumstellar disc of radius 225 AU \\citep{wilner05}. In standard models of protoplanetary discs, the gas pressure decreases radially outward. Gas in the disc then moves at sub-Keplerian speeds. Solid particles on Keplerian orbits experience a ``head wind'' with a velocity $\\sim 10^3$ cm s$^{-1}$ --- this head wind drags $\\sim 10$ cm- to $\\sim 1$ m-sized particles into the central star on time scales of $\\sim 10$--100 years \\citep{weiden77a}. These time scales are much shorter than the characteristic time scale for these particles to collide and grow into larger particles that are unaffected by the head wind. \\cite{safronov69} and \\cite{gw73} suggested that this difficulty can be circumvented by dust settling into the midplane of the disc and triggering gravitational instability, but \\cite{weiden80} pointed out that turbulence generated by the settling impedes the process. Long-lived structures in the gas are a possible way of concentrating particles with sizes $\\sim 1$ mm to $\\sim 10$ cm and growing them to larger sizes. Such structures are usually high-pressure regions in the gas that are capable of concentrating particles, which is a manifestation of Bernoulli's principle\\footnote{Bernoulli's principle states that in inviscid flows, a decrease in fluid velocity is accompanied by an increase in pressure. Hence, locations of maximum pressure are also locations of minimum velocity.} \\citep{kundu04}. Vortices are examples of long-lived structures --- they are spiral, non-linear motions of fluid with closed streamlines. On Earth, ocean vortices have been observed to trap larval fish off the coast of western Australia \\citep{paterson08}. In astrophysical settings, the possible role of vortices in planet formation was suggested by \\cite{vw46}, based on the writings of Kant, in an article entitled {\\it Die Entstehung des Planetensystems} (``The Origin of Planetary Systems''). Since then, vortices have been suggested as possible nurseries for growth to $\\gtrsim 1$ m-sized particles \\citep{aw95,bs95,tanga96,bracco99,gl99,gl00,johansen04,bm05,fn05,kb06,inaba06,shen06,bodo07,mr09}. In stratified protoplanetary discs, there are conceivable locations where the flow is turbulent and quasi-2D. An attractive feature of turbulent, 2D, hydrodynamic flows is that the fluid robustly self-organizes into large, coherent, long-lived vortices amidst a backdrop of small eddies without a need to invoke special initial conditions \\citep{carn91,wm93,tabeling02}. Such a property arises from the fact that the so-called ``vortex-stretching term'' in the vorticity equation is absent in 2D, thereby allowing an inverse cascade of energy (and forward cascade of enstrophy). Turbulence thus becomes a friend and not a foe. If protoplanetary discs are capable of producing turbulent, quasi-2D flows, then these may seed large-scale vortices that may survive for many orbital time scales. Such optimism should be tempered by the fact that off-midplane vortices have not been observed in simulations of protoplanetary discs with dust settling \\citep{chiang08,johansen09}. \\emph{The main question we are addressing in this study is: assuming vortices can be generated and sustained in discs, what sizes of particles do they capture, where are the capture locations and when does capture occurs?} In \\S\\ref{sect:basic}, we discuss/review the order-of-magnitude physics associated with protoplanetary discs, particle-gas interactions and vortices. In \\S\\ref{sect:static}, we start with the simpliest case of a static, minimum mass solar nebula disc and show that there are generically four vortex zones within any disc with surface density and temperature profiles that are decreasing, power-law functions of $r$. In \\S\\ref{sect:viscous}, we consider the next level of sophistication, which is the case of an evolving, viscously-heated disc; we show that there are preferred locations and particle radii for vortex capture. In \\S\\ref{sect:irradiated}, we generalize to the case of an evolving, viscous, irradiated disc. We demonstrate that discs which are able to both settle particles to their midplanes and capture them via vortices have upper limits to their masses that are consistent with most observed discs. We also show that the maximum particle radius for vortex capture in these discs is $\\sim 10$ cm, independent of disc model and weakly dependent on stellar, disc and dust properties. In \\S\\ref{sect:discussion}, we summarize our conclusions, discuss the open questions concerning the physics of vortices and describe the relevance of our results to observations. Table \\ref{tab:parameters} lists the fiducial values adopted for the parameters of our models. ", "conclusions": "\\label{sect:discussion} \\subsection{Summary} We have examined particle trapping by vortices in evolving, viscous and/or irradiated discs as a function of radial distance, initial disc mass, dust opacity and viscosity. The salient points of our study are: \\begin{enumerate} \\item If the surface density and effective temperature of protoplanetary discs are decreasing, power-law functions of the distance from the star, then all discs contain four vortex zones for a fixed particle size (Figures \\ref{fig:schematic} and \\ref{fig:mmsn}). \\item There is an annulus at intermediate distances from the star where small particles reside and may grow via coagulation. The size of this annulus decreases with time and also shrinks as the particles grow (Figure \\ref{fig:distances}). Vortex capture is optimal near and at the boundaries of this annulus, and occurs within an orbital period. \\item The optimal capture radii ($\\sim 1$ mm to $\\sim 10$ cm; Figures \\ref{fig:vis} and \\ref{fig:rad}) of particles are comparable to or smaller than the sizes of particles ($\\sim 1$ m) that drift fastest through the protoplanetary disc. The enhanced particle concentration within the vortices may make the characteristic time for particle growth shorter than that for radial drift. \\item Vortices in older discs prefer to capture smaller particles, a phenomenon we term ``vortex aging'' (Figures \\ref{fig:vis} and \\ref{fig:rad}). If coagulation between sub-micron particles mixed with the gas can only produce small ($\\lesssim 1$ mm) particles, then vortices can capture them throughout the lifetime of the disc. However, if coagulation manufactures larger ($\\sim 10$ cm) particles when the disc is young, vortices must form early to capture them. This metaphorical dance between coagulation and vortex aging determines how efficiently vortices help the disc to retain its mass in solids. \\item More viscous, more dust-opaque and/or less massive discs can have vortices that trap a broader range of particle sizes throughout the lifetime of the disc. While the coagulation of grains with sizes $\\lesssim 1$ mm needs to be more in synch with the evolution of more massive discs, such discs are also expected to grow grains to larger sizes more rapidly. \\item The maximum size of particle that can be trapped by vortices is $\\sim 10$ cm, independent of disc model and weakly dependent on stellar, disc and dust properties. Discs where particles settle to the midplane (within $\\gtrsim 10$ yr) and are sequestered in vortices have upper limits to their masses ($\\gtrsim 0.1~M_\\odot$) that are consistent with those of most observed discs. If vertical mixing is present (e.g., via turbulence), the maximum masses can be much smaller. \\end{enumerate} \\subsection{The Physics of Vortices: Open Questions} Many open questions remain concerning the microphysics of vortices. While we have shown that particles can be gathered by vortices, the outcome of these captures is uncertain. Vortices may concentrate enough mass to enhance collisions rates by an order of magnitude (or more) or to trigger local gravitational instabilities \\citep{aw95,gl99,gl00,kb06}. Both paths leads to the formation of self-gravitating objects which will not drift through the disc. However, at least some of the published simulations are run in 2D (e.g., \\citealt{davis00,inaba06,lyra09}) --- in the absence of viscosity or particles, such vortices live forever, implying that the mass of the planetesimal, embyro or planet formed depends either on the time the simulation is executed or the breakup of the vortex by the non-linear feedback of the concentrated particles. It is more likely that if the centres of vortices are relatively quiescent, they then serve as nurseries for coagulation to occur between somewhat larger particles. Vortex formation is also uncertain. \\cite{lp10} show that the subcritical baroclinic instability (SBI) is a plausible way of seeding vortices. Discs develop this \\emph{non-linear} instability when they are (radially) convectively unstable, have non-negligible thermal diffusion, and are subjected to finite vorticity perturbations ($\\sim 0.1$). The associated Reynolds number for the shearing box simulations in this study is ${\\cal R} \\sim 10^5$; because the threshold vorticity amplitude for invoking the SBI decreases with increasing ${\\cal R}$, they speculate that the threshold amplitude could be very small (and possibly sub-sonic) in realistic discs. Another possibility for generating vortices is via the (linear) Rossby wave instability \\citep{lovelace99,vt06}, which was invoked by \\cite{inaba06} to consider 2D vortices in protoplanetary discs. \\cite{inaba06} found that the formation of vortices via the Rossby wave instability critically depends on the amplitude and width of an initial density bump placed within the disc. This bump appears to be most unstable to perturbations with an azimuthal mode number of 5 (i.e., ``$m=5$'' perturbations). \\cite{meheut10} performed 3D simulations of stratified discs and concluded that strong and persistent vortices emerge out of the flow via the Rossby wave instability. \\cite{davis00} have noted that Rossby waves are only supported by flows with non-vanishing vorticity gradients, implying that they are relevant mostly in incompressible, inviscid flows. The subject of vortex survivability is mired in deeper controversy. \\cite{lp09} subject vortices embedded in a shearing sheet to 3D perturbations; only vortices with $4 \\lesssim q \\lesssim 6$ survive the elliptical instability in unstratified discs. This stability region vanishes when stratified discs are considered. By contrast, \\cite{lithwick09} asserts that weak vortices ($q \\gg 1$) can survive in quasi-2D flows. \\cite{lp10} find that vortices develop bursts of turbulence in their cores before surviving as weaker vortices; the SBI amplifies the vortices and the cycle restarts. If turbulent vortex cores are generic and ubiquitous phenomena in realistic discs, then they may pose a setback to using vortices as mechanisms for concentrating particles. Many collective properties of vortices in turbulent, 2D, hydrodynamic flows remain poorly understood. Among these is the ``universal decay theory'', which is the empirical observation that the vortex density, radius, velocity, mean separation between vortices, enstrophy and kurtosis can be approximated by power laws of time parametrized by a single parameter --- a first-principles explanation is still being sought \\citep{tabeling02}. Statistical theories predict vortices to always ultimately merge, in contrast to experiments that show that above a critical separation, a pair of vortices may remain separated for many dynamical times \\citep{tabeling02}. Despite these uncertainties, vortices provide an interesting alternative to streaming instabilities \\citep[e.g.,][]{gp00, yg05, jy07, yj07, johansen09}, which require dust-to-gas ratios to approach of order unity (presumably near the disc midplane). Vortices require no special dust-to-gas ratio to trap particles and may exist off the disc midplane. In both cases, particles with $\\xi/\\Omega \\sim 1$ are captured and the capture size decreases as the gas in the disc dissipates. Therefore, vortex trapping and streaming instabilities may provide complementary mechanisms for particle concentration and/or growth. A key question to study and explore is the size distribution of planetesimals produced by each mechanism, since this might have an impact on the types of (exo)planets produced in a system. Preliminary analyses of the initial size distribution of Solar System planetesimals suggest a broad range of values, ranging from $\\sim 1$--10 km \\citep{kb10} to $\\sim 100$--1000 km \\citep{morb09}. \\subsection{Observational Relevance} Our study makes falsifiable predictions about the size of particles concentrated as a function of distance from the star. Particles smaller or larger than the optimal radii for vortex capture will be uniformly distributed throughout the disc, while those with sizes $\\sim 1$ mm to $\\sim 10$ cm will be concentrated near their respective transitional distances $r_{\\rm in}$ and $r_{\\rm out}$ (Figure \\ref{fig:schematic}). If particle concentration leads to the growth of larger particles, these particles will then decouple from the gas and migrate out of the vortices. The decoupling will be stronger in the inner regions of the disc as $\\xi/\\Omega \\propto a^{-2}$ (instead of $\\propto a^{-1}$ in the outer regions). In our Solar System, the smallest constituents (chondrules) of $\\sim 100$ km-sized asteroids have sizes $\\sim 1$ mm \\citep{hewins97}, comparable to the sizes of particles trapped by vortices. \\cite{cuzzi08} (and references therein) have pointed out that there is a spread of about 1 Myr between the formation times of the oldest and youngest objects in the same meteorite, implying that particle growth was fairly inefficient. This inference is in turn consistent with the limited temporal windows for particle trapping implied by vortex aging. Another interesting property of $\\sim 10$--100 km-sized asteroids is that many of them are formed from a physically and chemically homogeneous mix of particles of a similar size, consistent with the aerodynamic sorting property of vortices. In extrasolar settings, observations characterizing grain opacity (and hence grain growth) in protoplanetary discs can quantify particle populations as functions of radial distance, but these are still nascent and have only been accomplished for a small number of objects, e.g., HD 163296 \\citep{natta07}. Therefore, the hypothesis that vortices serve as nurseries for particle growth will need to be tested by future infrared, submillimetre and centimetre observations of protoplanetary discs, and may lead to constraints on the time scales for grain coagulation (Figures \\ref{fig:vis} and \\ref{fig:rad}). For example, the {\\it Atacama Large Millimeter Array (ALMA),} which operates at wavelengths between 0.3 and 3.6 mm, might be able to detect disc features associated with vortex capture. Coagulation-fragmentation simulations can subsequently be performed to convert the observed fluxes and $\\alpha$ values into constraints on the grain properties (e.g., mass/size distribution, porosity; \\citealt{birn10}). Even if vortices do not survive long enough to create self-gravitating structures, they will certainly play a strong role in redistributing matter throughout the disc. Whether the redistribution of matter by the vortices plays any significant role in the eventual formation of planetesimals is unknown. An implication of a size-dependent redistribution of matter is that if the amount of electric charge carried by a particle is proportional to its size, then charge separation in protoplanetary discs may be a fairly common phenomenon. Our study has shown that if vortices form in protoplanetary discs, they are important in discs with typical masses and for particles that are likely to condense out of the protostellar nebula. The capture of particles also occurs at distances relevant to planet formation. With this study, we hope to (re)ignite the debate in connecting the microphysics of vortices with the global properties of protoplanetary discs, as the first step towards understanding the \\emph{efficiency} of planetesimal --- and eventually planet --- formation." }, "1005/1005.0385_arXiv.txt": { "abstract": "Bulges are commonly believed to form in the dynamical violence of galaxy collisions and mergers. Here we model the stellar kinematics of the Bulge Radial Velocity Assay ({\\sl BRAVA}), and find no sign that the Milky Way contains a classical bulge formed by scrambling pre-existing disks of stars in major mergers. Rather, the bulge appears to be a bar, seen somewhat end-on, as hinted from its asymmetric boxy shape. We construct a simple but realistic $N$-body model of the Galaxy that self-consistently develops a bar. The bar immediately buckles and thickens in the vertical direction. As seen from the Sun, the result resembles the boxy bulge of our Galaxy. The model fits the {\\sl BRAVA} stellar kinematic data covering the whole bulge strikingly well with no need for a merger-made classical bulge. The bar in our best fit model has a half-length of $\\sim 4\\;\\kpc$ and extends $20^\\circ$ from the Sun-Galactic Center line. We use the new kinematic constraints to show that any classical bulge contribution cannot be larger than $\\sim$ 8\\% of the disk mass. Thus the Galactic bulge is a part of the disk and not a separate component made in a prior merger. Giant, pure-disk galaxies like our own present a major challenge to the standard picture in which galaxy formation is dominated by hierarchical clustering and galaxy mergers. ", "introduction": "\\label{sec:intro} Astronomers commonly make the Copernican assumption that our Milky Way is in no way unusual; then they can exploit the fact that we live in it to study galaxy formation in special detail. Past assumptions about our Galactic bulge grew out of our developing understanding of galaxy formation. It is well known that spiral galaxies consist of three main components, an invisible dark matter halo, an embedded, flat disk, and a central bulge. The bulge of our Galaxy is $>$99\\% made of stars that are at least 5 Gyr old \\citep{cla_etal_08} with a wide range of metal abundances \\citep{mcw_ric_94,ful_etal_06,zoc_etal_08}. In this respect and many others, big bulges are similar to (diskless) elliptical galaxies. The formation of ellipticals is well understood. Hierarchical gravitational clustering of initial fluctuations in the cosmological density results in galaxy collisions and mergers that scramble flat disks into rounder ellipticals \\citep{toomre_77_merger,whi_ree_79,ste_nav_02,nak_nom_03}. Significant energy has been invested in developing this very successful theory of galaxy formation, and it was natural to think that our Galactic bulge is a product of it. There is little danger that the picture is fundamentally wrong \\citep{binney_04_sydney}. But it is incomplete. The theme of this paper is that our Galactic bulge is indeed normal but that it is prototypical of different formation processes than are usually assumed. A complementary suite of evolution processes shapes isolated galaxies. They evolve by rearranging energy and angular momentum; this grows central components that masquerade as classical bulges but that formed directly out of disks without any collisions \\citep{kormen_93,kor_ken_04}. To distinguish them from merger remnants, we call them ``pseudobulges''. They come in two varieties. Some are flattened; they are grown out of disk gas transported inward by nonaxisymmetries such as bars. Another variety is recognized only in edge-on galaxies. When bars form out of disks, they buckle vertically and heat themselves into thickened structures that look box-shaped when seen edge-on \\citep{com_san_81,com_etal_90,rah_etal_91}. Our Galaxy contains such a box-shaped \\citep{mai_etal_78,wei_etal_94,dwe_etal_95} pseudobulge \\citep{kor_ken_04}. The identification of this boxy structure as an edge-on bar is particularly compelling because infrared imagery shows a parallelogram-shaped distortion \\citep{mai_etal_78,wei_etal_94,dwe_etal_95} that is naturally explained as a perspective effect: the near end of the bar is closer to us than the far end. So its vertical extent is taller on the near side than on the far side \\citep{bli_spe_91}. \\citet{zhao_96} developed the first rapidly rotating bar model that fitted this distortion. Zhao's model was based on the \\citet{schwar_79} orbit superposition technique, so it was self-consistent and in steady state, but it did not evolve into that state from plausible initial conditions. Also, little stellar kinematic data were available to constrain Zhao's steady-state model and early $N$-body models \\citep{fux_97,fux_99,sev_etal_99}, and subsequent radial velocity data from a survey of planetary nebulae, although compared with a range of dynamical models \\citep{bea_etal_00}, led to only limited conclusions because of the small numbers and uncertain population membership of the planetary nebulae. In this paper, we simulate numerically the self-consistent formation of a bar that buckles naturally into a thickened state, and we scale that model to fit new kinematic data on bulge rotation and random velocities. The radial velocity observations are provided by the Bulge Radial Velocity Assay ({\\sl BRAVA}; \\citealt{ric_etal_07,how_etal_08}). This is a spectroscopic survey of the stellar radial velocities of M-type giant stars whose population membership in the bulge is well established. These giants provide most of the 2 $\\mu$m radiation whose box-shaped light distribution motivates bar models. {\\sl BRAVA} emphasizes measurements in two strips at latitude $b=-4^\\circ$ and $b=-8^\\circ$ and at longitude $-10^\\circ$ $< l <$ $+10^\\circ$. A strip along the minor axis ($l \\equiv 0^\\circ$) has also been observed. We use nearly 5,000 stellar radial velocities in this report. A preliminary analysis of data found strong cylindrical rotation \\citep{how_etal_09} consistent with an edge-on, bar-like pseudobulge, although a precise fit of a bar model to the data was not available. This success leads us here to construct a full evolutionary $N$-body model that we can fit to the radial velocity data. \\begin{figure} \\centerline{ \\includegraphics[angle=0.,width=\\hsize]{cont.ps} } \\caption{Upper three panels: Face-on and side-on views of the surface density of our best-fitting model as seen from far away. The Sun's position 8.5~kpc from the Galactic center is marked along the $+x$ axis. The Galaxy rotates clockwise as seen in the face-on projection. Bottom panel: Model surface brightness map in Galactic coordinates as seen from the Sun's location. Our perspective makes the box-shaped, edge-on bar look taller on its nearer side. The Galactic boxy bulge is observed to be similarly distorted.} \\label{fig:cont} \\end{figure} ", "conclusions": "In our models, a bar develops self-consistently from the initially unbarred, thin disk. Bar formation enhances the radial streaming motions of disk particles, so the radial velocity dispersion quickly grows much bigger than the vertical one. Consequently the disk buckles vertically out of the plane like a fire hose; this is the well known buckling or corrugation instability \\citep{toomre_66,com_etal_90,rah_etal_91}. It raises the vertical velocity dispersion and increases the bar's thickness. This happens on a short dynamical timescale and saturates in a few hundred million years. The central part of the buckled bar is elevated well above the disk mid-plane and resembles the peanut morphology of many bulges including the one in our Galaxy \\citep{com_etal_90,rah_etal_91}. Out of a large set of $N$-body models, we find the one that best matches our {\\sl BRAVA} kinematic data after suitable mass scaling. The barred disk evolved from a thin exponential disk that contains $M_{\\rm d}=4.25\\times 10^{10}\\Ms$, about 55 \\% of the total mass at the truncation radius (5 scale-lengths). The scale-length and scale-height of the initial disk are $\\sim$ 1.9~kpc and 0.2~kpc, respectively. The disk is rotationally supported and has a Toomre-Q of 1.2. The amplitude of the final bar is intermediate between the weakest and strongest bars observed in galaxies. The bar's minor-to-major axial ratio is about 0.5 to 0.6, and its half-length is $\\sim$ 4 kpc. Figure~1 (top three panels) shows face-on and side-on views of the projected density of the best-fitting model. A distinctly peanut shaped bulge is apparent in the edge-on projection. Figure~1 (bottom panel) shows the surface brightness distribution in Galactic coordinates as seen from the Sun's vantage point. Nearby disk stars dilute the peanut shape, but the bar still looks boxy. Moreover, from close up, an asymmetry in the longitudinal direction is apparent; this means that the bar cannot be aligned with the direction from the Sun to the Galactic center. Rather, its near end is at positive Galactic longitude, so it looks taller in that quadrant, and it extends farther from the Galactic center on the near side than on the far side. Both the boxy shape and the asymmetry are in good agreement with the morphology revealed by the COBE satellite near-infrared images \\citep{wei_etal_94,dwe_etal_95}. Figure~2 compares the best-fitting model kinematics (solid lines) with the mean velocity and velocity dispersion data from the {\\sl BRAVA} and other surveys \\citep{ran_etal_09}. All velocities presented here have been converted to Galactocentric values (the line-of-sight velocity that would be observed by a stationary observer at the Sun's position). For the first time, our model is able simultaneously to match the mean velocities and velocity dispersions along two Galactic latitudes ($-4^\\circ$ and $-8^\\circ$) and along the minor axis. Figure~3 constrains the angle between the bar and the line that connects the Sun to the Galactic center. It compares the model results with the data in the $b=-4^\\circ$ major-axis strip as we vary the above angle. Clearly the smallest bar angles give the best match to the velocity dispersions. Intriguingly, we find that the velocity dispersions provide much stronger constraints than the mean velocity profile. A bar angle of $0^\\circ$ also matches the kinematics well. However, the photometric asymmetry excludes a bar that is pointed at the Sun. We therefore conclude that the overall best-fitting model has a bar angle of $\\sim 20^\\circ$. Other studies converged on a similar bar angle \\citep{sta_etal_97,freude_98,fux_97,fux_99,bis_ger_02}. The excellent match to the data in Figures 1 -- 3 strongly supports the suggestion that the boxy pseudobulge of the Milky Way is an edge-on, buckled bar that evolved from a cold, massive disk. The thickened disk in the pseudobulge-forming process may have contributed to the thick disk of the Milk Way, as hinted from chemical similarities of Galactic bulge and local thick disk stars \\citep{alv_etal_10,ben_etal_10}. The model in Figures 1 -- 3 contains no classical bulge component. Could a small classical bulge also be present? Could it have been spun up by the formation of a bar, flattened thereby and made hard to detect? To constrain such multi-component models with {\\sl BRAVA} kinematics, we also constructed models with a pre-existing classical bulge. The distribution function for the live classical bulge component was generated iteratively \\citep{deb_sel_00} to ensure that both the disk and the classical bulge were initially in equilibrium, and we required that the bulge parameters are close to the fundamental plane for classical bulges and ellipticals \\citep{kor_etal_09}. The setup of the disk is the same as in the disk-only model. We show in Figure~4 that inclusion of a classical bulge -- one widely thought to be typical of Sbc spiral galaxies like our own -- greatly worsens the model fit to the data. The degradation is especially obvious along the Galaxy's minor axis. Including a classical bulge with just 8\\% of the disk mass considerably worsens the fit of the model to the data. Our models rule out that the Milky Way has a significant classical bulge whose mass is $>\\sim$ 15 \\% of the disk mass. \\begin{figure*}[!ht] \\centerline{ \\includegraphics[angle=0.,width=0.8\\hsize]{solar_v_3panel_vary_classical_bulge2.ps}} \\caption{Fits to the kinematic data (cf.~Figure~2) of models that include a pre-existing classical bulge. The heavy black lines from Figure 2 represent the model without a classical bulge. The red, green, and blue lines are for models whose classical bulges have masses of 8\\%, 15\\%, and 30\\%, respectively, of the disk mass $M_{\\rm disk}$. Including a classical bulge significantly worsens the model fits to the data, especially along the minor axis. } \\label{fig:varybulge} \\end{figure*} Could a smaller, merger-built bulge hide inside the boxy bar? The only result that we are aware of that might point to such a conclusion is the observed drop in stellar metal abundances with increasing height above the Galactic plane \\citep{zoc_etal_08,zoccal_10}. \\citet{zoc_etal_08} argue that this means that the bulge must consist of both a classical and an edge-on bar component. However, no kinematic gradient or transition corresponding to the abundance gradient is observed. Moreover, an abundance gradient can be produced within the context of secular pseudobulge formation if some of the vertical thickening is produced by resonant heating of stars that scatter off the bar \\citep{pfe_nor_90}. If the most metal-poor stars are also the oldest stars, then they have been scattered for the longest time and now reach the greatest heights. Our results have important implications for galaxy formation. We demonstrate that the boxy pseudobulge is not a separate component of the Galaxy but rather is an edge-on bar. Bars are parts of disks. To be sure, the stars in our Galactic bar are older than most disk stars. But those stars could have formed over a short period of time but long before the bar structure formed \\citep{wyse_99,freema_08_IAU}, their old age \\citep{zoc_etal_03,ful_etal_07} is therefore not an argument against the internal secular evolution model. Our kinematic observations show no sign that the Galaxy contains a significant merger-made, ``classical'' bulge. So, from a galaxy formation point of view, we live in a pure-disk galaxy. Our Galaxy is not unusual: it is very similar to another giant edge-on galaxy with a boxy bulge, NGC~4565. \\citet{kor_bar_10} recently show that NGC~4565 does not contain even a small classical bulge component and that it therefore is another giant, pure-disk galaxy that contains no sign of a merger remnant. In fact, giant, pure-disk galaxies are common in environments like our own that are far from rich clusters of galaxies \\citep{kor_etal_10}. Classical-bulge-less, pure-disk galaxies present an acute challenge to the current picture of galaxy formation in a Universe dominated by cold dark matter -- growing a giant galaxy via hierarchical clustering ($\\Vc \\simeq 220$ km~s$^{-1}$ in the Milky Way) involves so many mergers that it seems almost impossible to avoid forming a substantial classical bulge \\citep{pee_nus_10,age_etal_10}. How did our Galaxy grow so large with no observational sign that it suffered a major merger after the time 9 -- 10 Gyr ago \\citep[e.g.][]{win_kep_08} when the first disk stars formed?" }, "1005/1005.2441_arXiv.txt": { "abstract": "We determine the phase diagram for dense carbon/ oxygen mixtures in White Dwarf (WD) star interiors using molecular dynamics simulations involving liquid and solid phases. Our phase diagram agrees well with predictions from Ogata et al. and Medin and Cumming and gives lower melting temperatures than Segretain et al. Observations of WD crystallization in the globular cluster NGC 6397 by Winget et al. suggest that the melting temperature of WD cores is close to that for pure carbon. If this is true, our phase diagram implies that the central oxygen abundance in these stars is less than about 60\\%. This constraint, along with assumptions about convection in stellar evolution models, limits the effective $S$ factor for the $^{12}$C($\\alpha,\\gamma$)$^{16}$O reaction to $S_{300}\\leq170$ keV barns. ", "introduction": " ", "conclusions": "" }, "1005/1005.0996_arXiv.txt": { "abstract": "The Extragalactic Background Light (EBL) includes photons with wavelengths from ultraviolet to infrared, which are effective at attenuating gamma rays with energy above $\\sim 10$ GeV during propagation from sources at cosmological distances. This results in a redshift- and energy-dependent attenuation of the $\\g$-ray flux of extragalactic sources such as blazars and Gamma-Ray Bursts (GRBs). The Large Area Telescope onboard {\\em Fermi} detects a sample of $\\g$-ray blazars with redshift up to $z\\sim 3$, and GRBs with redshift up to $z\\sim 4.3$. Using photons above 10 GeV collected by {\\em Fermi} over more than one year of observations for these sources, we investigate the effect of $\\g$-ray flux attenuation by the EBL. We place upper limits on the $\\g$-ray opacity of the Universe at various energies and redshifts, and compare this with predictions from well-known EBL models. We find that an EBL intensity in the optical--ultraviolet wavelengths as great as predicted by the ``baseline'' model of \\citet{Stecker06} can be ruled out with high confidence. ", "introduction": "The {\\em Fermi} Gamma Ray Space Telescope was launched 2008 June 11, to provide an unprecedented view of the $\\g$-ray Universe. The main instrument onboard {\\em Fermi}, the Large Area Telescope (LAT), offers a broader bandpass \\citep[$\\sim$ 20~MeV to over 300~GeV;][]{FERMI} and improved sensitivity (by greater than an order of magnitude) than that of its predecessor instrument EGRET onboard the {\\em Compton Gamma Ray Observatory}~\\citep{EGRET}, and the Italian Space Agency satellite {\\em AGILE}~\\citep{AGILE}, which was launched in 2007. The LAT observes the full sky every 3 hr in survey mode leading to a broadly uniform exposure with less than $\\sim 15\\%$ variation. The Gamma-ray Burst Monitor, the lower energy ($\\sim 8$~keV -- 40~MeV) instrument onboard {\\em Fermi}, observes the full un-occulted sky at all times and provides alerts for transient sources such as GRBs. A major science goal of {\\em Fermi} is to probe the opacity of the Universe to high-energy (HE) $\\g$-rays as they propagate from their sources to Earth. Such energetic photons are subject to absorption by production of electron-positron ($e^-e^+$) pairs while interacting with low energy cosmic background photons~\\citep{Nishikov61,Gould66,Fazio70} if above the interaction threshold: $\\epsilon_{\\rm thr}=(2 m_e c^2)^2/(2E(1-\\mu))$ where $\\epsilon$ and $E$ denote the energies of the background photon and $\\g$ ray, respectively, in the comoving frame of the interaction, $m_ec^2$ is the rest mass electron energy, and $\\theta = \\arccos(\\mu)$ the interaction angle. Because of the sharply peaked cross section close to threshold, most interactions are centered around $\\epsilon^*\\approx 0.8(E/{\\rm TeV})^{-1}$eV for a smooth broadband spectrum. Thus, the extragalactic background light (EBL) at UV through optical wavelengths constitutes the main source of opacity for $\\g$-rays from extragalactic sources (Active Galactic Nuclei: AGN and GRBs) in the LAT energy range. The effect of absorption of HE $\\g$-rays is then reflected in an energy- and redshift dependent softening of the observed spectrum from a distant $\\g$-ray source. The observation, or absence, of such spectral features at HEs, for a source at redshift $z$ can be used to constrain the $\\g\\g\\to e^+e^-$ pair production optical depth, $\\tau_{\\g\\g}(E,z)$. The EBL is dominated by radiation from stars, directly from their surface and via reprocessing by dust in their host galaxies, that accumulated over cosmological evolution. Knowledge of its intensity variation with time would probe models of galaxy and star formation. The intensity of the EBL from the near-IR to ultraviolet is thought to be dominated by direct starlight emission out to large redshifts, and to a lesser extent by optically bright AGN. At longer wavelengths the infrared background is produced by thermal radiation from dust which is heated by starlight, and also emission from polycyclic aromatic hydrocarbons~\\citep[see e.g.][]{driver08}. Direct measurements of the EBL is difficult due to contamination by foreground zodiacal and Galactic light~\\citep[e.g.,][]{Hauser01}, and galaxy counts result in a lower limit since the number of unresolved sources is unknown~\\citep[e.g.,][]{Madau00}. Furthermore, evolution of the EBL density in the past epochs ($z>0$) that is required to calculate the $\\g$-ray flux attenuation from distant sources cannot be addressed by measuring the EBL density at the present epoch (z=0). Hence, several approaches have been developed to calculate the EBL density as a function of redshift. The models encompass different degrees of complexity, observational constraints and data inputs. Unfortunately, the available direct EBL measurements do not constrain these models strongly at optical-UV wavelengths due to the large scatter in the data points. A description of the different models is beyond the scope of this work; we refer the reader to the original works on the various EBL models \\citep[e.g.,][]{Salamon98, Stecker06, Kneiske02, Kneiske04, Primack05, Gilmore09, Franceschini08, Razzaque09, Finke09_model}. We note that all recent EBL models, and in particular all models used in this paper, use almost identical parameters of a $\\Lambda$CDM cosmology model. For the analyses presented in this work we have made use of the optical depth values $\\tau(E,z)$ provided by the authors of these EBL models. These models are available via webpages \\footnote{ \\url{http://www.physics.adelaide.edu.au/~tkneiske/Tau_data.html} for Kneiske 2004; \\url{http://www.phy.ohiou.edu/~finke/EBL/index.html} for Finke et al. 2010}, analytical approximations (as in, e.g., \\citet{Stecker06}), published tables (as in, e.g., \\citet{Franceschini08}) or via private communications (which is the case for, e.g., \\citet{Salamon98, Primack05, Gilmore09, Finke09_model} for this work). Since the optical depth values are usually available in tabular form, for exact values of observed energy $E$ and redshift $z$ a linear interpolation of $\\tau(E,z)$ is used for arbitrary values of $E$ and $z$ in our calculations below. The range of predictions by these EBL models is illustrated in Figure~\\ref{fig:tau_vs_energy} as a function of observed $\\g$-ray energy for sources at different redshifts. The Universe is optically thin ($\\tau_{\\g\\g} < 1$) to $\\gamma$-rays with energy below $\\simeq 10$ GeV up to redshift $z\\simeq 3$, independently of the model (see also \\citet{dieter}). This is due to the rapid extinction of EBL photons shortwards of the Lyman limit. Gamma rays below $\\sim 10$~GeV are not attenuated substantially because of faint far-UV and X-ray diffuse backgrounds. \\begin{figure} \\epsscale{0.9} \\plotone{f1} \\caption{Attenuation as a function of observed gamma-ray energy for the EBL models of \\citep {Franceschini08} and \\citep{Stecker06}. These models predict the minimum and maximum absorption of all models in the literature, and thus illustrate the range of optical depths predicted in the {\\em Fermi}-LAT energy range.} \\label{fig:tau_vs_energy} \\end{figure} The primary sources of HE extragalactic $\\g$-rays are blazars and gamma ray bursts (GRBs). Blazars are active galactic nuclei (AGN) with relativistic plasma outflows (jets) directed along our line of sight. GRBs are associated with the core collapse of massive stars, or might be caused by binary mergers of neutron stars or neutron star - black hole systems. Some GRBs produce beamed high-energy radiation similar to the case of blazars but lasting for a short period of time. GRBs have not been used to constrain EBL absorption during the pre-{\\em Fermi} era mainly because of a lack of sensitivity to transient objects above 10 GeV. The sensitivity of EGRET decreased significantly above 10 GeV, and the field-of-view (FoV) of TeV instruments is small (typically $2-4^\\circ$) to catch the prompt phase where most of the HE emission occurs. The new energy window ($10- 300$ GeV) accessible by {\\em Fermi}, and the wide FoV of the LAT, makes GRBs interesting targets to constrain EBL absorption in this energy band. Evaluating the ratio of the putatively absorbed to unabsorbed fluxes from a large number of distant blazars and GRBs observed by {\\em Fermi} could result in interesting EBL constraints, as proposed by \\citet{chen04}, although intrinsic spectral curvature \\citep[e.g.,][]{Massaro06} or redshift dependent source internal absorption \\citep{Reimer07} could make this, or similar techniques, less effective. \\citet{georgan08} have proposed that Compton scattering of the EBL by the radio lobes of nearby radio galaxies such as Fornax A could be detectable by the {\\em Fermi}-LAT. If identified as unambiguously originating from such process, a LAT detection of Fornax A could constrain the local EBL intensity. Because the e-folding cutoff energy, $E(\\tau_{\\g\\g}=1)$, from $\\g\\g$ pair production in $\\g$-ray source spectra decreases with redshift, modern Cherenkov $\\g$-ray telescopes are limited to probing EBL absorption at low redshift due to their detection energy thresholds typically at or below 50~GeV to 100~GeV \\citep{Hinton09}. Ground-based $\\g$-ray telescopes have detected 35 extragalactic sources to date\\footnote{e.g.,\\url{http://www.mpi-hd.mpg.de/hfm/HESS/pages/home/sources/}, \\url{http://www.mppmu.mpg.de/~rwagner/sources/}}, mostly of the high-synchrotron peaked (HSP) BL Lacertae objects type. The most distant sources seen from the ground with a confirmed redshift are the flat spectrum radio quasar (FSRQ) 3C~279 at $z=0.536$ \\citep{Albert08} and PKS~1510-089 at $z=0.36$ \\citep{PKS1510}. Observations of the closest sources at multi-TeV energies have been effective in placing limits on the local EBL at mid-IR wavelengths, while spectra of more distant sources generally do not extend above 1 TeV, and therefore probe the optical and near-IR starlight peak of the intervening EBL \\citep[e.g.,][]{Stecker93, Stanev98, Schroedter05,Aharonian99,Aharonian02,Costamante04, Aharonian06a, Mazin07, Albert08, Krennrich08, Finke09}. The starting point for constraining the EBL intensity from observations of TeV $\\g$-rays from distant blazars with atmospheric Cherenkov telescopes is the assumption of a reasonable intrinsic blazar spectrum, which, in the case of a power law, $dN/dE \\propto E^{-\\Gamma_{int}}$ for example, that is not harder than a pre-specified minimum value, e.g., $\\Gamma_{int}\\geq\\Gamma_{min}=0.67$ or 1.5. Upper limits on the EBL intensity are obtained when the reconstructed intrinsic spectral index from the observed spectrum, $\\Gamma_{obs}$, presumably softened by EBL absorption of very high energy (VHE) $\\g$-rays, is required to not fall below $\\Gamma_{int}$. The minimum value of $\\G$ has been a matter of much debate, being reasoned to be $\\G_{int}=1.5$ by \\citet{Aharonian06a} from simple shock acceleration theory and from the observed spectral energy distribution (SED) properties of blazars, while \\citet{stecker07} argued for harder values (less than 1.5) under specific conditions based on more detailed shock acceleration simulations. \\citet{Katarzynski06} suggested that a spectral index as hard as $\\Gamma_{int}=0.67$ was possible in a single-zone leptonic model if the underlying electron spectrum responsible for inverse-Compton emission had a sharp lower-energy cutoff. \\citet{boett08} noted that Compton scattering of the cosmic microwave background radiation by extended jets could lead to harder observed VHE $\\g$-ray spectra, and \\citet{aharonian08} have argued that internal absorption could, in some cases, lead to harder spectra in the TeV range as well. A less model dependent approach uses the (unabsorbed) photon index as measured in the sub-GeV range as the intrinsic spectral slope at GeV-TeV energies. This method has recently been applied to PG~1553+113 \\citep{lat1553} and 1ES~1424+240 \\citep{lat1424,prandini10} to derive upper limits on their uncertain redshifts, and to search for EBL-induced spectral softening in {\\em Fermi} observations of a sample of TeV-selected AGN \\citep{TeVselected}. Attenuation in the spectra of higher redshift objects ($z \\gtrsim 1$) may be detectable at the lower energies that are accessible to the {\\em Fermi}-LAT, i.e., at $E\\approx 10-300$ GeV. Gamma rays at these energies are attenuated mainly by the evolving UV background, which is produced primarily by young stellar populations and closely traces the global star-formation rate. Observations with {\\em Fermi} of sources out to high redshift could therefore reveal information about the star-formation history of the Universe, as well as the uncertain attenuation of UV starlight by dust. In this paper we present constraints on the EBL intensity of the Universe derived from {\\em Fermi}-LAT observations of blazars and GRBs. The highest-energy $\\g$-rays from high redshift sources are the most effective probe of the EBL intensity, and consequently a powerful tool for investigating possible signatures of EBL absorption. In contrast to ground-based $\\g$-ray detectors, {\\em Fermi} offers the possibility of probing the EBL at high redshifts by the detection of AGN at $\\gtrsim 10$ GeV energies out to $z>3$, and additionally by the detection of GRB 080916C at a redshift of $\\sim 4.35$ \\citep{abdo09,Greiner09}. GRBs are known to exist at even higher redshifts \\citep[GRB 090423 is the current record holder with $z\\sim$ 8.2]{Tanvir09}. Therefore observations of these sources with {\\em Fermi} are promising candidates for probing the optical-UV EBL at high redshifts that are not currently accessible to ground-based (Cherenkov) telescopes. In Section~\\ref{sec:analysis} we describe our data selections, the {\\em Fermi} LAT AGN and GRB observations during the first year of operation and analysis, and we discuss potential biases in the selection. Our methodology and results are presented in Section~\\ref{sec:methods}. We discuss implications of our results in Section~\\ref{sec:discussion}, and conclude in Section~\\ref{sec:conclusion}. In the following, energies are in the observer frame except where noted otherwise. ", "conclusions": "\\label{sec:conclusion} Using the high-energy 11-month photon data set collected by {\\em Fermi} from distant blazars, and two GRBs we have (i) placed upper limits on the opacity of the Universe to $\\g$ rays in the $\\sim$10--100~GeV range coming from various redshifts up to $z\\approx 4.3$; and (ii) ruled out an EBL intensity in the redshift range $\\sim$ 1 to 4.3 as great as that predicted by \\citet{Stecker06} in the ultraviolet range at more than $4\\sigma$ post trials in two independent sources (blazars). The overall rejection significance is found to be $>10 \\sigma$ post trials therefore making this result very robust. Our most constraining sources are blazars J1504+1029, J0808-0751 and J1016+0513 with $(z, \\langle E_{max} \\rangle)$ combinations of (1.84, 48.9~GeV), (1.84, 46.8 GeV) and (1.71, 43.3 GeV), respectively. Although a likelihood ratio analysis of the latter source indicates that the sensitivity of our analysis method is approaching the EBL flux level of the ``high UV model'' of Kneiske et al (2004), multi-trial effects markedly reduced the rejection significance. The two most constraining GRBs are GRB 090902B and GRB 080916C, both of which rule out the ``baseline'' EBL model of \\citet{Stecker06} in the UV energy range at more than 3$\\sigma$ level. The ``fast evolution'' model of \\citet{Stecker06} predicts higher opacities in the LAT energy range at all redshifts, and therefore is also ruled out. Together with the results from VHE observations (e.g., \\citet{aharonian07,Mazin07}) the models by \\citet{Stecker06} seem now disfavored in the UV and mid-IR energy range. We have also calculated model-independent optical depth upper-limits $\\tau_{\\g\\g, \\rm UL} (z, \\langle E_{max} \\rangle)$ at 95\\% CL in the redshift $z\\simeq 1-2.1$ and $E_{\\rm max}\\approx 28-74$GeV ranges. As the high-energy photon data set collected by {\\em Fermi} grows in the future and more blazars and GRBs are detected at constraining energies, the $(E, z)$ phase space that constrains $\\tau_{\\g\\g}$ will become more populated. This will provide us with unique opportunities to constrain the opacity of the Universe to $\\g$-rays over a large energy and redshift range, and eventually help us further understand the evolution of the intensity of the extragalactic background light over cosmic time." }, "1005/1005.1959_arXiv.txt": { "abstract": "{Using photometry of NGC 1097 from the {\\em Herschel} PACS (Photodetector Array Camera and Spectrometer) instrument, we study the resolved properties of thermal dust continuum emission from a circumnuclear starburst ring with a radius $\\sim 900$ pc. These observations are the first to resolve the structure of a circumnuclear ring at wavelengths that probe the peak (i.e. $\\lambda \\sim 100$ \\micron) of the dust spectral energy distribution. The ring dominates the far-infrared (far-IR) emission from the galaxy---the high angular resolution of PACS allows us to isolate the ring's contribution and we find it is responsible for 75, 60 and 55\\% of the total flux of NGC 1097 at 70, 100 and 160 \\micron, respectively. We compare the far-IR structure of the ring to what is seen at other wavelengths and identify a sequence of far-IR bright knots that correspond to those seen in radio and mid-IR images. The mid- and far-IR band ratios in the ring vary by less than $\\pm 20$\\% azimuthally, indicating modest variation in the radiation field heating the dust on $\\sim 600$ pc scales. We explore various explanations for the azimuthal uniformity in the far-IR colors of the ring including a lack of well-defined age gradients in the young stellar cluster population, a dominant contribution to the far-IR emission from dust heated by older ($> 10$ Myr) stars and/or a quick smoothing of local enhancements in dust temperature due to the short orbital period of the ring. Finally, we improve previous limits on the far-IR flux from the inner $\\sim 600$ pc of NGC 1097 by an order of magnitude, providing a better estimate of the total bolometric emission arising from the active galactic nucleus and its associated central starburst.} ", "introduction": "The central regions of galaxies host some of the most intense star-formation that we can observe in the local Universe in circumnuclear starburst rings. Starburst rings are believed to be the consequence of the pile-up of inflowing gas and dust, driven by a non-axisymmetric potential from a stellar bar, on orbits located near the Inner Lindblad Resonance of the bar \\citep{combes85,athanassoula92}. The high surface densities that exist in the ring lead to high star-formation rates. Indeed starburst rings are one of few regions in non-interacting galaxies where the formation of ``super star clusters'' commonly occurs \\citep{maoz96}. The stars formed in the ring can be numerous enough to drive the structural evolution of the galaxy \\citep{norman96,kormendy04} and can be the dominant power source for the galaxy's infrared (IR) emission. Star-formation in circumnuclear rings occurs under conditions not normally found in the disks of galaxies: in addition to their high gas surface densities, these regions have dynamical timescales that are comparable to the lifetimes of massive stars. Understanding star formation in circumnuclear rings has been a long-standing problem \\citep{combes96}. There are two main models: the ``popcorn'' model \\citep{elmegreen94}, where star-formation is driven by stochastic gravitational fragmentation along the ring, and the ``pearls on a string'' model, where gas flowing into the ring is compressed near the contact points (i.e. locations where the dust lanes intersect the ring) and then forms stars a short distance downstream \\citep[e.g.,][]{boker08}. The ``pearls on a string'' model predicts a gradient in the ages of young stellar clusters as one moves away from the contact points. This has been observed in a number of starburst rings \\citep[e.g., ][]{mazzuca08,boker08}. Conversely, many well-studied rings show no evidence for an age gradient \\citep{maoz01}. It is not obvious, however, that a single mode of star-formation must occur in all rings or even at all times in a given ring \\citep{vandeven09}. KINGFISH (Key Insights into Nearby Galaxies: A Far-Infrared Survey with {\\em Herschel}, PI R. Kennicutt) is an Open-Time Key Program to study the interstellar medium (ISM) of nearby galaxies with far-IR/sub-mm photometry and spectroscopy. Among the unique aspects of the KINGFISH science program is the ability to observe thermal dust emission at unprecedented spatial resolution ($\\sim$ 5.6, 6.8 and 11.3\\arcsec at 70, 100 and 160 \\micron) using PACS (Photodetector Array Camera and Spectrometer) imaging. High spatial resolution is crucial for observing processes occurring in the central regions of galaxies. These regions represent our best opportunity to study in detail the interplay between dynamics, star-formation and feedback that regulate the fueling of nuclear activity, be it a starburst or an active galactic nucleus (AGN). Below we present PACS imaging of the galaxy NGC 1097, one of the first KINGFISH targets observed during the {\\em Herschel} Science Demonstration Program (SDP) (for PACS spectroscopy of NGC 1097 see Beir\\~{a}o et al. 2010 and for {\\em SPIRE} observations see Engelbracht et al. 2010). The source NGC 1097 is a barred spiral galaxy located at a distance of 19.1 Mpc \\citep[][1\\arcsec$\\approx 92$ pc]{willick97}. In its central kpc it hosts an intensely star-forming \\citep[$\\sim 5$ \\msun\\ yr$^{-1}$;][]{hummel87} ring with a radius of $\\sim 900$ pc. The ring's rotation speed of $\\sim 300$ km s$^{-1}$ \\citep[corrected for inclination,][]{storchi-bergmann96}, corresponds to a rotation period of $\\sim 18$ Myr. The galaxy's nucleus is classified as a LINER from optical emission line diagnostics \\citep{phillips84}, but is shown to be a Seyfert 1 by its double-peaked H$\\alpha$ profile \\citep{storchi-bergmann93}. UV spectroscopy has revealed a few Myr old burst of star-formation in the central 9 pc of the galaxy \\citep{storchi-bergmann05}. With the high spatial resolution of {\\em Herschel} PACS, we can resolve the starburst ring and inner 600 pc of NGC 1097 for the first time at wavelengths near the peak of the dust spectral energy distribution (SED). ", "conclusions": "We have presented {\\em Herschel} PACS observations from KINGFISH of the inner kpc of the barred spiral galaxy NGC 1097. These are the first observations to resolve a starburst ring at wavelengths probing the peak of the dust SED. We show a comparison of the ring in a variety of tracers and find similar bright knots in the mid- and far-IR and radio continuum. These knots do not correspond to the same knots traced by CO. We find modest variation azimuthally in the mid- and far-IR band ratios suggesting that either there is no azimuthal age gradient, as would be predicted by the ``pearls on a string'' mode of star-formation, that dust heating is dominated by an older stellar population and/or that the dust heating variations get quickly erased over the short ring orbital period ($\\sim 18$ Myr). Finally, we place an order-of-magnitude tighter constraint on the far-IR emission originating in the central $\\sim 600$ pc of the galaxy." }, "1005/1005.2165.txt": { "abstract": "The apparent accelerating expansion of the Universe is forcing us to examine the foundational aspects of the standard model of cosmology -- in particular, the fact that dark energy is a direct consequence of the homogeneity assumption. We discuss the foundations of the assumption of spatial homogeneity, in the case when the Copernican Principle is adopted. We present results that show how (almost-) homogeneity follows from (almost-) isotropy of various observables. The analysis requires the fully nonlinear field equations -- i.e., it is not possible to use second- or higher-order perturbation theory, since one cannot assume a homogeneous and isotropic background. Then we consider what happens if the Copernican Principle is abandoned in our Hubble volume. The simplest models are inhomogeneous but spherically symmetric universes which do not require dark energy to fit the distance modulus. Key problems in these models are to compute the CMB anisotropies and the features of large-scale structure. We review how to construct perturbation theory on a non-homogeneous cosmological background, and discuss the complexities that arise in using this to determine the growth of large-scale structure. ", "introduction": "The standard model of the Universe -- the LCDM ``concordance\" model -- is a perturbed FLRW model containing cold dark matter and dark energy in the form of a cosmological constant $\\Lambda$. This model is highly successful in being able, up to now, to fit all cosmological observations, with the same small set of parameters~\\cite{1}. However, there is as yet no satisfactory explanation for the value of $\\Lambda$, which appears in Einstein's field equations in precisely the same form as the total vacuum energy of quantum fields. This term is responsible for the late-time acceleration of the expansion of the Universe within the spatially homogeneous FLRW framework. The unresolved nature of $\\Lambda$ and of alternative forms of dark energy throws into sharp focus the foundations of the standard model -- in particular, the spatial homogeneity assumption. We will address two questions in this paper: \\begin{itemize} \\item[{\\bf (Q1)} ] What is the basis for the spatial homogeneity assumption? \\item[{\\bf (Q2)} ] If we drop this assumption, can we find a model that is consistent with all cosmological observations? \\end{itemize} In summary, the current answer to (Q1) is that there are various results which motivate homogeneity (assuming the Copernican Principle); the most important one is that the observed (almost-)isotropy of the CMB provides strong evidence for (almost-)homogeneity of the Universe on large scales~\\cite{2}. But crucial open questions remain, which we identify and discuss. Standard cosmological perturbation theory is powerless here -- the argument depends on a fully nonlinear analysis, since one cannot assume an FLRW background. The current answer to (Q2) is that the simplest inhomogeneous cosmological models -- isotropic LTB (Lema\\^\\i tre-Tolman-Bondi) models without any dark energy, and with our galaxy at the centre of a large void~-- are able to fit current supernova and some other data~\\cite{MT1,MT2,MHE,PascualSanchez:1999zr,celerier1,Tomita:2000jj,IKN,Mof1,Mof2,celerier2,VFW,CR,EM,BMN,bolejko,enqvist,YKN,AAG,Alnes:2006pf,alnes,ABNV,celerier3,Romano:2007zz,gbh1,gbh2,Enqvist,ZMS,CFL,BW,CCF,CFZ,CBKH,FLSC,mortsell,Mof3,Yoo:2010qy,Romano:2009ej,Romano:2009mr,Romano:2009qx}. In this framework, the late-time acceleration becomes a mis-interpretation of an effect due to nonlinear inhomogeneity and curvature. In order to test these models against the full range of cosmological observations, it is necessary to develop perturbation theory on an LTB background. Although standard cosmological perturbation theory provides some qualitative guidance, perturbations on an inhomogeneous background are fundamentally different -- in particular, there is no longer a simple separation of scalar, vector and tensor modes at first order. We review recent developments on this~\\cite{CCF}, and discuss remaining open questions. ", "conclusions": "Without a theoretical understanding of the value of the cosmological constant the concordance model remains phenomenological no matter how strongly observations appear to support the model. The concordance model suffers an important coincidence problem: that dark energy starts to dominate at around the time our solar system forms. Until we demonstrate observationally that the Universe is homogeneous on large scales, we should consider inhomogeneous spacetimes even if they are philosophically uncomfortable, particularly in light of the fact that in their simplest incarnation they can explain away the dark energy problem through inhomogeneity, without apparently causing other problems. It is worth reflecting on how remarkable this is. Measurements of a nonzero $\\Lambda$ may in fact be thought of as an explicit consequence of the Copernican assumption. It is important therefore to re-examine the basic assumptions of the standard concordance model -- both to test the Copernican Principle at the heart of modern cosmology, and to understand what freedom we have to develop new cosmological models that may be able to explain cosmic acceleration using curvature and inhomogeneity rather than a cosmological constant. In order to do this, we need non-perturbative methods and also new types of nonlinear perturbations where the nonlinearity arises from the inhomogeneous background. Moreover, the whole inflationary paradigm needs to be re-examined to see if inhomogeneities this large could arise naturally. We have shown how the Copernican assumption when combined with the high isotropy of the CMB implies FLRW under fairly weak assumptions~-- mainly, that the CMB rest frame is geodesic. Provided that the dark matter is CDM that is comoving with the CMB, this geodesic property will hold. But in interacting dark energy-dark matter models, for example, the geodesic radiation frame is an additional assumption. In order to test the Copernican Principle at the foundation of the FLRW models, observations which look inside our past lightcone to estimate how distant observers see the CMB~\\cite{gbh2,goodman,CS,Jia:2008ti} are necessary. In order to observationally prove FLRW, however, we also must demonstrate that galaxies follow geodesics, and that dark energy is comoving with the CMB. If galaxies were non-geodesic, then this would leave a signature as a dipole in the Hubble law which grows linearly with distance (and so is distinguishable from a normal bulk flow, which is constant)~\\cite{CB} \\iffalse : in an inhomogeneous model the Hubble law becomes~\\cite{KS} \\be H_0={\\frac{1}{3}\\Theta}+ {A_\\a e^\\a}+ {\\sigma_{\\a\\nu}e^\\a e^\\nu},~~ e_\\a u^\\a=0\\,,~e_\\a e^\\a=1\\,, \\label{hubble from series} \\ee \\fi which is derived from the term which is linear in redshift in the area distance-redshift relation~-- see Eq.~(\\ref{hubble from series}). %(Note that the number of $e$'s tell us the spherical harmonic $\\ell$: e.g., the shear term is a quadrupole.) Possible dark energy flux would leave a similar signal in the quadratic term in the area distance-redshift relation. The generalized deceleration parameter, in a fully inhomogeneous spacetime is, following~\\cite{Cthesis}, \\ba %\\fl %\\begin{array}{rl} H_0^2\\, q_0&=&\\;16\\rho+\\;12p-\\;13\\Lambda -\\;23\\D_\\a\\udot^\\a +\\;{12}{5}\\sigma_{\\a\\nu} \\sigma^{\\a\\nu}-\\;23\\omega_\\a \\omega^\\a \\nonumber\\\\&& +e^\\a\\Bigg[-\\;23\\Theta \\udot_\\a -\\;{3}{5}\\sdel_\\a\\Theta-\\dot A_\\a+\\udot^\\nu\\sigma_{\\a\\nu}+\\;25q_\\a \\nonumber\\\\ &&%\\left.\\left. ~~~~~~~~~+\\;35\\varepsilon_\\a^{~ \\nu\\gamma}\\left(\\sdel_\\nu\\omega_\\gamma +2\\udot_\\nu\\omega_\\gamma\\right)\\Bigg]%\\right. \\nonumber\\\\ && \\left.+e^\\a e^\\nu\\left[-2\\sdel_{\\<\\a}\\udot_{\\nu\\>}+\\udot_{\\<\\a}\\udot_{\\nu\\>} +E_{\\<\\a\\nu\\>}-\\;12\\pi_{\\<\\a\\nu\\>}+ \\;97\\sigma^{\\vphantom{c}}_{\\gamma\\<\\a} \\sigma_{\\nu\\>}^{\\phantom{\\nu\\>}\\gamma} \\right.\\right. \\nonumber\\\\ &&\\left.\\left. +\\omega_{\\<\\a}\\omega_{\\nu\\>}- \\;23\\sigma^{\\vphantom{c}}_{\\gamma\\<\\a} \\omega_{\\nu\\>}^{\\phantom{\\nu\\>}\\gamma} \\right]\\right. + e^\\a e^\\nu e^\\gamma\\left[-5\\udot_{\\<\\a}\\sigma_{\\nu\\gamma\\>}-\\sdel_{\\<\\a} \\sigma_{\\nu\\gamma\\>}\\right]\\nonumber\\\\ && -3e^\\a e^\\nu e^\\gamma e^\\delta\\sigma_{\\<\\a\\nu}\\sigma_{\\gamma\\delta\\>} .\\label{q_0-observable} %\\end{array} \\ea If it were confirmed observationally that the CMB frame is geodesic, and that the CMB is almost isotropic around distant observers, then this $H_0^2q_0$ should be an isotropic observable~-- a potentially useful consistency check. Although it seems obvious that CDM and baryons are just geodesic dust, we have to be careful because of issues highlighted from the averaging problem. While a distribution of non-relativistic CDM particles obey a dust equation of state at all times, more and more of the CDM particles are locked up as galaxies and clusters form, and yet those virialized clusters then become the `dust particles', within the standard paradigm. Are these geodesic? It is not yet clear how to treat this situation properly, and it is conceivable that there is a backreaction effect which changes the effective equation of state of matter at late times~\\cite{BR} (amongst other effects). Depending on precisely what is calculated, this backreaction can be divergent in the concordance model~\\cite{CAL}, and remains a significant problem to calculate non-linearly (although see~\\cite{Baumann:2010tm} who find no such divergences using a different method of averaging). Off-lightcone observations will be able to place limits on the spatial gradients of the CMB multipoles as well as the multipoles themselves, which we have shown is also a requirement for deducing almost-FLRW geometry from almost-isotropic CMB (the almost-EGS result). The important exact ETM result states that if the first three multipoles are zero then all the rest must be too. It seems reasonable that this would extend to an almost-ETM result if the first 3 multipoles are small. In this case observations which look inside our past lightcone might only need to confirm that the first three multipoles around distant observers are small in order to observationally confirm FLRW. These issues deserve further investigation, and are important because it might be possible to construct a genuinely inhomogeneous model which solves the dark energy problem, but which also satisfies the Copernican principle, maintaining the isotropy of the CMB for all or `most' observers (as claimed in~\\cite{Wiltshire:2007zj}, for example). %All results here rely on general relativity being the correct %description of gravity. In modified gravity, it is reasonably %straightforward to generalize to those cases too by treating the %extra gravitational degrees of freedom as effective matter terms %(e.g., in scalar-tensor gravity the EGS theorem holds~\\cite{CCO}). %In this case, effective energy flux terms and anisotropic pressure %terms would need to be observed to be small from local observations %as well as the CMB. Dropping the Copernican Principle, we have the key exact result that isotropy of matter observations down our past lightcone implies an LTB geometry. However, it is an open question whether almost-isotropy of matter observations on the past lightcone implies an almost-spherically symmetric LTB model. We have discussed a complete analysis of structure growth at first order, which is the next step in developing LTB models, and shown how this is much more involved than the linear perturbations of FLRW. However, we argue that it is worth the effort. All tests of the Copernican Principle, including those which can be done by examining the background dynamics down our past lightcone~\\cite{CBL,UCE}, are difficult and sensitive. Only by developing a family of inhomogeneous spacetimes to the same level of sophistication as the standard concordance model, and directly comparing them side by side, will we really be able to understand whether $\\Lambda$ is real, or actually a consequence of our homogeneity assumption. ~\\\\{\\bf Acknowledgments:} We thank Timothy Clifton, Ruth Durrer, George Ellis, Pedro Ferreira, David Matravers, John Moffat, Syksy R\\\"as\\\"anen and Marco Regis for comments and/or discussions. CC is supported by the National Research Foundation (NRF, South Africa) and RM is supported by the UK Science \\& Technology Facilities Council. This work was partially supported by the South African National Institute for Theoretical Physics and by a Royal Society (UK)/ NRF exchange grant between the Universities of Portsmouth and Cape Town. \\newpage \\appendix" }, "1005/1005.0114_arXiv.txt": { "abstract": "We present a new code, CASTRO, that solves the multicomponent compressible hydrodynamic equations for astrophysical flows including self-gravity, nuclear reactions and radiation. CASTRO uses an Eulerian grid and incorporates adaptive mesh refinement (AMR). Our approach to AMR uses a nested hierarchy of logically-rectangular grids with simultaneous refinement in both space and time. The radiation component of CASTRO will be described in detail in the next paper, Part II, of this series. ", "introduction": "In this paper, Part I of a two-part series, we present a new code, CASTRO, that solves the multicomponent compressible hydrodynamic equations with a general equation of state for astrophysical flows. Additional physics include self-gravity, nuclear reactions, and radiation. CASTRO uses an Eulerian grid and incorporates adaptive mesh refinement (AMR). Our approach to AMR uses a nested hierarchy of logically-rectangular grids with simultaneous refinement of the grids in both space and time. Spherical (in 1D), cylindrical (in 1D or 2D), and Cartesian (in 1D, 2D or 3D) coordinate systems are supported. The radiation component of CASTRO will be described in detail in the next paper, Part II, of this series. There are a number of other adaptive mesh codes for compressible astrophysical flows, most notably, ENZO \\citep{ENZO}, FLASH \\citep{flash}, and RAGE \\citep{RAGE}. CASTRO differs from these codes in several ways. CASTRO uses an unsplit version of the piecewise parabolic method, PPM, with new limiters that avoid reducing the accuracy of the scheme at smooth extrema; the other codes are based on operator-split hydrodynamics, though the most recent release of FLASH, version 3.2, includes an unsplit MUSCL-Hancock scheme. The different methodologies also vary in their approach to adaptive mesh refinement. RAGE uses a cell-by-cell refinement strategy while the other codes use patch-based refinement. FLASH uses equal size patches whereas ENZO and CASTRO allow arbitrary sized patches. ENZO and FLASH enforce a strict parent-child relationship between patches; i.e., each refined patch is fully contained within a single parent patch; CASTRO requires only that the union of fine patches be contained within the union of coarser patches with a suitable proper nesting. Additionally, FLASH and RAGE use a single time step across all levels while CASTRO and ENZO support subcycling in time. All four codes include support for calculation of self-gravity. It is worth noting that CASTRO uses the same grid structure as the low Mach number astrophysics code, MAESTRO (see, e.g., \\cite{multilevel}). This will enable us to map the results from a low Mach number simulation, such as that of the convective period and ignition of a Type Ia supernova, to the initial conditions for a compressible simulation such as that of the explosion itself, thus taking advantage of the accuracy and efficiency of each approach as appropriate. ", "conclusions": "We have described a new Eulerian adaptive mesh code, CASTRO, for solving the multicomponent compressible hydrodynamic equations with a general equation of state for astrophysical flows. CASTRO differs from existing codes of its type in that it uses unsplit PPM for its hydrodynamic evolution, subcycling in time, and a nested hierarchy of logically-rectangular grids. Additional physics includes self-gravitation, nuclear reactions, and radiation. Radiation will be described in detail in the next paper, Part II, of this series. CASTRO is currently being used in simulations of Type Ia supernovae and core-collapse supernovae; examples of simulations done using CASTRO can be found in \\citet{Joggerstetal:2009,woosley-scidac2009}. Further details on the CASTRO algorithm can be found in the CASTRO User Guide \\citep{CASTROUserGuide}." }, "1005/1005.0863_arXiv.txt": { "abstract": "We present a model for the rotational evolution of a young, solar mass star interacting with an accretion disk. The model incorporates a description of the angular momentum transfer between the star and disk due to a magnetic connection, and includes changes in the star's mass and radius and a decreasing accretion rate. The model also includes, for the first time in a spin evolution model, the opening of the stellar magnetic field lines, as expected to arise from twisting via star-disk differential rotation. In order to isolate the effect that this has on the star-disk interaction torques, we neglect the influence of torques that may arise from open field regions connected to the star or disk. For a range of magnetic field strengths, accretion rates, and initial spin rates, we compute the stellar spin rates of pre-main-sequence stars as they evolve on the Hayashi track to an age of 3~Myr. How much the field opening affects the spin depends on the strength of the coupling of the magnetic field to the disk. For the relatively strong coupling (i.e., high magnetic Reynolds number) expected in real systems, all models predict spin periods of less than $\\sim3$ days, in the age range of 1--3~Myr. Furthermore, these systems typically do not reach an equilibrium spin rate within 3~Myr, so that the spin at any given time depends upon the choice of initial spin rate. This corroborates earlier suggestions that, in order to explain the full range of observed rotation periods of approximately $1$--$10$ days, additional processes, such as the angular momentum loss from powerful stellar winds, are necessary. ", "introduction": "\\label{sec_intro} The last three decades of observational surveys of pre-main-sequence stars have resulted in the measurement of rotation periods and/or rotational velocities ($v \\sin i$) of thousands of stars \\citep[see reviews by][]{herbstea07, scholz09}. The analysis of these data have revealed a number of mysteries regarding the spin rates of young stars. Perhaps the most prominent of these mysteries is that which was first noted by \\citet{vogelkuhi81}. Specifically, a large fraction of nearly solar mass pre-main-sequence stars rotate much more slowly than expected. The fact that stars are built up from material with high specific angular momentum, and that these young stars are still contracting, leads to the expectation of a rotation rate near the breakup velocity. Solar mass stars with ages less than a few Myr have rotation periods typically in the range of 1--10 days but with a tail in the distribution extending to around 20 days \\citep[e.g.,][]{attridgeherbst92, choiherbst96, stassunea99, rebull01, rebull3ea04, coveyea05, scholz09}. The statistics show that about half of these stars are indeed fairly rapid rotators and do seem to spin up as they approach the main sequence \\citep{bouvier3ea97, delarezapinzon04, scholzea07, irwinea08}. However, roughly half of the stars younger than a few Myr rotate at about 10\\% or less of breakup speed \\citep[][]{rebull3ea04, herbstea07, scholz09}. Thus, there is some mechanism operating that is capable of removing significant amounts of angular momentum during the pre-main-sequence phase. Following the suggestion of \\citet{edwardsea93}, the modeling work of \\citet{bouvier3ea97} and \\citet{rebull3ea04} showed that the observed range of spins could be reproduced reasonably well, by assuming that the presence of an accretion disk somehow results in a constant stellar spin period of around one week. There is some observational evidence that the population of stars with disks on average rotate more slowly than those without disks \\citep[e.g.,][]{edwardsea93, choiherbst96, stassunea99, herbstea00, littlefairea05, rebullea06, ciezabaliber07}. Also, the idea that the angular spin rate is held at a constant by the presence of the disk is loosly based on physical models for the interaction between a star and surrounding disk. There are generally two prominent theoretical ideas for how accreting stars can maintain a slow spin rate. One idea is that the torques arising from the magentic connection between the star and disk can remove substantial angular momentum \\citep[e.g.][]{ghoshlamb78, camenzind90, konigl91, shuea94}. When these torques are strong enough to enforce an equilibrium stellar spin rate, this idea is generally referred to as ``disk locking'' \\citep{choiherbst96}. The other idea to explain slowly rotating accretors is that powerful stellar winds are primarily responsible for removing angular momentum from the star \\citep{hartmannmacgregor82, mestel84, hartmannstauffer89, toutpringle92, paatzcamenzind96, mattpudritz05l, mattpudritz08II, mattpudritz08III}. Numerical, dynamical simulations of the star-disk interaction \\citep[such as the investigations of spin equilibrium by][]{romanovaea02, long3ea05} typically exhibit significant torques arising both from the magnetic star-disk connection and from winds. The goal of the present paper is to further examine the possibility of disk locking as an explanation for the slow rotation of young stars. To this end, we develop a model for the time-evolution of stellar spin rates, under the influence of magnetic star-disk interaction torques. In this work, we neglect the influence of stellar winds, as is customary in the disk-locking models. There are generally two types of disk locking models. One is the X-wind model \\citep[e.g.,][]{shuea94, ostrikershu95, mohantyshu08}. This model is developed under the assumption that the star-disk system on average exists in a spin equilibrium state, in which the accreting star feels no torque. Thus, the X-wind model cannot be used to address whether or how the system reaches this equilibrium, nor how the system behaves when out of equilibrium (e.g., due to variability or due to the evolution of the star or disk). All other disk locking models \\citep[e.g.,][]{camenzind90, konigl91, wang95, mattpudritz04} are of the second type, based on the general picture developed by \\citet{ghoshlamb78} for accreting neutron stars. The Ghosh \\& Lamb model provides a method for calculating the torque on the star, due to the star-disk interaction, for any state of the system. The star may spin up or down, depending on conditions, and there is a hypothetical disk-locked state, in which the net torque on the star is zero. Thus, the Ghosh \\& Lamb model can be used to compute the evolution of accreting star spins and to assess how far from equilibrium a system may be. \\citet[][hereafter CC93]{cameroncampbell93}, \\citet{yi94, yi95}, and \\citet[][hereafter AC96]{armitageclarke96} computed the evolution of pre-main-sequence star spins, adopting a Ghosh \\& Lamb-type model for torques on the star, and also including stellar contraction, a decrease in accretion rate with time, and a prescription for the evolution of the magnetic field. For parameters that reasonably represent T Tauri stars, their models produced spin rates within the observed range. These models also showed that, regardless of the initial spin of the star, the spin rates rapidly (in less than $\\sim 1$~Myr) approached an equilibrium value and stayed near equilibrium during the first few Myr of evolution. That is, the initial spin conditions were ``erased,'' and the stellar spin was ``locked'' to a rate that depended only on the stellar mass and radius, magnetic field strength, and accretion rate. These results generally supported the idea that disk locking explains the slow rotation of accreting T Tauri stars. However, since that work, it has been recognized that there is a serious theoretical problem with a key assumption of the classical Ghosh \\& Lamb torque model. The assumption is that the stellar magnetic field lines connect to a large region of the disk and become highly twisted in the azimuthal direction. This twisting is what gives rise to the magnetic spin-down torque felt by the star. Several authors \\citep[e.g.,][]{vanballegooijen94, lyndenbellboily94, lovelace3ea95, hayashi3ea96, millerstone97, goodson3ea97, fendtelstner00, mattea02, uzdensky3ea02, romanovaea02, kuker3ea03, bessolazea08} have now shown that, when the dipole field is twisted azimuthally past a threshold of approximately $45^\\circ$, the magnetic pressure associated with the azimuthal component pushes the field outward. This leads to an inflation and opening of the field loops in a stellar rotation timescale, ultimately disconnecting the star from the disk. \\citet{uzdensky3ea02} developed a method for computing which magnetic field lines would open. The amount of field opening depends primarily on how strongly coupled the magnetic field is to the disk, since the coupling is responsible for the twisting, and the twisting is responsible for the field opening. In order to assess how the field opening affects the torques, \\citet[][hereafter MP05]{mattpudritz05} modified the Ghosh \\& Lamb formulation to include the effects of field opening, following the method of \\citet{uzdensky3ea02}. MP05 showed that, for the relatively strong coupling expected in these systems, the opening of the field significantly reduces the spin-down torque on the star, relative to the classical assumption of a closed field. Thus, the equilibrium spin rate predicted by the disk locking picture is much faster, which calls into question whether disk locking can explain the existence of slowly rotating accretors. However, since the MP05 model could only calculate the torque for a given set of parameters (at a given epoch), their conclusion is based on the calculation of the equilibrium spin rate. MP05 were not able to describe how the spin evolves in time. Therefore, there is a question as to whether the the stars actually evolve near equilibrium, and whether the evolution of the system, including changes in the accretion rate, stellar radius, and stellar spin rate may affect this conclusion. Also, it is important to develop a physical model for understanding the evolution of stellar spin over a range of ages. For these reasons, in this paper, we develop and utilize a stellar spin evolution model \\citep[similar to CC93; AC96;][]{yi94, yi95} that includes the updated torque theory of MP05. In this first effort, we do not attempt to explain all phenomena related to young star spins. Rather, we focus on testing whether models can produce spin rates within the observed typical range of $\\sim1$--10 days. This work extends the MP05 formulation to the time domain, during the first $3\\times 10^{6}$ yr (i.e., during the Hayashi phase). Section \\ref{sec_model} contains a description of the model and details of our calculations. Section \\ref{sec_results} contains the results of stellar spin calculations that include the effect of magnetic field opening. A discussion and summary of the conclusions of this work are contained in section \\ref{sec_conclusion}. As a test of the model, and for illustrative purposes, we include an Appendix containing the results of stellar spin calculations under the classical assumption of a completely closed field. ", "conclusions": "\\label{sec_conclusion} The star-disk interaction model of MP05 includes the effects of the field opening expected to arise from star-disk differential rotation ($\\gamma_c=1$), as well as the effects of magnetic coupling, parameterized by $\\beta$. Based essentially upon the idea that the disk likely exists in a state of high magnetic Reynolds number, MP05 argued that a value of $\\beta=0.01$ is appropriate for real systems. In the present work, we have extended the MP05 torque model into the time domain, in order to explore the consequences of magnetic star-disk coupling and connectedness on the evolution of stellar spins. To this end, we developed a stellar spin evolution model (\\S \\ref{sec_model}). The model follows an accreting, one solar mass star during the Hayashi phase (from $3\\times10^4$ yr to $3$ Myr) and considers the torques expected to arise from the magnetic star-disk interaction. The primary goals of this work were to examine the role of disk locking and to determine if the star-disk interaction torques alone are sufficient to explain the observed range of spin rates. In developing our torque and spin-evolution model, we have adopted many of the same assumptions as the disk-locking models in the literature, except that the MP05 torque formulation includes the effect of the opening of the magnetic field. Thus, while there are inherant uncertainties associated with the adoption of many of these assumptions, our results serve best to highlight the effect of field opening {\\it relative} to previous results in the literature. Given the expectation that $\\gamma_c=1$ and $\\beta=0.01$ best-represents the conditions in real systems, and for a range of accretion rates and field strengths appropriate for T Tauri stars, the two main conclusions of the present work are as follows: 1. The models presented in figures \\ref{fig_g1b500bet01} and \\ref{fig_g1b2000bet01} exhibit spin periods ranging from 3 days to less than 1/3 day, in the age range of 1--3~Myr (see \\S \\ref{sec_gam1}). These are consistent only with the fastest rotators in the observed spin period distribution, which is significantly populated from approximately 1--10~days (see \\S \\ref{sec_intro}). It is apparent that the torque arising from the magnetic connection between the star and disk is not sufficient to explain the relatively slowly spinning, young stars. 2. Furthermore, the stars in these models are generally not in spin equilibrium (with a net zero torque), and the spin rate at all times depends upon the initial spin rate (see \\S \\ref{sec_gam1}). Only the models with the strongest field strength and highest accretion rate neared their equilibrium spin rate in $\\sim1$~Myr (see figure \\ref{fig_g1b2000bet01}). However, this equilibrium had a relatively short spin period of less than a day. It is apparent that disk locking does not play a strong role, except possibly for the most rapid rotators. Note that the term ``disk locking'' generally refers to the idea that the angular spin rate of the star is nearly equal to that of the inner edge of the disk (i.e., $R_t \\approx R_{\\rm co}$). In our magnetic models, this condition was always true. In fact, the truncation radius was nearest to the corotation radius in the models that were furthest from spin equilibrium (e.g., compare panel (d) in figures \\ref{fig_g1b500bet01} and \\ref{fig_ginfb500bet1}). Thus, these stars do have an angular spin rate that is very close to that of the disk inner edge, although it is not appropriate to think of these systems as being ``locked'' to any particular spin rate. Instead, this is a consequence of how the disk is truncated. When the magnetic coupling in the disk is strong (small $\\beta$), the disk will generally be truncated close to the corotation radius (in State 2; see \\S \\ref{sec_rt}). But, the condition that $R_t \\approx R_{\\rm co}$ does not necessarily mean that the star experiences a net zero torque. In order to explain the existence of slowly rotating young stars, it is necessary to consider other possibilities. Since the magnetic torques depend strongly on the magnetic field strength, it is tempting to suggest that a stronger magnetic field may solve the problem. Using our model, we find that (for $\\beta=0.01$) a dipole field strength of $B_*=10^4$~G is required to maintain spin periods consistent with the slow rotators. However, no T Tauri star has been found to have surface field strengths greater than 3~kG \\citep{johnskrull07}, and the global (dipole) fields are generally even weaker \\citep[e.g.,][]{safier98, bouvierea07, donatiea07, donatiea08}. Given the relatively small number of stars for which there are magnetic field measurements, the present situation could be improved by more and improved observations of the global field strength and geometry. Since the models presented in section \\ref{sec_gam1} with weak magnetic coupling (figure \\ref{fig_g1b2000bet1}) exhibit slow spins, it is also appropriate to consider the possibility that real systems have weak coupling. As discussed above, the conditions that are expected to best represent young star-disk systems from ``first principles'' suggest values of $\\beta=0.01$. However, $\\beta$ is a highly uncertain parameter because the physics of magnetic coupling is not well understood, nor are the conditions that influence the coupling (e.g., the ionization fraction of disk gas, turbulence levels in the disk, or magnetic reconnection rates). In order for magnetic coupling to explain the slow rotators, the coupling strength must not be very different (neither larger nor smaller) than $\\beta=1$, since MP05 showed that the maximum spin-down torque occurs for a $\\beta$ of unity. The condition that $\\beta \\approx 1$ seems to require some fine-tuning, which makes this possiblity even more difficult to justify at present. Therefore, it appears necessary to consider effects that are alternative or additional to the magnetic torques arising from the star-disk connection. Real systems seem to be variable on all timescales \\citep[e.g.,][]{hartmann97}, and we have neglected variability here. It may be possible, for example, that short timescale (e.g., $\\la 10^4$~yr), large-magnitude variations of the accretion rate are important for the stellar mass and angular momentum evolution \\citep[e.g.,][]{popham96}. Finally, it is necessary to consider the angular momentum loss from pre-main-sequence stellar winds (see \\S \\ref{sec_intro}). The idea that stellar winds may be important during the accretion phase is well-supported \\citep[e.g.,][]{decampli81, hartmann3ea82, kwantademaru88, hartmannea90, fendt3ea95, fendtcamenzind96, safier98, bogovalovtsinganos01, sauty3ea02, edwardsea03, edwardsea06, dupreeea05, meliani3ea06, kurosawa3ea06, kwan3ea07, fendt09}. Stellar winds may be powered by the accretion process itself \\citep{toutpringle92, cranmer08} and be the key driver of angular momentum loss \\citep{hartmannmacgregor82, mestel84, hartmannstauffer89, paatzcamenzind96, mattpudritz05l, mattpudritz08II, mattpudritz08III}. Some recent magnetohydrodynamic (MHD) simulations of rapidly rotating stars \\citep[e.g.,][]{romanovaea05, romanovaea09} exhibit a type of ``propeller'' regime in which intermittent accretion occurs on typical timescales of several orbits of the inner disk, while most of the would-be accreting material is instead launched into a wind. For the calculations in the present work, the torque represents that which is averaged over a calculation timestep. Our timesteps range from approximately 400 years (at the earliest times) to $\\sim 10^5$ yrs. While some of the simulations in the work cited above have been run for thousands of dynamical times, this is still orders of magnitude shorter than the typical timestep in the present work. It is not clear whether the torque formulation we adopted accurately describes the time-average behavior of the accretion and closed field region in the propeller regime simulations. In addition, while these MHD simulations reliably demonstrate many of the complexities and sometimes episodic nature of the magnetic star-disk interaction, when considering the long-term evolution of the star and disk, it is not yet clear whether the duration of the episodic (e.g., propeller) regime is significant for the overall spin evolution of young stars. However, the MHD simulations such as those cited above often exhibit significant outflows and torques arising from open magnetic field regions. In this respect, our result that the torques arising only from closed field regions are not sufficient to significantly spin-down the star, agrees with the simulations. It is clear that the opening of field lines is an important effect, since including this in the models indicates considerably different physics at work than the closed field models, when comparing to the same observational data. In future work, we will extend the present analysis to include the angular momentum loss from accretion-powered stellar winds." }, "1005/1005.5441_arXiv.txt": { "abstract": "We analyze the most salient cosmological features of axions in extensions of the Standard Model with a gauged anomalous extra $U(1)$ symmetry. The model is built by imposing the constraint of gauge invariance in the anomalous effective action, which is extended with Wess-Zumino counterterms. These generate axion-like interactions of the axions to the gauge fields and a gauged shift symmetry. The scalar sector is assumed to acquire a non-perturbative potential after inflation, at the electroweak phase transition, which induces a mixing of the St\\\"uckelberg field of the model with the scalars of the electroweak sector, and at the QCD phase transition. We discuss the possible mechanisms of sequential misalignments which could affect the axions of these models, and generated, in this case, at both transitions. We compute the contribution of these particles to dark matter, quantifying their relic densities as a function of the St\\\"uckelberg mass. We also show that models with a single anomalous U(1) in general do not account for the dark energy, due to the presence of mixed $U(1)-SU(3)$ anomalies. ", "introduction": "Given its important role as a possible solution of the strong CP problem \\cite{Peccei:1977ur} as well as a candidate for the dark matter of the universe, the study of axions \\cite{Wilczek:1977pj,Weinberg:1977ma} \\cite{Dine:1981rt,Zhitnitsky:1980tq,Kim:1979if,Shifman:1979if} (see \\cite{Sikivie:2006ni} for an overview) has received momentum both at theoretical and experimental level along the years. The invisible axion owes its origin to a global $U(1)_{PQ}$ (Peccei-Quinn, PQ) symmetry which is spontaneously broken in the early universe and explicitly broken to a discrete $Z_N$ symmetry by instanton effects at the QCD phase transition \\cite{Sikivie:1982qv}. The breaking occurs at a temperature $T_{PQ}$ below which the symmetry is nonlinearly realized. Strings and domain walls relics, which are typical of axion models and are a problem in ordinary PQ cosmology, can be avoided by introducing inflation to account for their dilution, or by embedding the model into more general constructions based on theories of Grand Unification \\cite{Lazarides:1982tw}. The almost massless nature of the axion and its suppressed coupling to the fields of the Standard Model are consequences of the fact that this field is associated with the phase of a global anomalous symmetry. Both properties are related to the same scale, the axion decay constant $f_a\\sim 10^{10}-10^{12}$ GeV. The implications of the PQ axion in cosmology, both in supersymmetric and in non supersymmetric models, have been explored to a finer level of detail. For instance, the axion plays an important role in determining the structure of the primordial perturbations \\cite{Dimopoulos:2003ii,Dimopoulos:2003ss,Lazarides:2005ek}, where it can act as a curvaton. The gauging of an anomalous symmetry has some important effects on the properties of this pseudoscalar, first among all the appearance of independent mass and couplings to the gauge fields. This scenario allows a wider region of parameter space where to look for these particles. For this reason, axion-like fields, which are at the center of several investigations, are unlikely to find any significant and fundamental formulation without an underlying anomalous $U(1)$ gauge symmetry, as emphasized in previous works~\\cite{Coriano:2006xh,Coriano:2007xg}, \\cite{Coriano:2009zh}. So far only two complete models have been put forward for a consistent analysis of these types of particles, the MLSOM \\cite{Coriano':2005js} and the USSM-A \\cite{Coriano:2008xa}. The first of them is at the basis of the elaborations that we are going to provide in this work. Here we will be focusing on the phenomenological analysis of a scenario which is a direct consequence of the model introduced in \\cite{Coriano':2005js}, while more details on the supersymmetric construction will be discussed in a separate work. Although the natural framework that motivates these constructions is open string theory \\cite{Angelantonj:2002ct}, the effective actions describing these types of particles can be consistently defined at lower energy just by the inclusion of the relevant dimension-5 Wess-Zumino (Peccei-Quinn) interactions. These are necessary in order to guarantee the gauge invariance of the effective action and can be interpreted as counterterms. In fact, they balance the anomalous variation of the 1-loop effective action induced by the extra $U(1)$ symmetry, restoring the gauge symmetry. The gauging of an anomalous symmetry is the essential element in the construction of these effective actions and can be justified within intersecting brane models. The gauging is a variant of the standard Peccei-Quinn construction and is characterized by a new scale $M$, which is the St\\\"uckelberg mass. We recall that St\\\"uckelberg extensions of the Standard Model with a non-anomalous $U(1)$ have been analyzed in several recent works \\cite{Feldman:2007wj,Feldman:2006wb,Feldman:2010wy}. We are going to provide a physical perspective on the possible phenomenological implications of the anomalous case. In particular, we will try to connect the St\\\"uckelberg fields, which are in the spectrum of these models, to the physical axion which may appear as an extremely weakly interacting particle of a certain relic density in our current universe. Our assumption, in the identification of the physical axion, is that the original St\\\"uckelberg fields will mix at the electroweak phase transition with the Higgs sector. As a result, an almost massless state will emerge after the electroweak phase transition. One of the key mechanisms that we will try to adapt and extend from the PQ case is that of {\\em vacuum misalignment}. This phenomenon occurs whenever a quasi Nambu-Goldstone mode - generated by the breaking of a certain symmetry - acquires non-perturbatively a small potential, lifting one flat direction from the vacuum degeneracy. For axions characterized both by an $SU(2)$ and an $SU(3)$ charge the mechanism of vacuum misalignment becomes sequential, as we are going to show. From a more general perspective, we will also try to characterize the possible role of these types of particles as quintessence axions. These appear in models where the axions remain decoupled from the gluonic sector and their mass is purely of electroweak origin (see for instance \\cite{Nomura:2000yk}). In this case one tries to exploit the Nambu-Goldstone nature of these particles. The main idea behind this proposal is that a phase transition around the electroweak scale can generate a small curvature in the potential, capable of giving a tiny mass to this particle, smaller than the Hubble parameter at current time ($H_0$). For this to be possible, as we are going to show, one has to search for solutions of the anomaly equations for an anomalous $U(1)$ which has a vanishing mixed anomaly with the $SU(3)$ color group. In this case the only source of mass for these axions would come from the electroweak and not from the QCD phase transition, and as such could be extremely small. In the general models that we analyze, the anomaly equations do not allow for such a solution, although this would not exclude the possibility of finding others, in the presence of more complicated gauge structures, for instance in models with several U(1)'s. We will not address this specific point any further, leaving it as an option for future studies. Instead, we will concentrate on the general features of an axion-like field coming from a single anomalous $U(1)$ symmetry, characterized by the presence of mixed anomalies both with the $SU(2)$ and $SU(3)$ sectors. The phenomenological details of the model are rather intricate, and have been worked out before. For this reason we have summarized in the next section some of their salient features, which turn out to be necessary in order to proceed with a realistic estimate of the relic densities. This is the specific goal of our work. ", "conclusions": "We have discussed the most salient cosmological features of models containing gauged axions, obtained from the gauging of an anomalous symmetry. The gauging allows to define a consistent theory for axion-like particles, which generalize many of the properties of PQ axions. They have appeared for the first time in the study of intersecting branes, but their features are quite generic. They are constructed as effective theories containing minimal gauge interactions which restore gauge invariance of the effective action in the presence of an anomalous $U(1)$ symmetry, and no further requirements. Differently from the PQ case, here there is no concept of an original PQ symmetry, broken at a very large scale, with the axion taking the role of a Goldstone mode that acquires a mass at the QCD phase transition. Rather, the physical axion emerges directly at the electroweak phase transition, when Higgs-axion mixing occurs. Being charged under $SU(3)$ and $SU(2)$, we have a sequential misalignment of this field, and we have quantified its relic density as a function of the St\\\"uckelberg mass. We have shown that only very large values of the St\\\"uckelberg mass cause a significant contribution of this type of axions to the current dark matter content of the Universe, which otherwise remains negligible. The absence of an original PQ-like potential has some implications at cosmological level, such as the absence of isocurvature perturbations, since the St\\\"uckelberg is not a physical mode before the electroweak phase transition, in particular at the time of inflation. This feature is due to the presence of a local gauge symmetry, realized in the St\\\"uckelberg form, which allows to absorb $b$ into the longitudinal component of the anomalous gauge boson. Our analysis represents, more generally, a description of the fate of the St\\\"uckelberg field in cosmology, from the defining St\\\"uckelberg phase of the theory at a large scale (defined by the value of the St\\\"uckelberg mass), down to the electroweak and QCD phase transitions, when this field develops a physical component. Our analysis could be extended in several directions, for instance with the inclusion of the modifications induced on the computation of the relics due to the presence of non-homogeneities in $\\chi$ beyond the QCD horizon, a feature which is also present in PQ models when the PQ scale lays below the scale of inflation. However, even at this level of refinement, the only significant scale in the determination of the relic densities remains the value of the St\\\"uckelberg mass. Small values of this mass parameter in the TeV range leave the contribution of these particles to the relic densities of dark matter negligible, and sizeable for $M$ around an intermediate scale of $10^7$ GeV. In this case all the constraints coming from the neutral current sector are satisfied, being the extra $Z^\\prime$ of the theory completely decoupled from the low energy spectrum of the Standard Model. The appearance of this intermediate scale is a novel feature of this type of axions which could be used to set limits on their parameter space. \\vspace{1cm} \\centerline{\\bf Acknowledgements} We thank Pierre Sikivie and Nikos Irges for discussions. C.C. thanks the Physics Department at Thessaloniki for hospitality. This work is supported in part by the European Union through the Marie Curie Research and Training Network UniverseNet (MRTN-CT-2006-035863). \\begin{appendix}" }, "1005/1005.3784_arXiv.txt": { "abstract": "{We present a preliminary analysis of the small-scale structure found in new 70-520 \\um\\ continuum maps of the Rosette molecular cloud (RMC), obtained with the SPIRE and PACS instruments of the {\\it Herschel} Space Observatory. We find 473 clumps within the RMC using a new structure identification algorithm, with sizes up to $\\sim$1.0 pc in diameter. A comparison with recent {\\it Spitzer} maps reveals that 371 clumps are ``starless\" (without an associated young stellar object), while 102 are ``protostellar.\" Using the respective values of dust temperature, we determine the clumps have masses ($M_{C}$) over the range -0.75 $\\leq$ log~($M_{C}$/\\msun) $\\leq$ 2.50. Linear fits to the high-mass tails of the resulting clump mass spectra (CMS) have slopes that are consistent with those found for high-mass clumps identified in CO emission by other groups.} ", "introduction": "\\label{intro} Stars form within molecular clouds, after some fraction of cloud material is first condensed into smaller-scale structures. How this process unfolds is not well understood, however. For example, what conditions within clouds drive the formation of small-scale structures that in turn produce stars of any given mass, e.g., high-mass stars. Detailed observations of the small-scale structure within molecular clouds should provide valuable insight into how stars of various masses form out of dense material. The need for sensitive probes of small-scale structure in molecular clouds is the impetus behind the ``{\\it Herschel}\\/ OB Young Stellar objects\" (HOBYS) Key Project (see Motte et al. 2010). The HOBYS team is currently using the ESA {\\it Herschel} Space Observatory (Pilbratt et al. 2010) to obtain wide-field maps of $\\sim$15 molecular clouds which are forming high-mass stars within 3 kpc of the Sun. These clouds are being observed with both the {\\it Herschel} SPIRE (Griffin et al.\\/ 2010) and PACS (Poglitsch et al.\\/ 2010) instruments, ultimately to obtain wide maps of these clouds at 70-520 \\um\\ at diffraction-limited resolutions of $\\sim$18\\as\\ $\\times$ ($\\lambda$/250 \\um). Such data sample thermal emission from cold dust grains mixed with molecular gas within the cloud. Since continuum emission at submillimetre wavelengths is typically optically thin, detections of continuum emission from dust at high resolution can trace the organization of mass in clouds on small scales. By sensitively sampling emission across the peaks of the spectral energy distributions (SEDs) of this dust, SPIRE and PACS data can constrain both the column density and temperature of the dust and provide unparalleled censuses of the small-scale structure in molecular clouds. In this paper, we describe the small-scale structure detected with SPIRE and PACS in the Rosette molecular cloud (RMC), the first of the HOBYS target list that was observed as part of the early {\\it Herschel} ``science demonstration phase\" (SDP) campaign. The RMC is a rich location of star formation at a distance of 1.6 kpc (Johnson 1962; P\\'erez et al. 1987; Park \\& Sung 2002). The cloud is southeast of NGC 2244, the Rosette Nebula, and the expanding HII region associated with NGC 2244 has begun to interact with it. The structure of material within the RMC has been previously examined by Williams, Blitz \\& Stark (1995), Schneider et al.\\/ (1998) and Dent et al.\\/ (2009) using various CO observations. Alternatively, the populations of young stellar objects (YSOs) within the RMC have been studied by Phelps \\& Lada (1998), Li \\& Smith (2005), Poulton et al.\\/ (2008), and Rom\\'an-Z\\'u\\~niga et al.\\/ (2008) using infrared observations. Notably, these studies identified 10 embedded clusters within the RMC, i.e., PL01-07 and REFL08-10. A recent review of observations of the RMC has been made by Rom\\'an-Z\\'u\\~niga \\& Lada (2008; RZL). The SPIRE and PACS observations described in this paper provide a preliminary look into how {\\it Herschel}\\/ data trace small-scale structures in the RMC, and the relationships between these structures and YSOs. Analyses of other aspects within the SDP data of the RMC can be found in the accompanying papers by Motte et al.\\/ 2010, Henneman et al.\\/ 2010, and Schneider et al.\\/ 2010.) ", "conclusions": "\\label{summary} We have conducted a preliminary analysis of small-scale emission within the Rosette molecular cloud, as observed by the {\\it Herschel} Space Observatory. From SPIRE and smoothed PACS images, we identified 473 clumps using {\\it getsources}, a new algorithm that identifies structure over multiple wavelengths and scales. Using {\\it Spitzer}\\/ data, we classified these into 371 starless and 102 protostellar clumps. The clumps have dust temperatures of 10-30 K, with starless clumps surprisingly warmer on average than the protostellar clumps, owing to the proximity of many to NGC 2244. Masses were determined for each clump, revealing a range of -0.75 $\\leq$ log~($M_{C}$/\\msun) $\\leq$ 2.50. Linear least-squares fits to the high-mass tails of the CMS reveal slopes that are consistent with slopes fit to mass spectra of clumps identified through CO in the RMC. Such slopes are shallower than those seen in the high-mass tails of the mass spectra of small-scale continuum structures (``cores\") in closer clouds, possibly from crowding of such objects on scales smaller than probed here." }, "1005/1005.3267_arXiv.txt": { "abstract": "{} {It was shown in previous works the existence of weakly chaotic orbits in the plutino population that diffuse very slowly. These orbits correspond to long-term plutino escapers and then represent the plutinos that are escaping from the resonance at present. In this paper we perform numerical simulations in order to explore the dynamical evolution of plutinos recently escaped from the resonance. } {The numerical simulations were divided in two parts. In the first one we evolved $20,000$ test particles in the resonance in order to detect and select the long-term escapers. In the second one, we numerically integrate the selected escaped plutinos in order to study their dynamical post escaped behavior. } {Our main results include the characterization of the routes of escape of plutinos and their evolution in the Centaur zone. We obtained a present rate of escape of plutinos between 1 and 10 every 10 years. The escaped plutinos have a mean lifetime in the Centaur zone of $ 108$ Myr and their contribution to the Centaur population would be a fraction of less than $6 \\%$ of the total Centaur population. In this way, escaped plutinos would be a secondary source of Centaurs. } {} ", "introduction": "In the last few years the number of observed transneptunian objects has enormously grown thanks to the progresses in the astronomical observations. This fact has allow to define, in a more rigorous way, the different dynamical classes previously identified in the first years of discoveries. The transneptunian region (TNR) can be structured into 4 dynamical classes (Chiang et al. \\cite{Chiang07}). The Resonant Objects are those in mean motion resonance with Neptune, the Classical Objects are those non-resonant objects with semimajor axis $a$ greater than $\\sim$42 AU and low eccentricity orbits, the Scattered Disk Objects (SDOs) with perihelion distances $q > 30$ AU and large eccentricities, and the Centaurs Objects. This last group has perihelion distances inside the orbit of Neptune and they are transitory objects descendants of the other 3 classes, mainly from the SDOs, recently dislodge from the transneptunian zone by planetary perturbations (Levison \\& Duncan \\cite{Levison97}, Tiscareno \\& Malhotra \\cite{Tiscareno03}), Di Sisto \\& Brunini \\cite {Disisto07} ). Centaurs are sometimes defined according to their aphelion distance or to their semimajor axis ($a$), as for example the nomenclature of Gladman et al \\cite{Gladman08} that uses $a 1$ km every 10000 years or a flux rate of escape of 0.5 $\\%$ of plutinos in $10^{10}$ years. Very recently, Tiscareno \\& Malhotra (\\cite{Tiscareno09}) carried out 1-Gyr numerical integrations to study the characteristics of the 2:3 and 1:2 mean motion resonances with Neptune. Their main results include maps of resonance stability for a whole range of eccentricities and inclinations. They made integrations with and without Pluto, and concluded that it has only modest effects on the Plutino population. They calculated the fraction of remaining plutinos in the resonance as a function of time and extrapolated this fraction after 4 Gyr and also evaluated the fate of escaped particles. From those previous works, plutino removal by ``dynamic'' is much greater than plutino removal by collisions. We refer to ``dynamical'', those numerical simulations that take into account the gravitational forces to follow up the evolution of a particle and occasionally cause the removal. Then in this work, we perform ``dynamical'' numerical simulations in order to describe and characterize the routes of escape of plutinos and their contribution to the other minor bodies populations of the Solar System, specially to Centaurs. ", "conclusions": "We have performed two numerical simulations in order to first obtain particles representative to the plutinos that are escaping at present from the resonance and second in order to describe their dynamical post escape evolution. In the first simulation we integrate $20,000$ initial particles in the 2:3 resonance and find that $\\sim 88 \\%$ of the particles left out the integration and the rest remain in it. Considering a plutino population with radius grater than 1 km of $N_p \\sim 10^8 - 10^9$, we obtained a present rate of escape of plutinos between 1 and 10 every 10 years. From this integration we selected those particles that are representatives of the present escape plutinos and performed a second integration. From this last integration we obtained the dynamical evolution of plutinos once they escape from the resonance. From the 1183 initial particles, 1179 were removed from the integration and 4 remain in it. From the 1179 removed particles, 787 are ejected, 385 reached the Jupiter's zone and 4 collide with the planets. We found that the great majority of escaped plutinos have encounters with Neptune, and this planet governs their dynamical evolution. When a plutino escape from the resonance, it is transferred to the SD zone ($q>30$ AU) or to the Centaur zone ($q<30$ AU) but it eventually switches to those population, due to the dynamical influence of Neptune. The densest zone in the orbital element space of escaped plutinos corresponds to the ranges $30 < a < 100$ AU and $5^{\\circ} < i < 40^{\\circ}$ and perihelions near the orbit of Neptune. When escaped plutinos are transferred to the SD they are quickly locked into a mean motion resonance with Neptune (similar to the behavior of SDOs analyzed by Fern\\'andez et al. (\\cite{Fernandez04}) and Gallardo (\\cite{Gallardo06}). In the Centaur zone (this is the zone of $q < 30$ AU ) the distribution of escaped plutinos is similar to that of SDOs in the Centaur zone obtained by Di Sisto \\& Brunini (\\cite{Disisto07}). The orbital evolution of escaped plutinos in the Centaur zone can be grouped into the four dynamical classes proposed by Di Sisto \\& Brunini (\\cite{Disisto07}). There are more particles that have the dynamical behavior of the second class and it is notable the great frequency of the presence of Kozai resonances in all the four classes. There are also several mean motion resonances densely populated in the ranges of $30 < a < 50$ AU. The escaped plutinos have a mean lifetime in the Centaur zone of $ 108$ Myr, greater than that of Centaurs from SD of $ 72$ Myr . Escaped-plutinos live more time than SDOs in the greater-perihelion Centaur zone, causing a slower diffusion to the inner Solar System of escaped-plutino orbits than of SDOs orbits. The present rate of injection of plutinos with radius greater than 1 km to the Centaur zone is between 1.6 to 16 plutinos every 100 years and the number of plutino-Centaurs with radius greater than 1 km would be between $1.8 \\times 10^{6} - 1.8 \\times 10^{7}$. Both, the rate of injection and the number of Centaurs from plutinos are much less than the contribution from the SD obtained by Di Sisto \\& Brunini (\\cite{Disisto07}). Then, plutinos would represent a secondary source of Centaurs and their contribution would be a fraction of less than $6 \\%$ of the total Centaur population. \\vspace*{1cm} \\noindent{\\bf Acknowledgments:} We thank Matthew S. Tiscareno who, as referee, made valuable comments that helped to improve this manuscript." }, "1005/1005.1865_arXiv.txt": { "abstract": "{} {We study the structure of the medium surrounding sites of high-mass star formation to determine the interrelation between the H~{\\sc ii} regions and the environment from which they were formed. The density distribution of the surrounding{s} is key in determining how the radiation of the newly formed stars { interacts with the surrounds in a way that allows it to be used as a star formation tracer}. } {We present new {{\\it Herschel/SPIRE}} 250, 350 and 500~$\\mu$m data of LHA 120-N44 and LHA 120-N63 in the LMC. We construct average spectral energy distributions (SEDs) for annuli centered on the IR bright part of the star formation sites. The annuli cover $\\sim$10$-$$\\sim$100 pc. We use a phenomenological dust model to fit these SEDs to derive the dust column densities, characterise the incident radiation field and the abundance of polycyclic aromatic hydrocarbon molecules. We see a factor 5 decrease in the radiation field {energy} density {as a function of radial distance} around N63. N44 does not show a systematic trend. We construct a simple geometrical model to derive the 3-D density profile of the surroundings of these two regions.} {{{\\it Herschel/SPIRE}} data {have proven} very efficient in deriving the dust mass distribution. We find that the radiation field in the two sources behaves very differently. N63 is more or less spherically symmetric and the average radiation field drops with distance. N44 shows no systematic decrease of the radiation intensity which is probably due to the {inhomogeneity} of the surrounding molecular material and to the complex distribution of several star forming clusters in the region.} {} ", "introduction": "\\label{sec:introduction} High-mass star formation (SF) sites (hereafter HMSFSs) are the beacons by which we probe a large part of the physics of external galaxies. They generally represent the most important tracers of the properties of their host galaxies in terms of star formation rate (SFR) and general activity. The most frequent tracers of the star-formation activity generally use the fact that the abundant UV light coming from the hot, young stars is absorbed in the vicinity and reradiated in the form of line or continuum emission. This is true, for example, for H$\\alpha$ \\citep[e.g.][]{1998ARA&A..36..189K}, the aromatic emission bands in the mid-IR \\citep[e.g.][]{2007ApJ...666..870C} or the IR continuum due to solid-state materials \\citep[dust, e.g.][]{1986ApJ...303L..41S}. These tracers work relatively well and are used to characterise nearby star forming regions and star forming galaxies out to large redshifts, although interesting discrepancies have been noted for dwarf galaxies at low SFR \\citep[see][]{2009ApJ...706..599L}. One of the main assumptions that enters into the quantitative interpretation of these data is the geometry of the material surrounding the newly formed stars, in particular, where the UV light is being reprocessed. For example, if the UV photons escape from the ionised medium this may boost the aromatic feature strengths and strongly influence the lines originating from the surrounding photo-dissociation regions (PDRs). There are indications that this geometry in external galaxies may qualitatively and quantitatively differ from that observed in the Milky Way (MW) \\citep[e.g.][]{2006A&A...446..877M,2009A&A...508..645G}. One simple effect may be that at different metallicities the surrounding medium is more or less opaque and therefore the UV photons have a different mean free path. More complex scenarios are also discussed in the literature. For example, clumpiness of the molecular cloud may lead to small molecular cores surrounded by large PDRs. {\\it Herschel} with its unprecedented wavelength coverage and angular resolution at submillimeter (submm) wavelengths provides a unique opportunity to probe the cold interstellar medium (ISM) and sample the effects of the environment on the resulting SF tracers. In particular, it is well suited to trace the distribution of matter around HMSFSs and to map the way the UV radiation permeates and heats the surroundings. Here we present a study of two HMSFSs in the Large Magellanic Cloud (LMC) based on data taken in the HERITAGE program \\citep[PI. Meixner, see][]{meixner_special_issue}. ", "conclusions": "\\label{sec:disc--concl} Figs.~\\ref{figResults} and \\ref{geometry} show the power of {\\it Herschel} to determine the matter distribution around HMSFSs. We investigate the effect the new SPIRE constraints on the derived parameters and their uncertainties by also fitting the SEDs without using these SPIRE data. We find that the column density of dust is often very discrepant (by more than an order of magnitude) from the values derived using the SPIRE data. The derived $\\Sigma_{dust}$ profile for N44 (like Fig.~\\ref{figResults} top panel) is virtually constant at a value of 0.1~M$_{\\sun}$\\,arcsec$^{-2}$ due to the fact that the cooler dust, is not well traced by the MIPS data. For N63 the model over-predicts the derived dust masses for the annuli between 20 to 60 pc by a large factor ($\\sim$5). In these cases the model tends to find very small values of $U_\\mathrm{min}$. This artifact could perhaps be circumvented, when modelling those kind of regions, in cases where submm constraints are missing, by limiting the allowed range of $U_\\mathrm{min}$ in Eq.~(\\ref{eqn:dale}). The difficulty of constraining the radiation field parameters without the SPIRE bands is also reflected in the derived uncertainties. In particular the value of $U_\\mathrm{min}$ is ill-constrained which results in large, asymmetric error-bars on the derived column densities. N44 and N63 exhibit strikingly different behaviour in the radiation intensity profile (Fig.~\\ref{figResults}{\\bf d}.) The lack of a systematic decrease of $<$U$>$ around N44 indicates that we are observing dust with a wide range of temperatures along each line-of-sight. The inner annuli in N44 are affected by the superbubble to the NE of the OB association, where high values of $<$U$>$ are expected. The low values of $<$U$>$ for such a luminous SF region may reflect clumpiness. The profile is clearly incompatible with a centrally illuminated optically thin irradiation profile. It is clear from Fig.~\\ref{figMaps} that the studied regions are not very spherically symmetric (azimuthally smooth). In particular, N44 harbours several clusters and the peak of the X-ray emission is located in a cavity, $\\sim$ 20$-$30~pc away from the TIR peak (see Fig.~\\ref{figMaps}). Measuring the azimuthally averaged properties smears out some of the characteristics. This smearing could have been the cause for the lack of trend seen in the average $U$ as seen by the dust (Fig.~\\ref{figResults}). We have verified that this small range of $<$U$>$ is not simply an artifact of this averaging or a wrong choice of center of the annuli by studying the parameters we derive pixel by pixel in the maps which makes no assumptions about the geometry. Indeed the highest $<$U$>$ in N44 is found close to the center we chose. Except for the very center all other values, with their scatter, are within the range as depicted in Fig.~\\ref{figResults}. We conclude that the choice of center does not dominate the lack of trend of N44 in the average radiation field. Thus this lack of trend reflects the true broad range of irradiation conditions along all lines of sight in N44 , which is an indication of the inhomogeneity of the ISM around N44. A simple dust model shows a deficit in PAHs toward the centers of these two regions. We find no evidence for a submm excess. We have used the observed dust column densities surrounding N44 and N63 to derive a 3-D model for these regions for the first time." }, "1005/1005.4995_arXiv.txt": { "abstract": "Pulsars are believed to be magnetized neutron stars. Their surface magnetic field ranges from $10^8$ to $10^{12}$ G. On the other hand, the magnetars have surface magnetic field $10^{14}-10^{15}$ G. It is believed that at the center, the magnetic field may be higher than that at the surface. We study the effect of the magnetic field on the neutron star matter and hence on the mass-radius relation of neutron stars. We model the nuclear matter with the relativistic mean field approach considering the possibility of appearance of hyperons at higher density. We find that the effect of magnetic field on the matter of neutron stars and hence on the mass-radius relation is important when central magnetic field $\\ge 10^{17}$ G. Moreover, if the central field is of the order of $10^{19}$ G, then the matter becomes unstable which limits the maximum magnetic field at the center of magnetars. ", "introduction": "Soon after the discovery of radio pulsar \\cite{hewish}, the theoretical proposition of neutron star (NS) \\cite{bz} drew much attention. At the core of a NS, the matter, composed of neutrons, protons, electrons, and sometimes muons, is highly degenerate with density in the range of $10^{14}-10^{15}$ gm/cc. Naturally, the constituent particles therein interact via nuclear forces, forming a non-ideal nuclear fluid. However, the nature of nuclear forces at this high density and the internal composition, especially the composition of matter in the inner most core of the NS, are not well understood yet. There are many theoretical models of nuclear matter at high density describing different equations of state for the matter. Based on different equations of state, different mass-radius relations of NS are obtained. Consequently, only way to constrain the equations of state is comparing the theoretical results with observed properties of NS. With the detection of pulsars, it was also observed that the pulse period slowly increases in a very regular manner. Consequently, pulsars were successfully modeled as magnetized rotating NSs \\cite{pacini,gold,og}. This model successfully describes the spin-down power of the Crab pulsar and other pulsars. This description requires a surface magnetic field $10^{11}-10^{13}$ G for the pulsars detected at the beginning. Later on, more observations broadened this range of the magnetic field with the discovery of millisecond pulsars \\cite{backer}. This class of pulsars has a much lower surface magnetic field of $10^8-10^{10}$ G \\cite{alpar}. In the recent past pulsars with a very high rate of change of period (period derivatives) have been observed \\cite{morriset,mcet}, implying a high surface field up to $10^{14}$ G. Another class of pulsars has been observed which is the accreting X-ray pulsar \\cite{parmar}. For this class, the surface magnetic field varies from $10^8$ to $10^{13}$ G. Anomalous X-ray pulsars (AXPs) were discovered in early 80s, which show periods in a relatively narrow range of $5-11$ s and have large period derivatives \\cite{vg}. In early days, they were believed to be unusual accreting NSs, whereas, later surveys detected no companions for these candidates. The large value of period derivatives implies a surface magnetic field as high as $10^{14}-10^{15}$ G. This helps in explaining the X-ray luminosities, powered by magnetic field decay, without any accretion. However, until the discovery of soft $\\gamma$-ray repeaters (SGRs), such high values of surface magnetic field were not acceptable. SGRs show brief intense bursts of soft $\\gamma$-rays. From the peak luminosities of the bursts, SGRs are identified as NSs with very strong surface magnetic field ($\\sim 10^{14}-10^{15}$ G) \\cite{nature2,woodset}. Now both AXP and SGR are believed to be strongly magnetized NSs with different variations, which are popularly known as magnetars \\cite{vg,dt,usov,td9}. The existence of a magnetar motivates to study the effects of strong magnetic field on NS properties. A strong magnetic field affects, the structure of a NS through its influence on the underlying metric \\cite{bbgn,cpl} and equation of state (EoS) through the Landau quantization of charged particles and then the interaction of magnetic moments of charged particles (and even the anomalous magnetic moment (AMM) of neutral particles) with the magnetic field. In the present paper, we plan to study the effect of magnetic field on the EoS of highly dense matter. Generally, the effect of magnetic field on the EoS of matter is significant when the magnetic field ${\\cal B}>10^{18}$ G. For the nuclear matter with a $n$-$p$-$e$ system, the effect of magnetic field was studied by several authors \\cite{cbs,yz,bpl1,czl,wmkksg}. However, the composition of the core of a NS is very uncertain as the density reaches $10-15$ times the normal matter density. At this high density, the quarks may be deconfined, making quark matter, or the hyperons may appear, making hyperonic matter. Another hypothesis is that the whole star is made of deconfined quark matter, which is known as strange star (SS). The effect of magnetic field on quark matter using the MIT bag model has been studied in different literature \\cite{chakra,gc,fmro}. There are other models of quark matter with phenomenological density dependent quark masses \\cite{fowler,chak,d98,lxl}. Broderick et al. \\cite{bpl2} studied the effect of strong magnetic field on hyperonic matter. In their study, the field strength does not depend on density. However, in realistic situation, it is believed that the field strength is higher at core than that at surface of a NS. Therefore, with radius, and hence with density, the field strength should vary. In the present work, we plan to study the influence of the strong magnetic field on the hyperonic matter and hence on the mass-radius relation of NSs with hyperonic matter in the core with a density dependent magnetic field. The paper is organized as follows. In the next section we describe the model of hadronic matter at high densities in the presence of magnetic field. Subsequently, we discuss the numerical results in section \\ref{discuss}. Finally, we summarize the results in the last section. ", "conclusions": "\\label{discuss} Nucleon-meson coupling constants are chosen in such a way that the nuclear matter properties can be reproduced as the binding energy $E/B=-16.3$ MeV, the saturation density $n_0=0.153$ fm$^{-3}$, the asymmetry energy coefficient $a_{asy}=32.5$ MeV and the incompressibility $K=240$ MeV, and are taken from previous work \\cite{glend8}. The hyperon-$\\omega$ coupling constants are determined from SU(6) symmetry of the quark model \\cite{dg,sdgmgs,sm}. The hyperon-$\\sigma$ coupling constants, however, are determined from the potential depth of hyperons in normal nuclear matter given by \\begin{equation} U_Y=-g_{\\sigma Y} \\sigma+g_{\\omega Y} \\omega_0, \\end{equation} where $Y$ corresponds to hyperon. We take the potential depth for $\\Lambda$-hyperon $U_\\Lambda=-30~MeV$, as obtained from the analysis of $\\Lambda$-hypernuclei \\cite{dg,fet}. From various experimental data \\cite{fet,ket} for $\\Xi$-hypernuclei, we obtain the potential depth for $\\Xi$-hyperon $U_\\Xi=-18~MeV$. Recent $\\Sigma$-hypernuclei data indicate that the corresponding potential depth to be repulsive \\cite{bet} and we take $U_\\Sigma= 30~MeV$. Note importantly that earlier work \\cite{bpl2} assumed hyperon-meson coupling constants to be fixed fractions of respective nucleon-meson coupling constants and for all the hyperons these fractions are same for a given meson. Hence, unlike our model, in that work the first appeared hyperon is $\\Sigma^+$, which does not appear at all (see Table \\ref{threshold}) in our model. In a realistic situation, the magnetic field at the core may be higher than that at the surface and gradually may decrease towards the surface. For the magnetars, the maximum surface magnetic field is $10^{15}$ G. For this case, the core magnetic field may be of the order of $10^{17}-10^{18}$ G. Without any proper knowledge of magnetic field configuration inside a NS, following previous work \\cite{bcs} here we adopt a magnetic field profile \\begin{equation} {\\cal B}\\left(n_b/n_0\\right)={\\cal B}_s+{\\cal B}_c\\left\\{1-e^{-\\beta \\left( \\frac {n_b}{n_0} \\right)^\\gamma}\\right\\}, \\label{magprfl} \\end{equation} where $\\beta$ and $\\gamma$ are two parameters determining the magnetic field profile with given ${\\cal B}_s$ and ${\\cal B}_c$, and $n_b$ is the total baryon number density. We consider different values of field at the center, whereas surface field strength is taken to be ${\\cal B}_s=10^{15}$ G. We notice that the different values of ${\\cal B}_s$ do not affect the EoS considerably, and thus we do not show the results with other ${\\cal B}_s$. In the above parametrization, the magnetic field strength depends on the baryon number density. However, at each density the field is uniform and constant. \\subsection{Equation of states} We compute the EoS self-consistently with the condition of beta-equilibrium and charge neutrality. It is found that, as described by the Figures given below, the effect of magnetic field is important for ${\\cal B}_c \\ge 10^{17}$ G. However, for ${\\cal B}_c < 10^{18}$ G, the effect is significant only when the field reaches ${\\cal B}_c$ at a very low density, away from the center, and remains almost constant up to center. Therefore, we restrict our study with ${\\cal B}_c \\sim\\,10^{18}$ G. At lower densities, the matter is composed of only neutrons, protons and electrons. Hence, at the low density regime, the particles which are affected by the magnetic field are electrons and protons. Since the electrons are highly relativistic, electron Fermi momentum is very large compared to electron mass. Therefore, the number of occupied Landau levels by electrons is very large, eventhough the field strength under consideration is larger than the critical field strength of electron by several orders. On the other hand, the field strength under consideration is very less than the critical field strength of protons. Consequently, the number of occupied Landau levels by protons is also large. As density increases, the heavier particles appear gradually. In addition, the magnetic field increases with the increase of density. As a result, the number of occupied Landau levels gradually decreases for every species. The threshold densities for muons and other hyperons to appear for ${\\cal B} _c=10^{18}$ G are given in Table \\ref{threshold}. The threshold densities for various species to appear do not differ from their respective values when the magnetic field is absent. \\begin{table}[t] \\begin{center} \\begin{tabular}[t]{|c|c|} \\hline \\multicolumn{2}{|c|}{Threshold densities} \\\\ \\hline $\\mu^-$ & 0.9 \\\\ \\hline $\\Lambda$ & 2.6 \\\\ \\hline $\\Xi^-$ & 3.1 \\\\ \\hline $\\Xi^0$ & 5.7 \\\\ \\hline \\end{tabular} \\caption{Threshold densities for muons and hyperons to appear in units of $n_0$.} \\label{threshold} \\end{center} \\end{table} In Fig. \\ref{eos} we show the EoSs with and without magnetic field when the magnetic field profile corresponds to $\\beta =0.1$ and $\\gamma=1$ [see Eq. (\\ref{magprfl})]. We observe that the EoS becomes softer with the increase of ${\\cal B}_c$. Here we should mention that, when the field strength is high enough, the field energy and the field pressure are not negligible. In calculating the EoS, we thus add this contribution too which is necessary to construct the structure of NSs. \\begin{figure}[t] \\begin{center} \\includegraphics[angle=-90,width=4.5 in]{p-ndiffb.eps} \\caption{Variation of $P$ as a function of normalized baryon number density. The solid curve is for without magnetic field. The dashed, dotted and dot-dashed curves correspond to ${\\cal B}_c=10^{18}$, $2\\times10^{18}$ and $3\\times 10^{18}$ G respectively.} \\label{eos} \\end{center} \\end{figure} \\subsection{Mass-radius relations} With the above EoSs, the structure of NSs with and without rotation is constructed. In this preliminary study of NS structure, we do not include anisotropic solutions based on the general relativity which is important in strong magnetic field. Without any rotation, the Einstein's equation leads to the Tolman-Oppenheimer-Volkoff (TOV) equation \\begin{equation} \\frac{dP}{dr} = -\\frac{[\\varepsilon(r)+P(r)][M(r)+4\\pi r^3P(r)]} {r[r-2M(r)]}, \\end{equation} where $r$ is the distance from the center of the star and $M(r)$ the mass enclosed within $r$. The radius of a star ($R$) is the distance from the center at which the pressure becomes zero. For different central densities, stars with different masses and radii are formed. In Fig. \\ref{m-r} the mass-radius relations for static stars which correspond to EoSs described in Fig. \\ref{eos} are shown. Since the effect of magnetic field softens the EoS, in the presence of magnetic field the maximum mass is less than that for without magnetic field. Stars with rotation, as expected, have larger radii than their static counterpart. In Figs. \\ref{m-rr1} and \\ref{m-rr2} we show the corresponding mass-radius relations for rotating NSs. In both the cases, the maximum masses in presence of the magnetic field are less than that in absence of the magnetic field. \\begin{figure}[t] \\centering \\includegraphics[angle=-90,width=4.5 in]{m-rst.eps} \\caption{Mass-radius relations without rotation. The solid curve is for without magnetic field. The dashed, dotted and dot-dashed curves correspond to ${\\cal B}_c=10^{18}$, $2\\times 10^{18}$ and $3\\times10^{18}$ G respectively.} \\label{m-r} \\end{figure} \\begin{figure}[t] \\centering \\includegraphics[angle=-90,width=4.5 in]{m-rr2.eps} \\caption{Same as Fig. 2, except for rotating NSs with rotational frequency $\\Omega=2000$ s$^{-1}$.} \\label{m-rr1} \\end{figure} \\begin{figure}[t] \\centering \\includegraphics[angle=-90,width=4.5 in]{m-rr4.eps} \\caption{Same as Fig. 2, except for rotating NSs with rotational frequency $\\Omega=4000$ s$^{-1}$.} \\label{m-rr2} \\end{figure} \\subsection{Matter instability due to magnetic field} It has been noted that there is an onset of matter instability for certain sets of $\\beta$ and $\\gamma$ (which correspond to different magnetic field profiles) for every ${\\cal B}_c$. Now we explore EoSs for different magnetic field profiles to understand the instability. Figures \\ref{18eos}a and \\ref{318eos}a show the variations of $P$ at ${\\cal B}_c=10^{18}$ and $3 \\times 10^{18}$ G respectively, for different sets of $\\beta$ and $\\gamma$. For each $\\beta$, we take minimum $\\gamma$ to be $1$ for which ${\\cal B}$ varies very slowly with $n_b$ (see Figs. \\ref{18eos}b and \\ref{318eos}b). The maximum value of $\\gamma$ for every $\\beta$ has been chosen in such a way that beyond a certain value of $n_b$, $P$ ceases to increase (and decreases) with the further increase of $n_b$. This implies that the matter becomes unstable after that value of $n_b$. Keeping $\\beta$ constant if $\\gamma$ decreases, then instability disappears at some value of $\\gamma$ and the EoS tends to become that for without magnetic field. The point to note is that if $\\beta$ decreases, the instability occurs at larger values of $\\gamma$ and $n_b$, which is evident from Figs. \\ref{18eos}a and \\ref{318eos}a. The instability occurs because the magnetic pressure contributes negatively to the matter pressure as reflected from the Eqs. (\\ref{enden}) and (\\ref{press}). Since the field strength increases with the increase of density, more negative contribution is added to the pressure, and finally, at a certain density, the pressure ceases to increase with the increase of the density. This effect has not been discussed in any earlier literature to the best of our knowledge. This is consistent with the result of softening the EoS with the increase of magnetic field. Since the softening of the EoS implies that with the energy density the rate of increase of pressure decreases, above a certain value of magnetic field, the rate becomes zero, and subsequently becomes negative. That is why, above a certain value of magnetic field the matter becomes unstable. This constrains the value of central magnetic field of a NS having high surface magnetic field, e.g. magnetar. In this context, it is worth mentioning that the inclusion of AMM effect \\cite {bpl2} stiffens the EoS, which dominates over the effect due to Landau quantization on EoS when field strength is $>5\\times10^{18}$ G. This stiffening is only possible when the field pressure is not considered in obtaining the EoS. However, the inclusion of field pressure further softens the EoS, as discussed above. When the field strength is $\\sim 5\\times 10^{19}$ G, the deviation in EoSs between the cases with and without AMM effect is significant throughout the density range \\cite{bpl2}. However, this deviation is much small compared to that due to the extra effect arised from field pressure. Hence, at such a strong magnetic field, AMM effect is unable to stiffen the EoS. It is, however, seen that when the field strength is near $5\\times10^{18}$ G, AMM effect is insignificant at high density, but significant at low density. However, one should also remember that this result is obtained by considering constant field strength over the whole range of density. Thus, in the realistic case, when the field strength is several orders lower than $10^{18}$ G at low density, i.e., near the surface, AMM effect does not play any role in modifying the EoS. Therefore, we do not consider the AMM effect in our calculation while ${\\cal B}_c\\leq10^{19}$ G. \\begin{figure}[t] \\begin{center} \\includegraphics[angle=-90,width=4.5 in]{18paper.eps} \\caption{Variation of (a) $P$, and (b) $B$ with normalized baryon number density for different magnetic field profiles with ${\\cal B}_c=10^{18}$ G. The solid curve is for without magnetic field. The sets of dashed, dotted, dot-dashed and short-dashed curves are for $\\beta=0.1$, $0.01$, $0.001$ and $0.0001$ respectively. In each set (for each $\\beta$), the upper curve in (a) is for $\\gamma=1$. The values of $\\gamma$ for the lower curve in sets in (a) are given in Table \\ref{gamma}. In (b), for each $\\beta$, the curves which reach ${\\cal B}_c$ at a comparatively lower $n_b$ correspond to maximum $\\gamma$ cases.} \\label{18eos} \\end{center} \\end{figure} \\begin{figure}[t] \\begin{center} \\includegraphics[angle=-90,width=4.5 in]{318paper.eps} \\caption{Same as Fig. \\ref{18eos} except for ${\\cal B}_c=3\\times10^{18}$ G.} \\label{318eos} \\end{center} \\end{figure} \\begin{table}[t] \\begin{center} \\begin{tabular}[h]{|c|c|c|} \\hline $\\beta$ & \\multicolumn{2}{c|}{$\\gamma$} \\\\ \\cline {2-3} & ${\\cal B}_c=10^{18}$ G & ${\\cal B}_c=3 \\times 10^{18}$ G \\\\ \\hline 0.1 & 4.0 & 1.4 \\\\ 0.01 & 6.0 & 2.6 \\\\ 0.001 & 6.0 & 3.5 \\\\ 0.0001 & 7.5 & 4.4 \\\\ \\hline \\end{tabular} \\caption{$\\gamma$-s for different lower curves shown in Figs. \\ref{18eos}a and \\ref{318eos}a for different $\\beta$-s.} \\label{gamma} \\end{center} \\end{table} In Fig. \\ref{m-rallst}, the mass-radius relations for different ${\\cal B}_c$ as well as that without magnetic field have been shown. We notice that for a fixed magnetic field profile, e.g. $\\beta=0.1$ and $\\gamma=1$, the maximum mass decreases as ${\\cal B}_c$ increases from $10^{18}$ G to $4\\times10^{18}$ G. Now as ${\\cal B}_c$ increases further from $4\\times10^{18}$ G, with the same magnetic field profile (same $\\beta$ and $\\gamma$), $P$ starts decreasing with the increase of $n_b$ at a very low $n_b$, leading to unstable matter. For ${\\cal B}_c=5\\times10^{18}$ G, we obtain stable matter EoS for $\\beta=0.01$ and $\\gamma=2$, which deviates maximum from that without magnetic field, as shown in Fig. \\ref{518eos}. It is found that the maximum deviation of mass-radius relation for ${\\cal B}_c=5 \\times 10^{18}$ G with respect to that without magnetic field is less than that deviation for ${\\cal B}_c=4 \\times 10^{18}$ G (see Fig. \\ref{m-rallst}). Now if ${\\cal B}_c$ increases further to $10^{19}$ G, we note that (see Fig. \\ref{19eos}) with the above said profile ($\\beta=0.01$ and $\\gamma=2$) the matter becomes unstable at a very low density. We have stable matter up to a reasonably high density with ${\\cal B}_c=10^{19}$ G for $\\beta=0.001$ and $\\gamma=3$ with maximum deviation in EoS from that without magnetic field. However, Fig. \\ref{m-rallst} shows that in this case the deviation of EoS from that without magnetic field case is even lesser. Hence we can conclude that the effect of magnetic field is significant when ${\\cal B}_c$ is around $4\\times10^{18}$ G depending on the magnetic field profile. Beyond that, the magnetic field profiles needed to create significant effect of magnetic field on the mass-radius relation make the matter unstable. \\begin{figure}[t] \\begin{center} \\includegraphics[angle=-90,width=4.5 in]{m-rallst.eps} \\caption{Mass-radius relations for static star with different values of ${\\cal B}_c$. The solid curve is for without magnetic field. The dashed curves are for ${\\cal B}_c=10^{18}$, $3\\times 10^{18}$ and $4\\times10^{18}$ G respectively from top to bottom for $\\beta =0.1$ and $\\gamma=1$. The dotted curve corresponds to ${\\cal B}_c=5\\times10^{18}$ G for $\\beta=0.01$ and $\\gamma=2$, and the dot-dashed curve is for ${\\cal B}_c=10^{19}$ G when $\\beta=0.001$ and $\\gamma=3$.} \\label{m-rallst} \\end{center} \\end{figure} \\begin{figure}[t] \\begin{center} \\includegraphics[angle=-90,width=4.5 in]{518eos.eps} \\caption{$P$ as a function of normalized baryon number density with ${\\cal B}_c=5\\times10^{18}$ G for different magnetic field profiles. The solid curve is for without magnetic field, the dashed curve for $\\beta=0.1$ and $\\gamma=1$, the dotted curve for $\\beta=0.01$ and $\\gamma=3$, the dot-dashed curve for $\\beta=0.01$ and $\\gamma=2$ and the short-dashed curve for $\\beta=0.001$ and $\\gamma=4$.} \\label{518eos} \\end{center} \\end{figure} \\begin{figure}[t] \\begin{center} \\includegraphics[angle=-90,width=4.5 in]{19eos.eps} \\caption{$P$ as a function of normalized baryon number density with ${\\cal B}_c=10^{19}$ G for different magnetic field profiles. The solid curve is for without magnetic field. The dashed curve is for $\\beta=0.01$ and $\\gamma=2$, the dotted curve for $\\beta=0.01$ and $\\gamma=1$ and the dot-dashed curve for $\\beta =0.001$ and $\\gamma=3$.} \\label{19eos} \\end{center} \\end{figure} As we have mentioned earlier, the effect of magnetic field on quark matter using the MIT bag model has been studied in many earlier work. Therefore, further it will be interesting to see the effect of magnetic field on other models, e.g. of Dey et al. \\cite{d98}, which is in progress." }, "1005/1005.2526_arXiv.txt": { "abstract": "{ The goal of local helioseismology is to elicit three-dimensional information about the sub-surface (or far-side) structure and dynamics of the Sun from observations of the helioseismic wave field at the surface. The physical quantities of interest include flows, sound-speed deviations and magnetic fields. However, strong surface magnetic fields induce large perturbations to the waves making inversions difficult to interpret. The purpose of this paper is to outline the methods of analysis used in local helioseismology, review discoveries associated with the magnetic Sun made using local helioseismology from the past three years, and highlight the efforts towards imaging the interior in the presence of strong magnetic fields. } ", "introduction": "Helio- and astero-seismology provide the means to ``see\" beneath the surface of the Sun or star, and to gain an understanding of the physics of stellar interiors. The analysis of the frequencies of the global modes of oscillation allows us to study the internal stellar structure only as a function of radius and unsigned latitude. With high spatial resolution observations of the velocity signal on the Sun's surface, local helioseismology is able to image the three dimensional subsurface structure of the Sun and image magnetic activity on the far-side. Local helioseismology aims to interpret the full wave field observed at the solar surface, not only the frequencies of the normal modes. We restrict our attention, here, to local helioseismology. The Doppler signal, which is the line-of-sight velocity of the solar surface, is the fundamental observable. Full-disk high-quality data is obtained from instruments including the Michelson Doppler Image (MDI) onboard the Solar and Heliospheric Observatory (SOHO) \\citep{MDI95}, the Global Oscillation Network Group (GONG) \\citep{Harveyetal1996}, and the Helioseismic and Magnetic Imager (HMI) onboard the recently launched Solar Dynamics Observatory (SDO). MDI also runs higher-resolution campaigns on smaller fields-of-view. The Magneto-Optical filter at Two Heights \\citep{Finsterle2004} observes simultaneously at two heights in the atmosphere allowing an analysis of the vertical propagation of waves. Various local helioseismic techniques exist to analyse the oscillations. The main ones are time-distance analysis \\citep{DJHP93}, ring diagram analysis \\citep{Hill1988} and acoustic imaging/holography \\citep{CCL97,LB97}. Ring-diagram analysis is most closely related to global mode helioseismology. By analysing frequency shifts over small regions of the solar surface, the direction and amplitude of subsurface flows can be determined. Time-distance analysis computes the travel time of a wave packet travelling between two points on the solar surface. Helioseismic holography (acoustic imaging) reconstructs the subsurface wave fields by propagating waves either forward or backward in time. Other methods include Fourier-Hankel analysis \\citep{BDL87}, designed specifically to analyse waves surrounding a sunspot. \\cite{Woodard2009} used direct modelling to interpret the wave correlations in wavevector-frequency space. The forward problem is then to compute the expected seismic observations from a particular model of the Sun. The inverse problem is to infer the internal properties of the Sun from the seismic observables. Local helioseismology is the only way to elicit the three-dimensional structure of the solar interior, including structures and flows on a range of scales. In this paper we focus on the local helioseismology of sunspots, active regions and large scale flows varying with the solar cycle. There are particular problems associated with doing helioseismology in the immediate vicinity of strong surface magnetic fields. The forward problem used to infer the subsurface properties has previously been based exclusively on solving hydrodynamic equations in the absence of a magnetic field. Additionally, the actual inversion method assumes that any perturbation to the waves is small, however it has been demonstrated that strong surface magnetic fields cause large perturbations to the waves. Knowing the subsurface structure of sunspots and active regions will constrain sunspot models and hopefully determine their deeper structure. The immediate need now is to quantitatively model wave propagation through magnetic fields and, subsequently, account for the effects when inverting for the subsurface structure. An important question in solar physics is to understand the mechanics of the solar dynamo. Local helioseismology has identified large scale flow variations that appear to be intimately connected to the surface activity. Understanding such connections will help to decipher the larger scale dynamics of the magnetic Sun. Other goals of local helioseismology are to measure the meridional flow and changes at the base of the convection zone, which would help constrain models of the dynamo. This short review will cover the basic techniques used in local helioseismology in Section~\\ref{at}. How the interior structure is subsequently inferred is described in Section~\\ref{inv}. Section~\\ref{obs} will then cover some of the results, over the past three years, from local helioseismic analysis tied to solar activity. A concentrated effort towards modelling waves propagating through magnetic fields to understand the observations will be reviewed in Section~\\ref{mod}. Recent reviews of local helioseismology include \\cite{KHH06,GT07,Birch2008}. For more extensive reviews of local helioseismology see, for example, \\cite{GBS10}. ", "conclusions": "Further confirmation of the surface effects of strong magnetic fields have been found in observations, casting further doubt on over-simplified inversions for wave-speed and flows below magnetic active regions and sunspots. The effects of filtering, source suppression and mode conversion of the waves in strong surface magnetic fields have yet to be fully quantified for each of the analysis techniques. In turn, a concerted effort is underway to quantitatively simulate the seismic signature of the Sun in the presence of magnetic fields to explore these effects. The simulations are still being fully analysed and in the near future a quantitative measure of the effects may be used to infer the correct subsurface properties of a sunspot. Future efforts in local helioseismology will focus on deeper inversions, particularly of the base of the convection zone, and steps towards this have been taken using time-distance techniques \\citep{ZHKM09}. A challenging goal is to detect the emergence of magnetic flux before it is observed at the surface. Some progress has been made towards this \\citep[e.g.][]{KMHH08,KHH09,HKZM10}, but strong definitive signatures have not been observed. Future studies of the effects of magnetic flux emergence will require systematic helioseismic analysis of many emerging active regions and realistic simulations. The HMI instrument onboard the recently launched \\linebreak SDO very soon provide increased resolution observations of the seismic Sun and provide accompanying vector magnetograms on a regular basis. Thanks to the improved spatial resolution we anticipate being able to do local helioseismology closer to the limb (and poles)." }, "1005/1005.5076_arXiv.txt": { "abstract": "{We point out some of the outstanding challenges for embedding inflationary cosmology within string theory studying the process of reheating for models where the inflaton is a closed string mode parameterising the size of an internal cycle of the compactification manifold. A realistic model of inflation must explain the tiny perturbations in the cosmic microwave background radiation and also how to excite the ordinary matter degrees of freedom after inflation, required for the success of Big Bang Nucleosynthesis. We study these issues focusing on two promising inflationary models embedded in LARGE volume type IIB flux compactifications. We show that phenomenological requirements and consistency of the effective field theory treatment imply the presence at low energies of a \\textit{hidden sector} together with a {\\it visible sector}, where the Minimal Supersymmetric Standard Model fields are residing. A detailed calculation of the inflaton coupling to the fields of the hidden sector, visible sector, and moduli sector, reveals that the inflaton fails to excite primarily the visible sector fields, instead hidden sector fields are excited copiously after the end of inflation. This sets severe constraints on hidden sector model building where the most promising scenario emerges as a pure $N=1$ SYM theory, forbidding the kinematical decay of the inflaton to the hidden sector. In this case it is possible to reheat the Universe with the visible degrees of freedom even though in some cases we discover a new tension between TeV scale SUSY and reheating on top of the well-known tension between TeV scale SUSY and inflation.} \\preprint{DESY 10-072} \\begin{document} \\tableofcontents \\bigskip ", "introduction": "Primordial inflation has been a very successful paradigm which can explain the observed temperature anisotropy in the cosmic microwave background (CMB) radiation~\\cite{WMAP7} (for a recent review see~\\cite{RM}). Any successful model of inflation must also explain how to excite the Standard Model (SM) quarks and leptons after the end of inflation through the process of reheating (for a review see~\\cite{Reheat-rev}) required for the success of Big Bang Nucleosynthesis~\\cite{BBN}. In this regard only visible sector models of inflation can be considered to be safe where the inflaton carries the SM charges~\\cite{MSSM-inf}. Such models are primarily based on the Minimal Supersymmetric Standard Model (MSSM), where supersymmetry (SUSY) helps in obtaining a {\\it cosmologically flat} potential via $D$-flat directions (for a review on MSSM flat directions see~\\cite{MSSM-rev}). The advantage of visible sector models embedded within the MSSM is that it can also reproduce the required dark matter abundance necessary for the structure formation~\\cite{ADM}. On the contrary there exists a plethora of models of inflation where the inflaton belongs to the hidden sector since it is a SM gauge singlet. Typically these models can explain the temperature anisotropy in the CMB~\\cite{RM}, but fail to explain how the inflaton energy density gets transferred primarily to visible sector degrees of freedom (\\textit{dof}), and not to hidden sector ones. Note that \\textit{a priori} a gauge singlet inflaton has no preference to either the visible or the hidden sector. Therefore, it is a challenge to construct a model of inflation where the inflaton belongs to the hidden sector and motivate the inflaton couplings to hidden and visible \\textit{dof}. It is typical to encounter such a scenario in string theory which provides many closed string modes that are good inflaton candidates. In particular, the low energy effective action of string compactifications on Calabi-Yau three-folds typically has a large number of uncharged massless scalar fields with a flat potential, called moduli. These moduli can be the ideal candidates to drive inflation. Since their vacuum expectation value (VEV) determines various parameters (like masses and coupling constants) of the four-dimensional effective theory, it is important to stabilise them by providing masses to these moduli. Otherwise, the presence of these massless scalars with effective gravitational coupling would mediate unobserved long range fifth forces. It is well known by now that nonzero background fluxes induce potentials for some of the moduli~\\cite{gvw,drs,gkp}. However these fluxes severely backreact on the Calabi-Yau geometry generically modifying the internal space to a manifold with a more complicated structure. This is not the case for type IIB compactifications where the fluxes induce just a warp factor. For this reason, moduli stabilisation is best understood in the context of type IIB which is therefore well suited for the discussion of phenomenology and cosmology. However the stabilisation of all the geometric moduli requires to take into account various perturbative and non-perturbative effects \\cite{kklt, LVS, GenAnalofLVS}. One of the nice properties of type IIB flux compactifications is that global (or bulk) issues decouple from local (or brane) issues. The former are model independent issues which depend on moduli stabilisation and some of them are: (a)~Different hierarchical scales,~(b)~SUSY breaking,~(c)~Soft SUSY breaking mass terms,~(d)~Cosmological constant, and ~(e)~Inflation. On the other hand, the local issues depend on the particular local $D$-brane construction and some of them are: (a)~ Gauge group,~(b)~Chiral spectrum,~(c)~Yukawa couplings,~(d)~Gauge coupling unification, ~(e)~Proton stability, and ~(f)~Baryogenesis. An excellent example of type IIB compactifications with stabilised moduli is given by the LARGE Volume Scenario (LVS) originally proposed in \\cite{LVS}. This is very appealing for particle physics phenomenology and cosmology. In these compactifications $\\alpha'$ and $g_s$ corrections can be combined with non-perturbative effects in order to generate a potential for all the K\\\"{a}hler moduli \\cite{GenAnalofLVS}, whereas the background fluxes induce a potential for the dilaton and the complex structure moduli. Unlike the KKLT set-up \\cite{kklt}, the moduli stabilisation is performed without fine tuning of the values of the internal fluxes and the Calabi-Yau volume is fixed at an exponentially large value (in string units). As a consequence, one has a very reliable four-dimensional effective description, as well as a tool for the generation of phenomenologically desirable hierarchies. There are two different choices to embed the visible sector within LVS: \\begin{enumerate} \\item {\\it Geometric regime}: the visible sector is built via magnetised intersecting $D7$ branes wrapping a blow-up 4-cycle which is stabilised at a value larger than the string scale~\\cite{GeomRegime}; \\item {\\it Quiver locus}: the visible sector is built via fractional $D$-branes at the singularity obtained by shrinking down a blow-up mode~\\cite{quiver}. \\end{enumerate} Both of them have different phenomenology, the scales and hierarchies are different and we will briefly review them in the following section. One of the interesting features for all these setups is that at low energies they always come with a hidden and a visible sector. Therefore these models will serve as a perfect playground to study the inflaton couplings to both the \\textit{dof}. Furthermore, the two sectors do not have any direct coupling among themselves, except indirectly via the closed string sector. In this respect the two sectors are isolated from each other, nevertheless their total energy density will govern the dynamics of the Universe. It is then crucial to find an answer to each of the following questions: how does the inflaton couple to each sector? how does perturbative reheating take place? is it at all possible to excite predominantly the MSSM \\textit{dof}? Thankfully, these questions can be answered within this setup which involves an interesting interplay between local and global issues. We shall stress that reheating can set severe constraints both on the hidden sector physics and the scale of inflation which should be compatible with CMB predictions, successful reheating, and TeV scale SUSY\\footnote{Reheating and thermalisation have been studied previously in the context of closed string excitations in brane inflation~\\cite{Many-reheat}, in the case of warped geometry~\\cite{Warped-reheat}, and in closed string models of inflation~\\cite{Green-reheat, Preheating}. In all these cases, no clear distinction has been made between hidden and visible \\text{dof}. In this respect such studies are interesting but leave little impact on how the MSSM \\textit{dof} are excited, whose answer is relevant for the success of BBN.}. In order to study these issues, we shall focus on two very promising models of string inflation embedded in LVS, where the inflaton is a closed string mode, more precisely a K\\\"{a}hler modulus parameterising the size of an internal 4-cycle: (1)~{\\it Blow-up inflation} (BI) \\cite{kahlerinfl}: in this case the inflaton is the size of a blow-up mode yielding a small field inflationary model which is in good agreement with current observational data; and (2)~{\\it Fibre inflation} (FI) \\cite{fiberinfl}: in this case the inflaton is the size of a K3 fibre over a $\\mathbb{C}P^1$ base producing a large field inflationary model which can be a potential candidate if primordial gravity waves are discovered. These models represent some of the few known examples where inflation can be achieved without fine-tuning any parameter in the potential. In fact it is the typical no-scale structure of the potential that allows to solve the $\\eta$-problem in a natural way. In spite of the successes of these models we shall discover two outstanding challenges for closed string inflationary scenarios: \\begin{enumerate} \\item Most of the inflaton energy gets dumped to hidden sector \\textit{dof}. This will lead to overproduction of hidden sector dark matter if there exists a stable species. This problem seems to rule out all the inflationary models unless the hidden sector consists just of an $N=1$ pure SYM theory. This severe constraint on hidden sector model building can be relaxed only for the geometric regime case with the inflaton wrapped by the visible sector $D7$-stack. \\item Incompatibility between TeV scale SUSY and reheating on top of the well-know tension between inflation and TeV scale SUSY \\cite{InfvsSUSY}: assuming that the density perturbations can be generated in a non-standard way by a second curvaton or modulating field \\cite{BCGQTZ}, one could set the scale of the scalar potential in order to get inflation, the correct amount of CMB temperature fluctuations and TeV scale SUSY at the same time. However it might still be impossible to achieve an efficient reheating of the visible sector due to the fact that the inflaton might decay after BBN. This problem seems to rule out all the FI models (with the possible exception of FI at the quiver locus) and BI in the geometric regime with the inflaton not wrapped by the visible sector. The only models which are left over are BI in the geometric regime with the inflaton wrapped by the visible sector or BI at the quiver locus. It is important to stress that both these models require fine-tuning to have a successful inflationary scenario. \\end{enumerate} This paper is organised as follows. In section 2, we give a brief review of LVS stressing the different features of the geometric regime and the quiver locus case. We then discuss the dynamics of BI and FI presenting their predictions for the cosmological observables. We finally briefly describe the process of perturbative and non-perturbative reheating and thermalisation. In section 3, we discuss the dynamics of the hidden sector and the constraints on hidden sector \\textit{dof}. In section 4, we describe the inflaton couplings to both hidden and visible \\textit{dof} estimating the maximal reheating temperature in the approximation of sudden thermalisation of the visible sector for BI and FI, respectively. In section 5, we discuss our main results and conclude. In appendix \\ref{appendice}, we present a detailed derivation of the moduli canonical normalisation, mass spectrum, couplings and decay rates to any particle present in our models both for BI and FI. More precisely, we have derived the moduli couplings to fermions/sfermions, Higgs/Higgsinos, gauge bosons/gauginos of the visible and hidden sector, and the moduli self couplings. ", "conclusions": "Once the inflaton has completely decayed into visible sector \\textit{dof}, the system evolves to a complete thermalisation and the thermal bath generates a moduli-dependent finite temperature scalar potential\\footnote{Note that within the MSSM, as it is the case for many scenarios, the final reheating temperature can be as low as $1$ TeV.}. The behaviour of this potential for an arbitrary LVS has been studied in \\cite{LVSatFiniteT} where it was found that the finite temperature corrections do not develop any new minimum, and so they cannot induce any phase transition, but they can still lead the system to a dangerous decompactification limit. The computation of the decompactification temperature for LVS gives \\cite{LVSatFiniteT}: \\be T_{max}\\simeq\\frac{M_P}{\\vo^{3/4}}, \\label{Tmax} \\ee and this is valid for all the different scenarios that we have studied here. This temperature sets the maximal temperature of our Universe, and so one has to make sure that $T_{RH}^{max}10^7$, the inflaton fluctuations are not able anymore to generate the right amount of density perturbations, which could be, on the other hand, produced by some non-standard mechanism, like a curvaton or a modulating field \\cite{BCGQTZ}. \\medskip Let us now summarise the main results we have found in this paper: \\begin{itemize} \\item In any LVS hidden sectors are always present along with the visible sector. \\item The ubiquitous presence of a hidden sector wrapping the inflaton 4-cycle implies that BI needs fine-tuning whereas FI does not. \\item The hidden sector and the visible sector are not directly coupled, and furthermore the inflaton coupling to visible sector \\textit{dof} can never be made stronger than its coupling to hidden sector \\textit{dof}. \\item Two generic problems arise: \\begin{enumerate} \\item Inflaton energy dumping to hidden, instead of visible, sector \\textit{dof}. As such this is not a problem provided the hidden sector \\textit{dof} are relativistic and there exists a curvaton mechanism which is responsible for creating the perturbations and the visible sector matter. Such a scenario was considered in past~\\cite{EKM}, where the MSSM Higgs would dominate the energy density during its oscillations and would create all the visible \\textit{dof}. It would also be interesting to investigate if this problem could be solved by building a curvaton model along the lines of \\cite{BCGQTZ} where the curvaton dominates the energy density. \\item Overproduction of hidden sector dark matter if there exists a massive long lived species, otherwise there would be too many new relativistic \\textit{dof}, which would spoil the BBN predictions. \\end{enumerate} \\item A possible solution could be to render the inflaton decay to hidden sector \\textit{dof} kinematically forbidden. This sets severe constraints on hidden sector model building leaving as the most promising possibility just a pure $N=1$ SYM theory. This is always the case for each hidden sector at the quiver locus and in the geometric regime if the inflaton is not wrapped by the visible sector. On the contrary, in the geometric regime case with the inflaton wrapped by the visible sector, the requirement of having just a pure SYM theory for the hidden sector not wrapping the inflaton 4-cycle can be relaxed. \\item This constraint seems to be in contrast with the generic presence of hidden photons whose study received a lot of attention recently \\cite{Mark}. More precisely, the appearance of hidden photons living on small or large cycles seems to be compatible with reheating only for models in the geometric regime with the inflaton 4-cycle wrapped by the visible sector. All the other cases seem to be incompatible with the existence of hidden photons. \\end{itemize} Assuming that these problems can be solved by an appropriate brane construction of the hidden sector, let us summarise our results for the maximal reheating temperature of the visible sector: \\begin{figure}[ht] \\begin{center} \\begin{tabular}{c||c|c} & Geometric Regime & Quiver Locus \\\\ \\hline\\hline \\\\ & & \\vspace{-0.9cm}\\\\ & Inflaton not wrapped by the visible sector: \\\\ Blow-up & $T_{RH}^{max}\\sim M_P \\vo^{-2}\\sim 10^{6- 8}$ GeV for $\\vo\\sim 10^{6- 7}$ & $T_{RH}^{max}\\sim M_P \\vo^{-5/2}\\sim 10^{2- 4}$ GeV \\\\ Inflation & Inflaton wrapped by the visible sector: & for $\\vo\\sim 10^{6- 7}$ \\\\ & $T_{RH}^{max}\\sim M_P \\vo^{-1}\\sim 10^{12- 13}$ GeV for $\\vo \\sim 10^{6- 7}$ \\\\ \\hline \\\\ & & \\vspace{-0.9cm}\\\\ & Inflaton not wrapped by the visible sector: \\\\ Fibre & $T_{RH}^{max}\\sim M_P \\vo^{-17/6}\\sim 10^{6- 9}$ GeV for $\\vo\\sim 10^{3- 4}$ & $T_{RH}^{max}\\sim M_P \\vo^{-10/3}\\sim 10^{4- 7}$ GeV \\\\ Inflation & Inflaton wrapped by the visible sector: & for $\\vo\\sim 10^{3- 4}$ \\\\ & $T_{RH}^{max}\\sim M_P \\vo^{-5/2}\\sim 10^{8- 10}$ GeV for $\\vo\\sim 10^{3- 4}$ \\end{tabular}\\\\ \\end{center} \\end{figure} We stress that only the model of BI with the visible sector at the quiver locus gives rise to a viable reheating with TeV scale SUSY at the same time. In all other cases $M_{soft}\\gg 1$ TeV. Let us now assume that the inflaton is only responsible to drive inflation but the density fluctuations are generated in a non-standard way by a second curvaton or modulating field \\cite{BCGQTZ}. In this way we still obtain inflation but the volume could be set so to obtain TeV scale SUSY. The resulting maximal reheating temperature in the visible sector would turn out to be: \\begin{enumerate} \\item BI in the geometric regime with the inflaton not wrapped by the visible sector: for $\\vo\\simeq 10^{14}$ (as needed to obtain TeV scale SUSY in the geometric regime), we would obtain $T_{RH}^{max}\\simeq 10$ eV. This is in clear disagreement with the lower bound coming from BBN: $T_{RH}^{max}>1$ MeV. Thus we infer that this model does not seem to be compatible with TeV scale SUSY. \\item BI in the geometric regime with the inflaton wrapped by the visible sector: for $\\vo\\simeq 10^{14}$, we would obtain $T_{RH}^{max}\\simeq 5\\cdot 10^{5}$ GeV which is not in contrast with any observation. Thus we infer that this model could be made compatible with TeV scale SUSY. \\item FI in the geometric regime with the inflaton not wrapped by the visible sector: for $\\vo\\simeq 10^{14}$, and $c\\simeq 5\\cdot 10^{-2}$ we would obtain $T_{RH}^{max}\\simeq 10^{-14}$ eV, which is unrealistically small. Thus we infer that this model does not seem to be compatible with TeV scale SUSY. \\item FI in the geometric regime with the inflaton wrapped by the visible sector: for $\\vo\\simeq 10^{14}$, and $c\\simeq 5\\cdot 10^{-2}$ we would obtain $T_{RH}^{max}\\simeq 10^{-8}$ eV, which is unrealistically small. Thus we infer that this model does not seem to be compatible with TeV scale SUSY. On top of this, we would obtain an unrealistically small value for the visible sector gauge coupling: $\\alpha^{-1}_{vis}=\\langle\\tau_1\\rangle=c\\,\\vo^{2/3}\\simeq 10^8$. \\item FI at the quiver locus : for $\\vo\\simeq 10^6$ (as needed to obtain TeV scale SUSY at the quiver locus), we would obtain $T_{RH}^{max}\\simeq 1$ MeV which is of the same order of magnitude as $T_{BBN}$. Thus we infer that this model could be made compatible with TeV scale SUSY only if $T_{RH}^{max}$ turns out to be actually a bit higher than $T_{BBN}$ and the matter-antimatter asymmetry is generated in a non-thermal way. \\end{enumerate} Thus, even though FI does not require any fine-tuning of the coefficient of the string loop corrections, it seems very difficult to obtain a viable reheating for these models which is also compatible with TeV scale SUSY. On the contrary, both BI at the quiver locus and in the geometric regime with the inflaton wrapped by the visible sector, seems to be more promising to achieve inflation, reheating and TeV scale SUSY at the same time. We finally stress that our estimate of the reheating temperature was based on the approximation of sudden thermalisation of the visible sector \\textit{dof}. In reality it is more likely that the system slowly evolves to thermal equilibrium with an actual reheating temperature $T_{RH}< T_{RH}^{max}$. However we shall leave the detailed study of the thermalisation process for future investigation." }, "1005/1005.4593_arXiv.txt": { "abstract": "We show that we can obtain a good fit to the present day stellar mass functions (MFs) of a large sample of young and old Galactic clusters in the range $0.1 - 10$\\,\\Msolar with a tapered power law distribution function with an exponential truncation of the form $dN/dm \\propto m^\\alpha \\, [1 - e^{-(m/m_c)^\\beta}]$. The average value of the power-law index $\\alpha$ is $\\sim -2$, that of $\\beta$ is $\\sim 2.5$, whereas the characteristic mass $m_c$ is in the range $0.1 - 0.8$\\,M$_\\odot$ and does not seem to vary in any systematic way with the present cluster parameters such as metal abundance, total cluster mass or central concentration. However, $m_c$ shows a remarkable correlation with the dynamical age of the cluster, namely $m_c/M_\\odot \\simeq 0.15 + 0.5 \\times \\tau_{\\rm dyn}^{3/4}$, where $\\tau_{\\rm dyn}$ is the dynamical age taken as the ratio of cluster age and dissolution time. { The small scatter seen around this correlation is consistent with the uncertainties on the estimated value of $\\tau_{\\rm dyn}$.} We attribute the observed trend to the onset of mass segregation via two-body relaxation in a tidal environment, causing the preferential loss of low-mass stars from the cluster and hence a drift of the characteristic mass $m_c$ towards higher values. { If dynamical evolution is indeed at the origin of the observed trend, it would seem plausible that high-concentration globular clusters, now with median $m_c \\simeq 0.33$\\,M$_\\odot$, were born with a stellar MF very similar to that measured today in the youngest Galactic clusters and with a value of $m_c \\simeq 0.15$\\,M$_\\odot$. This hypothesis is consistent with the absence of a turn-over in the MF of the Galactic bulge down to the observational limit at $\\sim 0.2$\\,M$_\\odot$ and, if correct, it would carry the implication that the characteristic mass is not set by the thermal Jeans mass of the cloud.} ", "introduction": "There is general consensus that stars do not form in isolation but rather from the fragmentation of molecular clouds that leads to star clusters (e.g. Lada \\& Lada 2003; Elmegreen 2010). This makes stellar clusters ideal places to study the properties of star formation and its outcome, the stellar initial mass function (IMF). However, it is also equally well established today that a large majority of stars, in our Galaxy and elsewhere, are not in clusters but in the field, since clusters disrupt over time (see e.g. Gieles 2010). Therefore, any attempt to set constraints on the star formation mechanisms from the analysis of the present day stellar mass function (MF) of but the youngest clusters cannot ignore the consequences of their dynamical evolution. This issue becomes particularly important if we want to compare the results of star formation in physically different environments and with various ages. The obvious example is addressing differences in the way low-mass stars ($< 1$\\,M$_\\odot$) formed in globular clusters (GCs), at $z \\simeq 5$ or about 12\\,Gyr ago, and in young clusters (YCs) in the local universe. While in both cases the raw data show a broad plateau in the mass distribution, suggesting a characteristic mass of order a few tenths of M$_\\odot$ (e.g. Elmegreen et al. 2008), until the effects of dynamical evolution are properly understood, no meaningful conclusion can be drawn either on the uniformity of the star formation process over time or on the possible role of the environment. In order to cover a wide range of metallicities and initial densities one is forced to compare stellar systems that formed a Hubble time apart, such as the YCs and GCs mentioned above. Therefore, the universality of the IMF over time and location remains as yet an unresolved issue. On the other hand, even though the two-body relaxation process that governs the cluster's dynamical evolution and that leads to preferential loss of low-mass stars is today rather well understood (e.g. Spitzer 1987), it is in general not possible to roll back the effects of dynamics and derive the IMF from the present day MF, since we cannot trace back the trajectories of stars that have escaped the cluster. It is however possible and statistically meaningful to study the evolution of the stellar MF on a global scale by looking at the differences between clusters at different evolutionary stages. This requires a homogeneous sample of high quality observations of Galactic (open and globular) clusters, treated in a uniform way and with reliable errors. Early in this decade, this type of homogeneous study became possible for GCs (Paresce \\& De Marchi 2000), mostly thanks to the Hubble Space Telescope. In the meanwhile, high quality data have become available for YCs as well, mostly from wide field ground based surveys, thereby making this study possible on a global scale. \\begin{center} \\begin{deluxetable*}{clllllrllcclc}[t] \\tablecaption{The sample of stellar clusters used in this study \\label{tab1}} \\tablehead{ \\colhead{ID} & \\colhead{Name} & \\colhead{$\\alpha$} & \\colhead{$\\beta$} & \\colhead{$m_c$} & \\colhead{$m$ range} & \\colhead{$[Fe/H]$} & \\colhead{$\\log M_{\\rm tot}$} & \\colhead{$c$} & \\colhead{$\\log t$} & \\colhead{$\\log t_{\\rm dis}$} & \\colhead{$\\tau_{\\rm dyn}$} & \\colhead{Ref} \\\\ \\colhead{} & \\colhead{} & \\colhead{} & \\colhead{} & \\colhead{[M$_\\odot$]} & \\colhead{[M$_\\odot$]} & \\colhead{} & \\colhead{[M$_\\odot$]} & \\colhead{} & \\colhead{[yr]} & \\colhead{[yr]} & \\colhead{} & \\colhead{} } \\startdata 1 & $\\rho$\\,Oph & $-1.9 \\pm 0.2$ & $1.9 \\pm 0.2$ & $0.17 \\pm 0.03$ & $0.02$ -- $7$ & $0.08$ & $3.3$ & & $ 5.7$ & $8.7$&$0.001$ & a \\\\ 2 & Orion N. C. & $-1.9 \\pm 0.2$ & $2.0 \\pm 0.3$ & $0.19 \\pm 0.05$ & $0.02$ -- $50$ & $-0.01$ & $1.8$ & $0.3$ & $ 6.0$ & $8.7$ &$0.002$ & b, c \\\\ 3 & Taurus & $-2.3 \\pm 0.2$ & $1.8 \\pm 0.2$ & $0.51 \\pm 0.05$ & $0.03$ -- $3.5$& $-0.03$ & $1.7$ & & $ 6.0$ & $6.3$ & $0.5$ & d \\\\ 4 & IC\\,348 & $-1.9 \\pm 0.1$ & $3.3 \\pm 0.2$ & $0.12 \\pm 0.05$ & $0.03$ -- $2.2$& $-0.01$ & $1.8$ & $0.3$ & $ 6.3$ & $8.7$ & $0.004$ & e \\\\ 5 & $\\sigma$\\,Ori & $-2.0 \\pm 0.1$ & $3.2 \\pm 0.2$ & $0.14 \\pm 0.02$ & $0.02$ -- $2.4$& $-0.01$ & $2.3$ & $0.6$ & $ 6.5$ & $8.7$ & $0.006$ & f \\\\ 6 & $\\lambda$\\,Ori& $-1.9 \\pm 0.1$ & $2.3 \\pm 0.2$ & $0.12 \\pm 0.02$ & $0.04$ -- $2.7$& $-0.01$ & $2.8$ & $0.6$ & $ 6.7$ & $8.7$ & $0.01$ & g \\\\ 7 & Cha\\,I & $-1.9 \\pm 0.2$ & $2.3 \\pm 0.3$ & $0.15 \\pm 0.03$ & $0.01$ -- $3.5$& $-0.11$ & $2.6$ & & $ 6.7$ & $8.7$ &$0.01$ & h \\\\ 8 & IC\\,2391 & $-2.1 \\pm 0.1$ & $2.4 \\pm 0.3$ & $0.14 \\pm 0.02$ & $0.05$ -- $0.5$& $-0.09$ & $2.2$ & $0.6$ & $ 7.7$ & $8.8$ & $0.06$ & i \\\\ 9 & Blanco\\,1 & $-1.7 \\pm 0.1$ & $1.8 \\pm 0.2$ & $0.22 \\pm 0.03$ & $0.05$ -- $1.75$& $0.14$ & $3.5$ & $0.6$ & $ 8.1$ & $9.1$ & $0.09$ & j \\\\ 10 & Pleiades & $-2.2 \\pm 0.1$ & $2.3 \\pm 0.2$ & $0.27 \\pm 0.03$ & $0.04$ -- $9$& $0.03$ & $2.9$ & $0.6$ & $ 8.1$ & $9.3$ & $0.06$ & k \\\\ 11 & M\\,35 & $-1.7 \\pm 0.1$ & $2.4 \\pm 0.2$ & $0.30 \\pm 0.03$ & $0.09$ -- $1.4$& $-0.21$ & $3.2$ & $0.7$ & $ 8.2$ & $9.4$ & $0.08$ & i \\\\ 12 & Coma Ber & $-1.3 \\pm 0.4$ & $1.5 \\pm 0.3$ & $0.45 \\pm 0.10$ & $0.12$ -- $1.1$& $-0.08$ & $2.4$ & $0.4$ & $ 8.6$ & $8.9$ & $0.48$ & b \\\\ 13 & Praesepe & $-2.0 \\pm 0.2$ & $4.0 \\pm 0.5$ & $0.20 \\pm 0.02$ & $0.07$ -- $1.5$& $0.00$ & $3.3$ & $0.7$ & $ 8.8$ & $9.5$ & $0.18$ & l, m\\\\ 14 & Hyades & $-2.1 \\pm 0.1$ & $2.8 \\pm 0.2$ & $0.45 \\pm 0.05$ & $0.07$ -- $2.7$& $0.14$ & $2.6$ & $0.4$ & $ 8.8$ & $9.2$ & $0.37$ & n \\\\[0.15cm] 15 & NGC\\,104 & $-2.0$ & $1.8 \\pm 0.15$ & $0.33 \\pm 0.02$ & $0.11$ -- $0.7$ & $-0.76$ & $5.84$ & $2.03$ & $10.11$ & $10.93$& $0.15$ & p \\\\ 16 & NGC\\,5139 & $-2.0$ & $2.3 \\pm 0.2$ & $0.33 \\pm 0.02$ & $0.13$ -- $0.7$ & $-1.62$ & $6.17$ & $1.61$ & $10.06$ & $10.75$ & $0.21$ & p \\\\ 17 & NGC\\,5272 & $-2.0$ & $2.4 \\pm 0.2$ & $0.33 \\pm 0.02$ & $0.16$ -- $0.7$ & $-1.57$ & $5.65$ & $1.84$ & $10.06$ & $10.92$ & $0.14$ & p \\\\ 18 & NGC\\,6121 & $-2.0$ & $2.8 \\pm 0.25$ & $0.35 \\pm 0.02$ & $0.09$ -- $0.5$ & $-1.20$ & $4.83$ & $1.59$ & $10.10$ & $10.16$ & $0.88$ & p \\\\ 19 & NGC\\,6254 & $-2.0$ & $1.9 \\pm 0.2$ & $0.33 \\pm 0.02$ & $0.13$ -- $0.6$ & $-1.52$ & $5.00$ & $1.40$ & $10.06$ & $10.39$ & $0.47$ & p \\\\ 20 & NGC\\,6341 & $-2.0$ & $2.0 \\pm 0.2$ & $0.30 \\pm 0.02$ & $0.11$ -- $0.7$ & $-2.28$ & $5.30$ & $1.81$ & $10.12$ & $10.40$ & $0.52$ & p \\\\ 21 & NGC\\,6397 & $-2.0$ & $2.3 \\pm 0.2$ & $0.33 \\pm 0.02$ & $0.09$ -- $0.55$ & $-1.95$ & $4.65$ & $2.50$ & $10.10$ & $10.27$ & $0.67$ & p \\\\ 22 & NGC\\,6656 & $-2.0$ & $2.3 \\pm 0.3$ & $0.31 \\pm 0.03$ & $0.10$ -- $0.6$ & $-1.64$ & $5.46$ & $1.31$ & $10.10$ & $10.62$ & $0.31$ & p \\\\ 23 & NGC\\,6752 & $-2.0$ & $2.6 \\pm 0.1$ & $0.34 \\pm 0.02$ & $0.10$ -- $0.6$ & $-1.56$ & $5.15$ & $2.50$ & $10.07$ & $10.50$ & $0.37$ & p \\\\ 24 & NGC\\,6809 & $-2.0$ & $2.3 \\pm 0.3$ & $0.30 \\pm 0.02$ & $0.11$ -- $0.6$ & $-1.81$ & $5.04$ & $0.76$ & $10.09$ & $10.32$ & $0.58$ & p \\\\ 25 & NGC\\,7078 & $-2.0$ & $2.7 \\pm 0.3$ & $0.23 \\pm 0.02$ & $0.15$ -- $0.7$ & $-2.26$ & $5.75$ & $2.50$ & $10.07$ & $10.93$ & $0.14$ & p \\\\ 26 & NGC\\,7099 & $-2.0$ & $1.6 \\pm 0.2$ & $0.25 \\pm 0.02$ & $0.15$ -- $0.7$ & $-2.12$ & $5.00$ & $2.50$ & $10.11$ & $10.39$ & $0.53$ & p \\\\[0.15cm] 27 & NGC\\,2298 & $-2.0$ & $3.0 \\pm 0.2$ & $0.54 \\pm 0.01$ & $0.20$ -- $0.8$ & $-1.85$ & $4.50$ & $1.28$ & $10.02$ & $10.25$ & $0.59$ & q \\\\ 28 & NGC\\,6218 & $-2.0$ & $3.2 \\pm 0.4$ & $0.62 \\pm 0.03$ & $0.30$ -- $0.8$ & $-1.48$ & $4.94$ & $1.39$ & $10.10$ & $10.33$ & $0.60$ & q \\\\ 29 & NGC\\,6712 & $-2.0$ & $3.2 \\pm 0.2$ & $0.80 \\pm 0.03$ & $0.30$ -- $0.8$ & $-1.01$ & $4.97$ & $0.90$ & $10.02$ & $10.22$ & $0.63$ & q \\\\ 30 & NGC\\,6838 & $-2.0$ & $2.6 \\pm 0.2$ & $0.60 \\pm 0.02$ & $0.30$ -- $0.8$ & $-0.73$ & $4.23$ & $1.15$ & $10.14$ & $10.21$ & $0.84$ & q\\\\ \\enddata \\tablecomments{Table columns are as follows: ID number; cluster name; best fitting value of $\\alpha$ and $1\\,\\sigma$ uncertainty; best fitting value of $\\beta$ and $1\\,\\sigma$ uncertainty; best fitting value of $m_c$ and $1\\,\\sigma$ uncertainty; mass range over which the TPL fit is performed; cluster metallicity; total cluster mass; central concentration, or the logarithmic ratio of tidal radius and core radius; cluster age; estimated time to dissolution; dynamical time, or the ratio of the cluster age and the time to dissolution; bibliographic reference for the MF data, as follows. For YCs: (a) \\cite{luh99}; (b) \\cite{kra07}; (c) \\cite{sle04}; (d) \\cite{sal10}; (e) \\cite{mue03}; (f) \\cite{cab08}; (g) \\cite{bar04a}; (h) \\cite{luh07}; (i) \\cite{bar04b}; (j) \\cite{mor07}; (k) \\cite{mor03}; (l) \\cite{cha05}; (m) \\cite{wil95}; (n) \\cite{bou08}. For dense GCs: (p) \\cite{pdm00}. For loose GCs: (q) \\cite{dem07}. For YCs, the metallicity and age are from the bibliography in column Ref. and references therein, whereas total mass and concentration are from Piskunov et al. (2008), { although the value of M$_{\\rm tot}$ is from K\\\"uc\\\"uk \\& Akkaya (2010) for Taurus, from Hodapp et al. (2009) for $\\sigma$\\,Ori and from Dolan \\& Mathieu (2002) for $\\lambda$\\,Ori.} For GCs, the metallicity, total mass and concentration are from the 2003 revision of the GC catalogue of Harris (1996), whereas the age is from Forbes \\& Bridges (2010). The values of $t_{\\rm dis}$ have been determined as explained in Section\\,5, whereas $\\tau_{\\rm dyn}$ is the dynamical age of each cluster, expressed as $t/t_{\\rm dis}$.} \\end{deluxetable*} \\end{center} ", "conclusions": "At least qualitatively, the trend seen in Figure\\,\\ref{fig2} is fully consistent with the onset of mass segregation via two-body relaxation in a tidal environment, causing the preferential loss of low-mass stars from the cluster and hence a drift of the characteristic mass towards higher values. Vesperini \\& Heggie (1997) showed from N-body models that stellar evaporation, integrated over the cluster's orbit and further enhanced by the presence of the Galactic tidal field, causes a flattening of the MF, i.e. a selective depletion at the low-mass end. Their models do not reveal a drift of the characteristic mass towards higher values, but their IMF in all cases is a pure power-law. A drift of $m_c$ is visible in the N-body models of Portegies Zwart et al. (2001), although it proceeds very slowly and only appears when the study of the MF is limited to the regions inside the half-mass radius. {That mass segregation is present in both young and globular clusters has long been established observationally. Examples include many of the YCs in our sample, such as M\\,35 (e.g. Mathieu 1983; Barrado y Navascu\\'es et al. 2001), Pleiades (e.g. Converse \\& Stahler 2008), Hyades (Reid 1993; Perryman et al. 1998), Blanco\\,1 (Moreaux et al. 2007). As for GCs, mass segregation has been observed in all those objects where MF measurements exist at various radial distances (see e.g. De Marchi et al. 2007 and references therein) and can be satisfactorily explained as being the result of the two-body relaxation process (Spitzer 1987). Furthermore, De Marchi et al. (2007) have also shown that dense and loose GCs have today systematically different GMFs and this is consistent with a much stronger selective loss of low-mass stars in the latter. Loose GCs have systematically larger $m_c$ than denser clusters with otherwise similar metallicity or space motion parameters. This strongly suggests that the value of $m_c$ must drift as a result of the cluster's dynamical evolution, which is quicker in loose systems, even though the analysis of the N-body models (e.g. Portegies Zwart et al. 2001) has so far failed to detect this effect.} In fact, Kruijssen (2009) has recently developed a simple physical model for the evolution of the MF that builds on the pioneering work of H\\'enon (1969). His results (see Figure\\,14 in Kruijssen 2009) show a shift in the inflection point of the MF with evolutionary time, consistent with a drift in $m_c$ (D. Kruijssen, priv. comm.). { A potentially very interesting consequence of the trend shown in Figure\\,\\ref{fig2}, if dynamical evolution is indeed at its origin, is that it would seem plausible that all GCs, having now $m_c \\simeq 0.33$\\,M$_\\odot$ if they are dense or $\\sim 0.65$\\,M$_\\odot$ if they are loose, were born with a stellar MF like the one measured today in the youngest Galactic clusters, or more precisely with $m_c \\simeq 0.15$\\,M$_\\odot$. Although not necessarily required by Figure\\,\\ref{fig2}, this scenario is consistent with it. The careful reader will notice that this statement appears at odds with the conclusions of a previous work by our team. From a study of the GMFs of twelve halo GCs with widely different dynamical histories and noting that they show a very similar characteristic mass ($m_c=0.33 \\pm 0.03$\\,\\Msolar), Paresce \\& De Marchi (2000) concluded that this value of $m_c$ had to be a feature of the stellar IMF in GCs. However, it is now apparent that this conclusion was the result of a selection effect, due to the fact that all twelve GCs in the Paresce \\& De Marchi (2000) sample have high central concentration, since no reliable information on the low-mass end of the MF of loose GCs was available at that time. The analysis presented by De Marchi et al. (2007) suggests that dense GCs have lost less low-mass stars than looser clusters throughout their life, but this does not mean that they have lost none. Therefore, one can no longer conclude that the $m_c \\simeq 0.33$\\,M$_\\odot$ value observed in the present-day GMF of dense GCs reflects the characteristic mass of their IMF, since it could have been smaller. } The possibility that GCs were born with a much smaller $m_c$ value than what we measure today in their MF { would also be consistent} with the absence of a turn-over in the MF of the Galactic bulge down to the observational limit at $\\sim 0.2$\\,M$_\\odot$ (Zoccali et al. 2000). Regardless as to whether the bulge is primarily the result of fast formation at early epochs (e.g. Ballero et al. 2007) or a collection of the stars lost from disrupted clusters throughout the life of the Galaxy (e.g. Gnedin \\& Ostriker 1997), its very location at the bottom of the Galaxy's potential well makes it very hard for low-mass stars to escape from it. Therefore, while the present MF of the Galactic disc could represent a complex average over time and space, that of the bulge should have not been altered by dynamical evolution and, unlike the case of the comparably old GCs, should still reflect the properties of the IMF. Forthcoming HST observations of the Galactic bulge (see e.g. Brown et al. 2009) will provide insights on its MF below $0.15$\\,M$_\\odot$, where we expect a turn-over. {To be sure, N-body models of dense stellar systems able to deal with a large number of particles ($N > 10^6$) and incorporating a realistic treatment of stellar evolution and of the interaction with the Galactic tidal field (see Portegies Zwart et al. 2010) would be necessary to simulate the growth of the characteristic mass with time in GCs. Our analysis (see Section\\,5) suggests that young and globular clusters may share rather similar values of the $\\alpha$ and $\\beta$ parameters. Thus, if the picture sketched above is correct and, by strongly affecting the value of $m_c$, dynamical evolution act as the main source of variations in the present-day stellar MF of Galactic clusters, one might postulate that Population I and II stars should have similar IMFs. While not necessarily required by our analysis, this hypothesis is fully compatible with our results. Claims of the IMF universality that one finds in the literature (e.g. Gilmore 2001; Elmegreen et al. 2008) are based on the observation that stellar MFs in different environments have {\\em similar} $m_c$ values, within a factor of a few. Here we show that it is plausible that the original value of $m_c$ be actually {\\em the same} for Population I and II stars. If this is true, one would have to consider the possibility that the characteristic mass in the IMF is not necessarily set by the Jeans mass scale in the forming cloud (e.g. Larson 1998; Larson 2005) and that the physical conditions of the environment do not play a significant role in the outcome of the star formation process, i.e. on the IMF. This, however, would not necessarily mean that the star formation process itself is the same: recent numerical simulations proposed by Goodwin \\& Kouwenhoven (2009) and Bate (2009) suggest that the IMF shape is largely insensitive to parameters such as the binary fraction and mass ratio in binaries. If this is the case, the role of the IMF as an effective tool to probe the physics of star formation may have to be critically reconsidered.}" }, "1005/1005.4070_arXiv.txt": { "abstract": "{ We present evidence for the contribution of high-mass globular clusters to the stellar halo of the Galaxy. Using SDSS-II/SEGUE spectra of over 1900 G- and K-type halo giants, we identify for the first time a subset of stars with CN bandstrengths significantly larger, and CH bandstrengths lower, than the majority of halo field stars, at fixed temperature and metallicity. Since CN bandstrength inhomogeneity and the usual attendant abundance variations are presently understood as a result of star formation in globular clusters, we interpret this subset of halo giants as a result of globular cluster dissolution into the Galactic halo. We find that $2.5\\%$ of our sample is CN-strong, and can infer based on recent models of globular cluster evolution that the fraction of halo field stars initially formed within globular clusters may be as large as $50\\%$.} ", "introduction": "Hierarchical structure formation is presently the dominant explanation for galaxy formation, based on observed fluctuations in the cosmic microwave background \\citep{WMAP} and sophisticated numerical simulations of their evolution to the present day (e.g., Diemand et al. 2007\\nocite{DKM07}; Springel et al. 2008\\nocite{SWV08}). In this picture, galaxies like the Milky Way are formed through the coalescence of multiple low-mass galaxies which develop within a much larger dark matter halo. The initial disagreement between the calculated mass function of dark matter subhaloes in the simulations and the observed mass function of nearby dwarf galaxies (e.g., Klypin et al. 1999\\nocite{KKV99}; Moore et al. 1999\\nocite{MGG99}) is being addressed from several directions at the same time, from complex semianalytic simulations that include star formation and feedback processes, and calculate the chemodynamical evolution of Milky Way-like galaxies (e.g., Johnston et al. 2008\\nocite{JB08}; Tumlinson 2010\\nocite{T10} and references therein) to searches for extremely low-mass galaxies in the Local Group (e.g., Zucker et al. 2006a, 2006b\\nocite{ZB06a}\\nocite{ZB06b}; Belokurov et al. 2007\\nocite{BZE07}). The stellar halo of the Milky Way is thought to have been constructed mostly through the early ($\\simeq 10$ Gyr ago) accretion of low-mass protogalaxies. The halo exhibits considerable substructure in density and in velocity (e.g. Bell et al. 2008\\nocite{BZB08}), and kinematically distinct streams presently observed in the halo (e.g., Majewski et al. 2003\\nocite{MSW03}; Duffau et al. 2006\\nocite{DZV06}; Mart{\\'{\\i}}nez-Delgado et al. 2007\\nocite{MD07}) are interpreted as remnants of more recent or ongoing merger activity. The ``ECHOS'' identified in \\citet{SRB09} are interpreted as older substructure that has lost some spatial coherence with time. Studies of the abundance distributions and star formation histories in nearby dwarf galaxies (e.g., Koch et al. 2007a, 2007b, 2008a, 2008b\\nocite{KG07}\\nocite{KW07}\\nocite{KG08}\\nocite{KM08}; Kirby et al. 2008\\nocite{KSG08}; Aoki et al. 2009\\nocite{AAS09}; Frebel et al. 2009\\nocite{FKS09}, 2010\\nocite{FSGW10}) have shown that there is a reasonable concordance between the properties of the present-day Milky Way halo (as characterized by, e.g., Schoerck et al. 2009\\nocite{SCC09}) and the dwarf galaxies that would have been available as stellar contributors early in Galactic history (e.g., Font et al. 2006\\nocite{FJB06}; Carollo et al. 2007\\nocite{CB07}). However, the ongoing dissolution of globular clusters such as Palomar 5 (e.g., Odenkirchen et al. 2001, 2002, 2003\\nocite{OGR01}\\nocite{OGD02}\\nocite{OGD03}; Rockosi et al. 2002\\nocite{ROG02}; Grillmair \\& Dionatos 2006\\nocite{GD06}) and NGC 5466 \\citep{BEI06} implies that some fraction of halo stars are initially formed in globular clusters. This dissolution is driven by internal 2-body relaxation, stellar evolution processes and tidal interactions with the Galactic potential, and can cause significant mass loss over the lifetime of typical halo globular clusters (e.g., Gnedin \\& Ostriker 1997\\nocite{GO97}). One particular model of globular cluster formation, described in \\citet{DVD08}, posits that all old halo clusters surviving to the present day have lost at least 90\\% of their initial mass. Early in the development of the cluster, winds from AGB stars collected in the cluster center and formed a second generation of low-mass stars. Type Ia supernovae then caused the cluster to expand, and stars at large radii, mostly members of the first generation, were lost. In the model, the end result, $\\simeq 10$ Gyr later, is a $10^{5} - 10^{6} M_{\\sun}$ cluster with two stellar populations that differ slightly in age and abundance pattern. This two-generation model was developed specifically to explain the anomalous light-element abundance patterns observed in globular cluster stars, specifically the presence in every old globular cluster in the Milky Way of a subpopulation with typical Population II abundances, and a second subgroup with the same metallicity but enhanced N, Na, and Mg along with depleted C, O, and Al. This abundance bimodality has been studied extensively in globular clusters (Langer et al. 1992\\nocite{LSK92}, Kraft 1994\\nocite{K94}, and Gratton et al. 2004\\nocite{GSC04} all provide thorough reviews of the topic), and explanations for the second abundance subgroup have varied from pollution \\citep{CC98} to internal mixing \\citep{L85} to enrichment of star-forming gas by moderate- to high-mass stars (e.g., Cottrell \\& DaCosta 1981\\nocite{CD81}; Yong et al. 2008\\nocite{Y08}) or high-mass binaries \\citep{DMP09}. Surface pollution of already-formed stars would result in abundance anomalies that are erased at first dredge-up, while current models of deep mixing (e.g., Charbonnel \\& Zahn 2007\\nocite{CZ07}) indicate that it only begins to operate after first dredge-up. Both of these processes have been effectively ruled out as explanations for abundance bimodality by the presence of abundance variations at all evolutionary phases in globular clusters (e.g., Briley et al. 2002\\nocite{BCS02}; Harbeck et al. 2003\\nocite{HSG03a}). In the primordial enrichment scenario, there is ongoing discussion over the exact source of enriching material. In some models the source is moderately high-mass ($\\sim 4-5M_{\\sun}$) AGB stars (e.g., Parmentier et al. 1999\\nocite{PJM99}), while \\citet{DCM07} claim that rotating high-mass ($M \\ga 10 M_{\\sun}$) stars are a better source for processed material because of their very short lifetimes and \\citet{DMP09} prefer high-mass binaries because of their potentially strong mass loss and low wind velocities. Stars with these light-element abundance anomalies are readily identified through strong UV/blue CN molecular absorption and relatively weak absorption in the CH G band, and are hereafter called ``CN-strong stars'', with the understanding that the full abundance pattern from C through Al necessarily follows the CN variation. They are not observed to exist in open clusters (e.g., Smith \\& Norris 1984\\nocite{SN84}; Jacobson et al. 2008\\nocite{JFP08}; Martell \\& Smith 2009\\nocite{MS09}) or the halo field \\citep{GSC04}. This feedback process apparently only occurs in the high-density environment of globular clusters, and as a result the characteristic sawtooth abundance pattern from carbon through aluminium can be used as a marker of globular cluster origin. Given the contributions globular clusters are presently making to the halo field, and the significant mass loss predicted theoretically over the lifetime of the Galactic globular cluster system (e.g., Baumgardt et al. 2008\\nocite{BKP08}), it is intriguing that no CN-strong stars have to date been observed in the halo. We interpret this as a qualitative sign that the contributions to the halo field of globular clusters as we know them today are relatively minor, as is also suggested in \\citet{Y08}. To test this interpretation, we searched for CN-strong halo giants in the Sloan Extension for Galactic Understanding and Exploration (SEGUE) survey \\citep{Y09}. The SEGUE survey is a spectroscopic extension of imaging taken during the Sloan Digital Sky Survey \\citep{SDSS00}, with targets selected to address questions of halo substructure and Galactic formation history. Data were taken from 2005 August through 2008 July using a 640-fiber multiobject spectrograph and the same telescope at Apache Point Observatory that was used for SDSS imaging. The first portion of the SEGUE data was made publically available in 2008 as part of SDSS Data Release 7 (DR7), including roughly 240,000 spectra in 200 ``pencil beam'' lines of sight containing stars chosen for specific purposes (i.e., fields in the Sagittarius stream, M dwarfs to study extremely local kinematic substructure, G and K giants to study the distant halo). In addition to flux-calibrated spectra, DR7 also offers the products of the ``SEGUE Stellar Parameters Pipeline'' (SSPP), which include derived stellar parameters like effective temperature, [Fe/H] metallicity, and radial velocity, determined automatically through template matching, $\\chi^{2}$ minimization, cross-correlation or grid-matching methods, as appropriate. Lee et al. (2008a\\nocite{L08a}; 2008b\\nocite{L08b}) and \\citet{AP08} give thorough explanations of the SSPP pipeline and process. ", "conclusions": "Although light-element abundance variations have not been observed before in the halo, our identification of these CN-strong halo field stars is not wholly unexpected. There are several well-understood mechanisms for globular cluster mass loss, and theoretical studies of globular cluster formation and evolution (e.g., D'Ercole et al. 2008\\nocite{DVD08}; Baumgardt et al. 2008\\nocite{BKP08}) predict significant mass loss in individual clusters as well as a dramatic reshaping of the cluster mass function with time. The data set we analyzed, selected from the SEGUE survey, is not representative of the full halo, in mass, evolutionary phase, or metallicity. Distances to the candidate CN-strong stars range from 4 to nearly 40 kpc, with $93\\%$ found within 20 kpc of the Sun. However, our result for red giants is generalizable to all halo stars, since abundance bimodality is observed to exist at all masses and evolutionary phases in globular clusters. More fundamentally, all of the cuts we made in selecting the final data set are blind to CN and CH bandstrengths and light-element abundances, and all nitrogen-enhanced giants with metallicities above [Fe/H]$=-1.8$ ought to show clearly strong CN bands. Since approximately $2.5\\%$ of our halo red giants exhibit strong CN bands and weak CH bands, we expect that the same fraction of the entire halo will contain the same abundance enhancements and depletions. This prediction can be confirmed by observations of dwarfs in the halo field, because main sequence stars in globular clusters show the same abundance division as giants (e.g., Briley et al. 2002\\nocite{BCS02}). \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{f9.eps}} \\caption[Generalized histograms, all metallicities]{ Generalized histograms of $\\delta S(3839)$ for CN-weak and CN-strong stars in each [Fe/H] bin. As in Fig. \\ref{ff6}, maximum metallicity for each panel is given in the upper left corner. As in Fig. \\ref{ff5}, CN-strong histograms were amplified for clarity. The multiplicative factor for each panel is given in the upper right corner. } \\label{ff9} \\end{figure} In order to convert $f^{\\mathrm{p}}_{\\mathrm{h}}$, the present-day fraction of CN-strong halo stars, into $f^{\\mathrm{gc}}_{\\mathrm{h}}$, the fraction of globular cluster-originating stars in the halo field, we must consider what fraction CN-strong stars comprise of the stars originally formed in globular clusters. In the two-generation scenario of \\citet{DVD08}, roughly $90\\%$ of stars originally formed in a globular cluster, consisting entirely of first-generation stars (with halo-like chemistry), are lost between the epoch of cluster formation and the present day. This means that $f^{\\mathrm{m}}_{\\mathrm{gc}}$, the fraction of stars that remain as members of the globular cluster they were formed in, is around $0.1$. Since $f^\\mathrm{p}_\\mathrm{gc}$, the fraction of present-day globular clusters stars that are CN-strong, is around $0.5$ (e.g., Kraft 1994\\nocite{K94}), we can calculate that $f^{\\mathrm{gc}}_{\\mathrm{h}}=\\frac{f^{\\mathrm{p}}_{\\mathrm{h}}}{f^{\\mathrm{m}}_{\\mathrm{gc}}\\times f^{\\mathrm{p}}_{\\mathrm{gc}}}$. Since $f^{\\mathrm{p}}_{\\mathrm{h}}=0.025$ in the present study, this suggests that $f^{\\mathrm{gc}}_{\\mathrm{h}}=\\frac{0.025}{0.1\\times 0.5}=0.5$, and that a remarkable $50\\%$ of the halo field originally formed in the massive star clusters that were progenitors of the present-day globular cluster population, with a further unknown contribution of CN-weak stars made by globular clusters that did not survive to the present day and were not massive enough to self-enrich. While some numerical studies of galaxy formation (e.g., Boley et al. 2009\\nocite{BLR09}) have suggested that the halo could be constructed entirely from disrupted globular clusters, there is not presently a strong consensus on the role of globular clusters in cosmological-scale galaxy formation. Precise numerical study of the dynamical evolution of globular clusters is very complicated: the number of particles is large enough, and the relevant timescales short enough, that highly accurate simulations are very time-consuming. However, the development of semianalytic prescriptions for the mass evolution of globular clusters would allow single-halo-scale simulations like those of \\citet{JB08} to include them as a source of halo stars, and to predict what fraction of the halo field ought to originate in globular clusters." }, "1005/1005.4888_arXiv.txt": { "abstract": "The daily variation of the solar photocenter over some 11 years is derived from the Mount Wilson data reprocessed by Ulrich et al. 2010 to closely match the surface distribution of solar irradiance. The standard deviations of astrometric jitter are 0.52 $\\mu$AU and 0.39 $\\mu$AU in the equatorial and the axial dimensions, respectively. The overall dispersion is strongly correlated with the solar cycle, reaching $0.91~\\mu$AU at the maximum activity in 2000. The largest short-term deviations from the running average (up to 2.6 $\\mu$AU) occur when a group of large spots happen to lie on one side with respect to the center of the disk. The amplitude spectrum of the photocenter variations never exceeds $0.033$ $\\mu$AU for the range of periods $0.6$--$1.4$ yr, corresponding to the orbital periods of planets in the habitable zone. Astrometric detection of Earth-like planets around stars as quiet as the Sun is not affected by star spot noise, but the prospects for more active stars may be limited to giant planets. ", "introduction": "\\label{firstpage} The prospects of finding habitable planets orbiting nearby solar-type stars are to a large degree associated with the ultra-precise astrometric instruments under development or construction, such as the SIM Observatory \\citep{sha,unw,catpa} and Gaia \\citep{cas}. The Earth orbiting the Sun produces an observable astrometric wobble of 3 $\\mu$AU (micro-AU) and a radial velocity variation of 0.089 \\ms, if seen equator-on (inclination $i=90\\degr$). A pole-on configuration ($i=0\\degr$ or $180\\degr$) is optimal for astrometry, because the reflex motion signal is present in both dimensions of the sky projection, whereas the radial velocity amplitude, which is proportional to $\\sin i$, drops to zero. For a high signal-to-noise (S/N) ratio detection, the total error budget of the prospective exoplanet detection techniques should be well below these values. The astrophysical jitter caused by the rotation of magnetic features on the surface (spots, faculae) potentially becomes a significant part of the total observational error in this domain of accuracy, even for fairly common inactive dwarfs. The recently completed double blind test of planet-detection capabilities with the SIM Observatory produced encouraging results even for complex multiple systems \\citep{tra}, but magnetic jitter was not taken into account in those simulations. Semi-analytical considerations in \\citep{mak09} supported by Monte-Carlo simulations and indirect observational evidence determined that the astrometric method is much less sensitive to the effects of magnetic features, at least by an order of magnitude for solar-type stars, than the Doppler shift technique. Similar conclusions were drawn by \\citet{cat} for the Sun, based on a sophisticated model of sunspot activity and extensive numerical simulation. The variable distribution of surface brightness is the cause of both the astrometric jitter and the total flux variation. Therefore, the former can be estimated from the latter, given a model describing the number of spots (or bright features), their size and lifetimes. The dispersion of the total solar irradiance (TSI) observed for a few decades with a number of satellites \\citep{fro} sets the standard of quiet stars ($4\\cdot 10^{-4}$, or 400 ppm in normalized flux). The light curves from Kepler for some 120000 stars revealed that at least half of all solar-type dwarfs in the field are at least as quiet as the Sun \\citep{bas}, displaying lower levels of photometric dispersion. By extrapolation, about half of nearby dwarfs should be as amenable to exoplanet detection as the Sun. The aim of this paper is to determine the solar astrometric jitter directly from observations. The results can be used to verify and correct the existing models of spot activity for other stars. ", "conclusions": "We determined that the solar jitter caused by magnetic features is small and should not preclude astrometrists from detecting habitable planets as small as the Earth. A few additional considerations should be made when extrapolating this result to other stars or to actual astrometric observations. \\begin{figure}[htbp] \\plotone{SunHZdX_power.eps} \\caption{Amplitude spectrum of solar astrometric jitter in the equatorial dimension for the range of periods corresponding to the habitable zone.} \\label{pow.fig} \\end{figure} The ultra-precise photometric measurements with Kepler confirmed that about half of all solar-type dwarfs are as quiet as the Sun, or more \\citep{bas}. The stars that are more active than the Sun have much larger filling factors, and their magnetic features may be long-lived (up to a few weeks). Both photometric and astrometric variability is proportional to the average total area $A_s$ occupied by spots \\citep{mak09}. In the model currently adopted for planet detection simulations with SIM (to be published elsewhere), the total area is related to the index of chromospheric activity $\\log R'_{\\rm HK}$ through \\eb \\log A_s=12.468+1.75\\log R'_{\\rm HK}, \\ee which yields 6137 micro-stellar-hemispheres (MSH) for the Sun. As roughly half of nearby field stars are more chromospherically active than the Sun \\citep{gra}, they are expected to have larger spot groups and, therefore, jitter amplitudes. At $\\log R'_{\\rm HK}=-4.6$, the expected area and the jitter are 5 times the solar values, and at $\\log R'_{\\rm HK}=-4.2$, 25 times. This model gives fairly accurate predictions for the active star $\\kappa^1$ Ceti, carefully studied with MOST \\citep{wal}. Furthermore, an additional factor of 2--6 increase may ensue from the limited duration of an astrometric mission. Both SIM and Gaia will nominally operate for 5 years. This is considerably shorter than the solar cycle. If a solar analog is observed at the height of its activity, the level of magnetic jitter may be higher than the multi-year average. Thus, magnetic jitter can indeed pose the natural limit to the sensitivity of the astrometric method at stars significantly more active than the Sun. The amplitude spectrum periodogram shown in Fig.~\\ref{pow.fig} is derived from a high-frequency observational cadence of 2881 data points over 11 years. We expect only about 200 2D measurements to be taken for each target star with SIM Lite during its 5-year nominal mission. If the astrometric jitter is completely random and uncorrelated, the variance of the amplitude of a random harmonic in the spectrum is inversely proportional to the number of data points. In that case, the amplitude spectrum will rise by a factor of $\\sqrt{2881/200}=3.8$ for SIM. On the other hand, if the magnetic activity underlying the astrometric jitter is systematic and deterministic, the corresponding parts of the spectrum will stay as low is in Fig.~\\ref{pow.fig}. Such systematic variations may occur, for example, if sunspots and plages are not randomly distributed on the solar surface, but tend to appear in confined areas called ``active longitudes\" in the literature. Currently, there seems to be no strong evidence for the existence of such long-term structures. A possible north-south asymmetry in the distribution of magnetic features evolving with the solar cycle is another interesting route of investigation where the reconstructed TSI images can be used. The results derived in this paper apply to solar-type stars seen equator-on (inclination $\\simeq 90\\degr$). A pole-on configuration is optimal for astrometric detection, because both dimensions are equally engaged, but it is statistically less probable. Even at moderate inclinations to the line of sight, the distribution of magnetic perturbations in $y$-coordinate may be different from that we find for the Sun, if the magnetic features are confined to the equatorial zone. The features would normally be seen closer to the side of the limb opposite to the visible pole, producing a skewed distribution of $\\Delta y$ depending on whether the spots or bright areas are the dominating contributors. However, giant near-polar spots appear to be common on very active, fast-rotating stars." }, "1005/1005.1180_arXiv.txt": { "abstract": "We use the UKIDSS Ultra-Deep Survey to trace the evolution of galaxy clustering to $z=3$. Using photometric redshifts derived from data covering the wavelength range $0.3 - 4.5 {\\rm \\mu m}$ we examine this clustering as a function of absolute K-band luminosity, colour and star-formation rate. Comparing the deprojected clustering amplitudes, we find that red galaxies are more strongly clustered than blue galaxies out to at least $z=1.5$, irrespective of rest-frame K-band luminosity. We then construct passive and star-forming samples based on stellar age, colour and star-formation histories calculated from the best fitting templates. The clustering strength of star-forming galaxies declines steadily from $r_0\\simeq7h^{-1}$Mpc at $z\\simeq2$ to $r_0\\simeq3h^{-1}$Mpc at $z\\simeq0$, while passive galaxies have clustering strengths up to a factor of two higher. Within the passive and star-forming subsamples, however, we find very little dependence of galaxy clustering on K-band luminosity. Galaxy `passivity' appears to be the strongest indicator of clustering strength. We compare these clustering measurements with those predicted for dark matter halos and conclude that passive galaxies typically reside in halos of mass $M\\ge10^{13}M_{\\odot}$ while luminous star-forming galaxies occupy halos an order of magnitude less massive over the range $0.5>$1000) and magnetic fields of the order of $\\mu$Gauss in the intracluster volume. The physical mechanisms responsible for the origin of this non-thermal intracluster component are matter of debate (e.g. \\cite{1977ApJ...212....1J,Govoni04}), as well as the effects of intracluster cosmic rays (CRs) and magnetic fields on the thermodynamical evolution and mass estimate of galaxy clusters (e.g. \\cite{2009arXiv0909.0270S,2010A&A...510A..76L}). A deep understanding of the evolutionary physics of {\\em all} the different cluster components (dark matter, galaxies, thermal and non-thermal intracluster medium -- ICM) and of their mutual interactions is indeed essential for high-precision cosmology with galaxy clusters \\cite{2005bmri.conf...77A}. In the following, I will give an overview of our current knowledge of the non-thermal component of galaxy clusters. I will also stress the importance of a new generation of multi-wavelength telescopes -- such as the {\\it Low Frequency Array} ({\\it LOFAR}), and the Gamma- and hard X-ray (HXR) satellites {\\it Fermi} and {\\it NuSTAR} -- for a deep understanding of the non-thermal cluster physics. The $\\Lambda$CDM model with H$_0$=70 km ${\\rm s}^{-1} {\\rm Mpc}^{-1}$, $\\Omega_m=0.3$ and $\\Omega_{\\Lambda}=0.7$ has been adopted. ", "conclusions": "" }, "1005/1005.4907_arXiv.txt": { "abstract": "We investigate the role of host galaxy classification and black hole mass ($M_{BH}$) in a heterogeneous sample of 276 mostly nearby ($z<0.1$) X-ray and IR selected AGN. Around 90$\\%$ of Seyfert 1 AGN in bulge-dominated host galaxies (without disk contamination) span a very narrow range in the observed 12$\\micron$ to 2-10keV luminosity ratio ($199\\%$ confidence. Using ring morphology of the host galaxy as a proxy for lack of tidal interaction, we find that AGN luminosity in host galaxies within 70Mpc is independent of host galaxy interaction for $\\sim$ Gyrs, suggesting that the timescale of AGN activity due to secular evolution is much shorter than that due to tidal interactions. We find that LINER hosts have lower 12$\\micron$ luminosity than the median $12\\micron$ luminosity of normal disk- and bulge-dominated galaxies which may represent observational evidence for past epochs of feedback that supressed star formation in LINER host galaxies. We propose that nuclear ULXs may account for the X-ray emission from LINER 2s without flat-spectrum, compact radio cores. We confirmed the robustness of our results in X-rays by comparing them with the 14-195keV 22-month BAT survey of AGN, which is all-sky and unbiased by photoelectric absorption. ", "introduction": "\\label{sec:intro} Galactic nuclei generally considered to be 'active' have bolometric luminosities of $\\sim 10^{-2}$ to $\\sim 10^{4}$ times their host galaxy luminosity and are powered by accretion onto a supermassive black hole. The fundamental parameters that determine accretion onto the central black hole should be: (1) the black hole mass, (2) the amount of gas and dust in the galactic nucleus, (3) a mechanism to drive material in the galactic nucleus onto the black hole and (4) outflows as a result of accretion, which may affect (2) and (3) as feedback. While most galactic nuclei in the local Universe are believed to host very large mass black holes, only $\\sim 0.01-1\\%$ of these nuclei are highly luminous \\citep[e.g.][]{b49,b13}. By contrast, low luminosity nuclei are far more common, accounting for $\\sim 1/3$ of all galactic nuclei \\citep{b95}. Therefore some combination of the parameters determining accretion conspires to keep very luminous galactic nuclei relatively rare, while also providing a very wide range of observed activity at lower luminosities. Black hole mass is the simplest parameter involved in accretion. The greatest activity (or highest bolometric luminosity) in a galactic nucleus occurs when a black hole accretes mass quickly. Black hole mass correlates with several properties of the central regions of the host galaxy, including bulge stellar velocity dispersion ($\\sigma_{\\ast}$), bulge mass and bulge luminosity (see e.g. \\citet{b39} and references therein). This implies that the largest mass black holes live in galaxies with large bulges. If most galactic bulges grew early in the Universe \\citep[e.g.][]{b41}, then black holes grew fastest at that time and so activity in galactic nuclei may have been greatest at high redshift. However, periods of intense activity in a galactic nucleus could also arise at lower redshift, due to mergers between gas-rich galaxies \\citep[e.g.][]{b43} or when a large reservoir of cold gas builds up in the galactic disk \\citep{b24,b40}. The raw material for accretion onto the black hole is gas and dust in the galactic nucleus. How much gas and dust there is in the galactic nucleus depends on the local rate of formation of massive stars \\citep{b25} and on mechanisms driving gas and dust into the nucleus from elsewhere in the galaxy. Mechanisms driving material into the galactic nucleus from the outside could be internal or external to the galaxy. Internal mechanisms involve bars \\citep[e.g.][]{b33} or more generally, internal (secular) disk driven evolution \\citep{b24}; external mechanisms involve tidal disruptions or mergers \\citep[e.g.][]{b41} or nuclear bombardment \\citep{b60}. Not all of the material in the galactic nucleus needs to come from the rest of the galaxy. Material that gains angular momentum close to the accreting supermassive black hole can flow outwards to be recycled in the surrounding galactic nucleus \\citep[e.g.][]{b27,b1,b44,b8}. Indeed if the feedback from the accreting black hole is powerful enough it could disrupt star formation in the galactic nucleus and beyond \\citep{b1}. Outflows from the accreting black hole into the surrounding galactic nucleus could end up terminating inflows onto the black-hole itself, which leads to a picture of galactic nucleus activity as self-regulating \\citep[e.g.][]{b38}. Isolating the fundamental parameters that determine the mass accretion rate onto supermassive black holes is a major observational problem. Many active galactic nuclei (AGN) are shrouded by obscuring material \\citep{b2}, although the obscuration can be complicated \\citep[e.g.][]{b4,b3}. AGN are mostly distant enough that broad-band observations include host galaxy luminosity contributions (e.g. from hot diffuse gas, X-ray binaries and ultra-luminous X-ray sources in the X-ray band alone). One approach to solving this difficult observational problem is to compare broadband luminosities in AGN with fundamental accretion parameters, such as black hole mass, or simple observables such as host galaxy classification, which may be related to fundamental accretion parameters. In this work, we investigate the connection between black hole mass, host galaxy classification and the observed IR and X-ray luminosities of a heterogeneous sample of 276 AGN (mostly from \\citep{b99}). In section~\\ref{sec:sample} we investigate the connection between AGN luminosity, black hole mass and host galaxy classification. In section~\\ref{sec:hockey} we discuss the AGN luminosity distribution and in section~\\ref{sec:group2} we discuss the highly heterogeneous, non-Seyfert 1 AGN in our sample. In section~\\ref{sec:rings}, we discuss the importance of ringed morphology in host galaxies and the implications for AGN activity. We discuss issues of bias and completeness in our sample as well as the reliability of our conclusions in section~\\ref{sec:bias} and in section~\\ref{sec:conclusions} we summarize our conclusions. ", "conclusions": "\\label{sec:conclusions} We investigated the role of black hole mass ($M_{BH}$) and host galaxy morphology in a survey of the observed 2-10keV X-ray and 12$\\micron$ IR luminosities of a heterogeneous sample of 276 mostly nearby AGN. We find that:\\\\ (a) As black hole mass increases, the average observed IR and X-ray luminosity of the homogeneous Group 1 (Seyfert 1.X) AGN increases, maintaining a ratio in the range $R_{IR/X}=$[1,30] over $\\sim 3-4$ orders of magnitude in mass at a confidence level of $>90\\%$. The luminosities for Group 1 AGN remain in the range $\\sim 10^{-3}-10^{-1}$ of Eddington for each decade in mass. By contrast, among the heterogeneous, lower luminosity Group 2 AGN, the ratio $R_{IR/X}$ \\emph{decreases} as black hole mass increases. This is mostly due to a combination of host galaxy contamination and photoelectric absorption. (b) There is an average increase in IR luminosity of a factor of $\\sim 3$ among Group 1 AGN in disk-dominated hosts versus bulge-dominated hosts at a confidence level $>99\\%$. Presumably the increase is due to contamination from the unresolved galactic disk \\citep{b12}. Therefore a better measure of the underlying central engine is the range of $R_{IR/X}$ in bulge-dominated Group 1 AGN, which is $1 L_c$) to allow it to rise to the top of the convection zone. We have used our model to explore the critical length $L_c$ separating flux tubes which form stable magnetic loops in the convection zone ($L < L_c$) from those which erupt through the photosphere ($L > L_c$). Figure \\ref{figure:yosfigure2} shows the dependence of $L_c$ on the temperature defect parameter $\\eta$ for two values of the field strength at the base of the convection zone $B_0$, $10^4{\\rm \\ G}$, and $10^6{\\rm \\ G}$. We find that for perturbations on length scales $L > L_c(B_0,\\eta)$ the rise time is short compared to the duration of the solar cycle. Under the assumption that the solar cycle can be modeled as a kinematic ($e.g.$, $\\alpha \\omega$ or $\\alpha^2 \\omega$) dynamo model operating below the convection zone, we conclude that such perturbations must therefore be rare or nonexistent. We speculate that the magnetic flux which does erupt through the photosphere forms initially from perturbations with $L < L_c$, resulting in stable structures. These are subsequently destabilized either by thermal diffusion or by stretching of the anchor points until $L$ exceeds $L_c$. In either case, we expect that the anchor point separation $L$ should fall roughly within the range $L_c(\\eta = 0)$ to $L_c(\\eta = 1)$. The rise time of flux tubes is the time scale for conduction of heat into the tube in the first case, and the stretching time in the second case. Finally, we argue that active regions are formed from the emergence of a single flux tube segment." }, "1005/1005.0245_arXiv.txt": { "abstract": "{The stellar helium-to-metal enrichment ratio, $\\Delta Y/ \\Delta Z$, is a widely studied astrophysical quantity. However, its value is still not precisely constrained.} {This paper is focused on the study of the main sources of uncertainty which affect the $\\Delta Y/ \\Delta Z$ ratio derived from the analysis of the low-main sequence (MS) stars in the solar neighborhood.} {The possibility to infer the value of the helium-to-metal enrichment ratio from the study of low-MS stars relies on the dependence of the stellar luminosity and effective temperature on the initial helium and metal abundances. The $\\Delta Y/\\Delta Z$ ratio is obtained by comparing the magnitude difference between the observed stars and a reference theoretical zero age main sequence (ZAMS) with the related theoretical magnitude differences computed from a new set of stellar models with up-to-date input physics and a fine grid of chemical compositions. A Monte Carlo approach has been used to evaluate the impact on the result of different sources of uncertainty, i.e. observational errors, evolutionary effects, systematic uncertainties of the models. As a check of the procedure, the method has been applied to a different data set, namely the low-MS of the Hyades.} {Once a set of ZAMS and atmosphere models have been chosen, we found that the inferred value of $\\Delta Y/ \\Delta Z$ is sensitive to the age of the stellar sample, even if we restricted the data set to low luminosity stars. The lack of an accurate age estimate of low mass field stars leads to an underestimate of the inferred $\\Delta Y/ \\Delta Z$ of $\\sim 2$ units. On the contrary the method firmly recovers the $\\Delta Y/ \\Delta Z$ value for not evolved samples of stars such as the Hyades low-MS. Adopting a solar calibrated mixing-length parameter and the PHOENIX GAIA v2.6.1 atmospheric models, we found $\\Delta Y/\\Delta Z~= 5.3~\\pm~1.4$ once the age correction has been applied. The Hyades sample provided a perfectly consistent value. % } {We have demonstrated that the assumption that low-mass stars in the solar neighborhood can be considered as unevolved, does not necessarily hold, and it may indeed lead to a bias in the inferred $\\Delta Y/\\Delta Z$. The effect of the still poorly constrained efficiency of the superadiabatic convection and of different atmosphere models adopted to transform luminosities and effective temperature into colors and magnitudes have been discussed, too.} ", "introduction": "It is a well known and firm result of stellar evolution studies that the main structural, observational and evolutionary characteristics of a star of given mass depend sensitively on the original chemical composition, i.e. the initial helium and metal abundances, $Y$ and $Z$, respectively. As a consequence, these parameters affect also the observable quantities of stellar systems, from star clusters to galaxies. While the present $Z$ in a stellar atmosphere can be obtained by the direct spectroscopic measurements of some tracer element, mainly iron, with the additional assumption on the mixture of heavy elements, the situation of $Y$ is completely different. In the vast majority of stars the helium lines can not be observed, with the exception of those hotter than $20\\,000$ K. This means that for the low-mass stars, which are the most common and long-lasting objects in the Universe, helium can be directly observed only in advanced evolutionary phases, as the blue part of the horizontal branch or in the post-asymptotic giant branch. Thus, the actually measured helium abundance is not the original one, but the result of several complex processes, like dredge-up of nuclearly processed material, diffusion and radiative levitation, which severely alter the surface chemical composition. As a consequence, in order to evaluate the original $Y$, the only possibility is to rely on indirect methods. This explains why such an important parameter is still poorly constrained. As early suggested by \\citet{peimbert74} from the analysis of the chemical composition of the HII regions in the Large Magellanic Cloud, a common approach in both stellar and population synthesis models is to assume a linear relationship between the original $Y$ and $Z$, \\begin{equation} \\label{eq:dydzlin} Y = Y_\\mathrm{P} + \\frac{\\Delta Y}{\\Delta Z} \\times Z \\quad \\end{equation} where $Y_\\mathrm{P}$ is the primordial helium content, i.e. the result of the Big Bang nucleosynthesis, and $\\Delta Y / \\Delta Z$ the ratio which provides the stellar nucleosynthesis enrichment. In the last four decades there has been a continuous effort to try to constrain both the $Y_\\mathrm{P}$ \\citep[see e.g.][]{peebles66,churchwell74,peimbert74,peimbert76,lequeux79,kunth83,pagel86,kunth86,pagel92,mathews93, izotov94,olive95,izotov97,olive97,peimbert02,izotov07,peimbert07,spergel07,dunkley09} and the ratio $\\Delta Y / \\Delta Z$ \\citep{faulkner67,perrin77,lequeux79,peimbert80,peimbert86,pagel92,renzini94, fernandes96,pagel98,jimenez03,izotov04,fukugita06,casagrande07}. Indeed, as previously mentioned, the relationship between $Y$ and $Z$ adopted in stellar models directly affects some important quantities of stellar systems, both resolved and not, inferred by comparing observations and theoretical predictions. Thus, a precise determination of $Y_\\mathrm{P}$ and $\\Delta Y / \\Delta Z$ is of paramount importance for the studies not only of stellar evolution, but also of galaxy evolution. Furthermore, an accurate estimate of $Y_\\mathrm{P}$ is of great cosmological interest, as it constrains the early evolution of the Universe, when the Big Bang nucleosynthesis occurred. In this paper we will focus on the value of the $\\Delta Y / \\Delta Z$ ratio. In the past, several techniques have been used to determine such a ratio by means of HII regions, both galactic and extragalactic, planetary nebulae (PNe), the Sun and chemical evolution models of the Galaxy (see Sect. \\ref{sec:othermeth} for a brief summary of these results). An alternative and well established way to determine the value of the helium-to-metallicity enrichment ratio takes advantage of the dependence of the location of stars in the Hertzsprung-Russell (HR) diagram on their helium content. From the study of stellar populations in the galactic bulge \\citet{renzini94} inferred $2 \\leq \\Delta Y / \\Delta Z \\leq 3$. A frequently adopted approach relies on the analysis of the fine structure of the low-MS of the local field stars in the HR diagram. Pioneers of such an approach have been Faulkner, who found $\\Delta Y / \\Delta Z = 3.5$ \\citep{faulkner67} and Perrin and collaborators, who obtained $\\Delta Y / \\Delta Z = 5$ \\citep{perrin77}. Following these early studies, \\citet{fernandes96} constrained the value of $\\Delta Y / \\Delta Z$ to be larger than 2, by comparing the broadening in the HR diagram of the low-mass MS stars in the solar neighborhood and the theoretical ZAMS of several $Y$ and $Z$. With a similar approach but taking advantage of Hipparcos data, \\citet{pagel98} inferred $\\Delta Y / \\Delta Z = 3 \\pm 2 $. This kind of approach culminated recently in the works by \\citet{jimenez03} and \\citet{casagrande07} who provided, respectively, $\\Delta Y / \\Delta Z = 2.1 \\pm 0.4 $ and 2.1 $\\pm$ 0.9. The present analysis deals with the determination of the $\\Delta Y / \\Delta Z$ ratio by means of the comparison between the local K dwarf stars, for which accurate measures of both the [Fe/H] and parallaxes are available, and state-of-the-art stellar models. A great effort has been devoted to discuss the effect of the main uncertainties still present in stellar models on the inferred value of $\\Delta Y / \\Delta Z$. In Sect. \\ref{sec:models} we present the set of low-mass stellar models we have computed for this paper; in Sect. \\ref{sec:dataset} we describe the data set we have used; Section \\ref{sec:method} contains the description of the analysis method, while Sect. \\ref{sec:uncert} deals with the possible sources of uncertainty that could affect the method itself. Results for the adopted data set are presented in Sect. \\ref{sec:results}. In Sect. \\ref{sec:nonlin} we investigate the possibility of a non linear relation between $Y$ and $Z$. In Sect. \\ref{sec:Hyatest} we apply our method to an independent and unevolved set of stars, i.e. the Hyades low-main sequence. We compare our results to those obtained by other authors, with independent methods, in Sect. \\ref{sec:othermeth}. Section \\ref{sec:concl} contains the final discussion and summary of the whole paper. ", "conclusions": "\\label{sec:concl} The principal aim of this work was to test the reliability of the determination of $\\Delta Y/ \\Delta Z$ by the comparison between low-MS stars and theoretical ZAMS models. A very fine grid of stellar models has been computed for many values of $\\Delta Y/ \\Delta Z$, [Fe/H] and masses adopting two different mixing-length parameters $\\alpha$, namely 1.97 (our solar calibration) and 2.4, and two different sets of atmosphere models to transform luminosities and effective temperatures into magnitudes and color indices, the PHOENIX GAIA v2.6.1 models \\citep{brott} and the ATLAS9 ones \\citep{castelli03}, respectively. A detailed analysis of the capabilities of the method and of the main uncertainty sources affecting the derived results has been performed by means of many numerical experiments on synthetic data set produced under controlled conditions and with precisely known properties. One of the main findings of the paper is that the inferred value of $\\Delta Y/ \\Delta Z$ is quite sensitive to the age of the stellar sample, even in the case in which only very faint (i.e. $M_V > 6$ mag) MS stars are selected. By means of numerical experiment we showed that the lack of an age estimate of low mass field stars leads to an underestimate of the inferred $\\Delta Y/ \\Delta Z$ of about $2$ units. As a consequence, the face value of the helium-to-metals enrichment ratio provided by the recovery procedure applied to the solar neighborhood stellar sample must be corrected for this age-bias. Adopting our reference set of models (i.e. those with our ``solar'' calibrated $\\alpha$) transformed into the observational plane using the PHOENIX GAIA v2.6.1 model atmospheres, we found $\\Delta Y/\\Delta Z = 5.3 \\pm 1.4$. Such a result has been checked against an independent and very accurate data set, that is, the low-MS of the Hyades cluster. This sample has the additional advantage to be unaffected by the age-bias, since its very young age (i.e. $500 \\div 600$ Myr) guarantees that the low-mass ($<0.9 M_{\\sun}$) MS stars which constitutes the sample are essentially unevolved. The recovery method provided again $\\Delta Y/\\Delta Z = 4.75 \\pm 0.35$, in perfect agreement with the result for field stars. To further check the consistency of this final result, we calculated isochrones with $\\Delta Y/ \\Delta Z = 5$ (the closest in our grid of models) and the measured [Fe/H] of Hyades. The good agreement between these isochrones and the Hyades MS, not only in the faint part belonging to the sample but also at higher luminosities, is a further proof of the internal consistency of the recovery procedure. The effect of a change in the assumed efficiency of the superadiabatic convection (i.e. the $\\alpha$ parameter) in the stellar models used to build the ZAMS and in the adopted atmosphere models used to transform luminosities and effectives temperatures into magnitudes and color indices has been discussed, too. More in detail, the recovery method yield a nominal value of $\\Delta Y/\\Delta Z= 1.59 \\pm 1.01$ when adopting the set of theoretical models computed with $\\alpha=~2.4$ and transformed into the observational plane by means of PHOENIX atmosphere models and $\\Delta Y/\\Delta Z= 0.5$ when adopting our standard set of models with $\\alpha=~1.97$ but the ATLAS9 atmosphere models. These values become about 3.6 and 2.5, respectively, once corrected for the age bias. Our data have [Fe/H] determinations coming from two different sources, i.e. \\cite{Nordstrom04} and \\cite{Taylor05} catalogs. The 4 stars we have in common from the two catalogs show a disagreement in the [Fe/H] scale between the two catalogs. This disagreement was already identified in \\cite{Taylor05} and the author provides a Table to account for it and put the two catalogs on the same [Fe/H] scale. Nevertheless, even after the re-zeroing procedure suggested by \\cite{Taylor05}, we still find some disagreement in the final results on $\\Delta Y/ \\Delta Z$ when the two catalogs are considered separately. Probably an even more accurate study of the zero points of metallicity determinations is needed." }, "1005/1005.2847_arXiv.txt": { "abstract": "Terzan~5 is a globular cluster-like stellar system in the Galactic Bulge which has been recently found to harbor two stellar populations with different iron content and probably different ages \\citep{fe09}. This discovery suggests that Terzan~5 may be the relic of a primordial building block which contributed to the formation of the Galactic Bulge. Here we present a re-determination of the structural parameters (center of gravity, density and surface brightness profiles, total luminosity and mass) of Terzan~5, as obtained from the combination of high-resolution ({\\it ESO}-MAD and {\\it HST} ACS-WFC) and wide-field ({\\it ESO}-WFI) observations. We find that Terzan~5 is significantly less concentrated and more massive than previously thought. Still it has the largest collision rate of any stellar aggregate in the Galaxy. We discuss the impact of these findings on the exceptional population of millisecond pulsars harbored in this stellar system. ", "introduction": "Terzan~5 is commonly catalogued as a globular cluster (GC) located in the inner Bulge of our Galaxy. It is difficult to observe because it is heavily reddened, with an average color excess $E(B-V)=2.38$ (Barbuy et al. 1998; Valenti et al. 2007). Not only is the reddening large, but it strongly depends on the line of sight (differential reddening; see Ortolani et al. 1996, Valenti et al. 2007). Terzan~5 has an exceptionally large population of millisecond pulsars (MSPs). Indeed, the 33 MSPs detected so-far in Terzan~5 amount to about $25\\%$ of the entire sample of known MSPs in Galactic GCs (Ransom et al. 2005; see the updated list at {\\tt http://www.naic.edu/pfreire/GCpsr.html}) As part of a project (Ferraro et al. 2001, 2003; Cocozza et al. 2008) aimed at studying the properties of stellar populations harboring MSPs, we obtained a set of high-resolution images of Terzan~5 in the $K$ and $J$ bands using the multi-conjugate adaptive optics (AO) demonstrator MAD \\citep{marchetti07} temporally installed at the European Southern Observatory (ESO) Very Large Telescope (VLT). The ($K,~J-K$) color-magnitude diagram (CMD) obtained from these observations led to the discovery of two well-defined red horizontal branch (HB) clumps, clearly separated in luminosity ($\\delta K\\sim 0.3$) and color (see Ferraro et al. 2009, hereafter F09; also see Figure \\ref{fig:cmd}). A prompt spectroscopic follow-up demonstrated that the two populations have the same radial velocity (hence they belong to the same stellar systems) and their metal content is different: ${\\rm [Fe/H]} \\simeq -0.2$ and ${\\rm [Fe/H]}\\simeq +0.3$ for the fainter and the brighter group, respectively. These findings and the comparison with theoretical stellar isochrones confirm the existence of two distinct stellar populations in Terzan~5 and suggest that they possibly have been generated by two bursts of star formation with a time separation of a few ($\\approx 6$) Gyrs. While the age gap can be reduced by invoking a difference in the helium content of the two populations \\citep{dantona10}, the iron enrichment and the spatial segregation of the brightest clump, together with the extraordinary amount of MSPs found in Terzan 5, indicate that this system probably experienced a particularly troubled formation and evolutionary history (see Sect. 5). Terzan~5 is the first a GC-like system in the Galactic Bulge found to have a spread in the iron content and it could be the relic of one of the building blocks that contributed to the formation of the Bulge. Indeed the discovery might represent the observational evidence that even the innermost part of galactic spheroids form (at least partially) by the accretion/merging of small, previously formed and internally evolved stellar systems \\citep[e.g.,][]{immeli04}. In this paper we present the accurate re-determination of Terzan~5 structural parameters (surface density and surface brightness profiles, total luminosity, collision rate, etc.), obtained from a combination of high-resolution and wide-field observational data. These parameters provide basic information for a deeper understanding of the origin and the evolution of this puzzling system. ", "conclusions": "The star density and SB profiles can be used to derive the integrated luminosity of the cluster. From the best-fit King model we estimate that the percentage of cluster light within regions of radius $r=15\\arcsec$, $18\\arcsec$, and $20\\arcsec$ are roughly 30\\%, 36\\% and 40\\%, respectively. Using aperture photometry on the MAD images, we obtain integrated-light values of $K(r<15\\arcsec) = 3.44$, $K(r<18\\arcsec) = 3.3$ and $K(r<20\\arcsec) = 3.2$ mag, respectively. Adopting the color excess $E(B-V)=2.38$, the distance modulus $(m-M)_0=13.87$ (Valenti et al. 2007, corresponding to a distance of $d=5.9\\pm0.5$ kpc) and the bolometric correction $BC_K = 2.4$ appropriate for a population of intrinsic color $(J-K)_0 = 0.8$ (see Montegriffo et al. 1998), we estimate that the corresponding bolometric luminosity in the considered regions is: $L_{\\rm bol} (r < 15\\arcsec) = 3 \\times 10^5 L_\\odot$, $L_{\\rm bol} (r < 18\\arcsec) = 3.4 \\times 10^5 L_\\odot$ and $L_{\\rm bol} (r < 20\\arcsec) = 3.7 \\times 10^5 L_\\odot$. Considering the fraction of light sampled in each region we find that the total luminosity of the system is $L_{\\rm bol} = 9.5\\pm0.3 \\times 10^5 L_\\odot$. An independent estimate of the total luminosity of the stellar system can be derived from its stellar population, by using a simple relation \\citep{fuel86} linking the number of stars ($N_j$) observed in a given post-main sequence evolutionary stage $j$ and the luminosity of the entire parent cluster ($L_T$): \\begin{equation} N_j = B \\times t_j \\times L_T, \\end{equation} where $B$ is the specific evolutionary flux (for intermediate/old stellar populations $B = 2 \\times 10^{-11}\\,{\\rm stars\\, yr}^{-1} L_\\odot^{-1}$) and $t_j$ is the duration of the evolutionary stage. The number of HB stars counted in the two clumps by F09 is quite large: a total of about 1300 (with 800 and 500 belonging to the faint and the bright HB clumps, respectively). This population is comparable to (or even larger than) that observed in the largest Galactic GCs, like 47 Tucanae \\citep{bec_47tuc} and NGC~6388 \\citep{ema_6388}, and suggests that the overall size of Terzan 5 (in terms of luminosity and mass) is comparable to that of these systems. For a quantitative estimate, we insert the observed number of HB stars in the above relation and adopt $t_{\\rm HB} = 10^8$ yr. This provides a luminosity of $4 \\times 10^5 L_\\odot$ and $2.5 \\times 10^5 L_\\odot$ for the two parent populations, and a total luminosity of $6.5 \\times 10^5 L_{\\odot}$ for the entire stellar system. This estimate, which is distance and reddening independent, is quite consistent with the previous one, thus confirming that Terzan~5 has a considerable total luminosity (hereafter we adopt the average value $L_{\\rm bol} = 8 \\times 10^5 L_\\odot$), significantly higher than previously thought. By comparison, adopting the values of distance and reddening quoted above and a bolometric correction $L_{\\rm bol}\\simeq 1.4\\,L_V$, the total bolometric luminosity corresponding to the integrated magnitude ($V_t=13.85$) quoted by \\citet{harris96} would be only $L_{\\rm bol}\\simeq 10^5 L_\\odot$. The discrepancy is most probably due to the strong (differential) reddening affecting the system, especially in the optical bands. This effect is greatly reduced for our new estimate, since it is based on the observed $K$-band integrated magnitude and the number of HB stars. By assuming a mass-to-light ratio $M/L_{\\rm bol}=3$ \\citep[e.g.,][]{maraston98}, the total stellar mass of this system is $M_{\\rm T} \\simeq 2 \\times 10^6 M_\\odot$. \\citet{verhut87} first suggested that the collision rate of Terzan~5 is the highest among the Galactic GCs. We can now re-compute this quantity by adopting the newly determined parameters. Following \\citet{verhut87}, the collisional parameter ($\\Gamma$) for a King-model or virialized system can be computed as: $\\Gamma \\propto \\rho_0 \\times r_c^{0.5}$, where $\\rho_0$ is the central mass density. By using the values obtained above and equation (7) of \\citet{djorg93}, we find that the collision parameter of Terzan~5 is between 5 and 10 times higher than that of Liller 1 and of other massive clusters for which structural parameters have been recently re-determined \\citep[NGC6388, NGC6266, 47Tuc;][respectively]{ema_6388, bec_6266, mapelli06}. Hence we confirm that, even with the new structural parameters (suggesting a lower concentration and a larger mass than previously thought), Terzan~5 still has the largest known collision rate of any stellar aggregate in the Galaxy. The co-existence of two stellar populations with different iron content (and probably ages) suggests that the original mass of Terzan~5 was significantly larger in the past than observed today, large enough to retain the iron-enriched gas that, otherwise, would have been ejected out from the system by the violent supernova (SN) explosions. Indeed, the smallest systems with solid evidences of a spread in the iron content (and ages) are significantly more massive than GCs: the dwarf spheroidal satellites of the Milky Way typically have masses of $\\sim 10^7 M_\\odot$ \\citep[Strigari et al. 2008; see also][]{battaglia08} and, following recent chemo-dynamical models well reproducing the observations, their initial masses amounted to a few $10^8 M_\\odot$ \\citep{revaz09}. While a lower limit of $\\sim 10^7 M_\\odot$ for the proto-Terzan 5 could also be hazarded following \\citet{baumgardt08}, more detailed and extensive simulations are needed to firmly determine the smallest total mass necessary to retain the SN ejecta. The exceptionally high metallicity regime of the two stellar populations found in Terzan~5 also suggests a quite efficient enrichment process, that could have a relevant role in the origin of its population of MSPs. In particular, both the iron and the [$\\alpha/$Fe] abundance ratios measured in Terzan~5 \\citep[][Rich et al. 2010, in preparation]{origlia04} show a remarkable similarity with those of the Bulge stars. This strongly suggests that these two structures shared the same star formation and chemical enrichment processes. The many observations of Bulge stars \\citep[e.g.,][and references therein]{melendez08,origlia08,ryde09} indicate that they are all characterized by an old age, a high (close to solar) average metallicity [Fe/H], and an [$\\alpha/$Fe] ratio which is enhanced (due to SNII enrichment) up to a metallicity [Fe/H]$\\simeq 0$. These constraints suggest a scenario where the dominant stellar population of the Bulge formed early (thus explaining the old age\\footnote{\\scriptsize{Additional episodes of star formation mainly confined in the innermost ($\\sim 100$ pc) region could eventually explain the presence of younger stars \\citep[e.g.,][]{blum03,figer04}}.}), rapidly and with high efficiency (from a gas mainly enriched by SNII, thus explaining the [$\\alpha/$Fe] enhancement up to high iron contents\\footnote{\\scriptsize{The [$\\alpha/$Fe]--[Fe/H] relation shows a down-turn at a value of [Fe/H] which depends on the star formation rate: the higher the latter, the higher the metallicity at which the down-turn occurs. Such a value is [Fe/H]$\\simeq -1$ in the Old Halo/Disk, while it is significantly higher ([Fe/H]$\\simeq 0$) in the Bulge, testifying a much higher star formation rate in this dense environment.}}). Also chemical evolution models \\citep[e.g.,][]{ballero07,mcwilliam08} indicate that the abundance patterns observed in the Bulge require a quite high star formation efficiency and an initial mass function flatter than that in the solar neighbourhood, thus to rapidly enrich the gas up to about solar metallicity through an exceptionally large amount of SNII explosions. The assumption of a similar scenario for Terzan~5 would naturally explain its extraordinary population of MSPs, since the expected high number of SNII would produce a large population of neutron stars, most of which would have been retained by the deep potential well of the massive proto-Terzan~5 system. Then the high collision rate could have favoured the formation of binary systems containing neutron stars and promoted the re-cycling process that finally generated the large population of MSPs now observed in Terzan~5. If such a scenario is correct, many more MSPs still wait to be discovered in this system \\citep[see also][]{ransom05}, the 33 known objects probably being just the tip of the iceberg. Future deeper pulsar searches of Terzan~5, perhaps with larger telescopes such as the Square Kilometer Array, will shed additional light on the nature of this system." }, "1005/1005.0842_arXiv.txt": { "abstract": "We present a comprehensive spectral analysis of all \\integral\\/ data obtained so far for the X-ray--bright Seyfert galaxy NGC 4151. We also use all contemporaneous data from \\xte, \\xmm, \\swift\\/ and \\suzaku. We find a linear correlation between the medium and hard-energy X-ray fluxes measured by \\integral, which indicates an almost constant spectral index over six years. The majority of \\integral\\/ observations were made when the source was either at a very bright or very dim hard--X-ray state. We find that thermal Comptonization models applied to the bright state yields the plasma temperature of $\\simeq 50$--70 keV and its optical depth of $\\simeq 1.3$--2.6, depending on the assumed source geometry. For the dim state, these parameters are in the ranges of $\\simeq 180$--230 keV and $\\simeq 0.3$--0.7, respectively. The Compton parameter is $y\\simeq 1$ for all the spectra, indicating a stable geometry. Using this result, we can determine the reflection effective solid angles associated with the close and distant reprocessing media as $\\simeq 0.3\\times 2\\upi$ and $0.2 \\times 2\\upi$, respectively. The plasma energy balance, the weak disc reflection and a comparison of the UV fluxes illuminating the plasma to the observed ones are all consistent with an inner hot accretion surrounded by an outer cold disc. The disc truncation radius can be determined from an approximate equipartition between the observed UV and X-ray emission, and from the fitted disc blackbody model, as $\\sim 15$ gravitational radii. Alternatively, our results can be explained by a mildly relativistic coronal outflow. ", "introduction": "\\label{intro} Continuum properties of the hard X-ray and soft $\\gamma$-ray emission from radio-quiet Seyfert galaxies nuclei are relatively well known thanks to the many satellites operating during the last decades. The spectra are commonly approximated by a phenomenological e-folded power-law model, $F(E)\\propto E^{1-\\Gamma} \\exp(-E/E_{\\rm c})$ (where $\\Gamma$ is the photon index and $E_{\\rm c}$ is the e-folding, or cut-off, energy), accompanied by a Compton reflection component. A recent study based on a local ($z < 0.1$) sample of 105 Seyfert galaxies observed with \\sax\\/ presents the average parameters of this model for Seyfert 1s and 2s \\citep{Dadina2008}. The mean $\\Gamma$ is $\\approx 1.9$ (1.8) for Seyfert 1 (2) nuclei, $E_{\\rm c}$ is $\\simeq 230$ (380) keV, and the average relative strength of reflection, $R\\equiv \\Omega/2\\upi$ (where $\\Omega$ is the effective solid angle subtended by the reflector), is $\\simeq 1.2$ (0.9). See \\citet{Beckmann2009} for the similar \\integral\\/ results. This phenomenological model is too crude for studying of the physics of the central engine of Seyferts. There is a general consensus that their X$\\gamma$ photons are from (predominantly thermal) Comptonization by hot electrons of some seed soft photons. The seed photons may come from an optically-thick accretion disc or clouds in the vicinity of the hot plasma, or may be internally produced by the synchrotron process, e.g., \\citet{Xie2010}. For example, non-simultaneous data for the Seyfert NGC 5548 from \\rosat, \\ginga\\/ and \\gro/OSSE have yielded the plasma temperature, $kT_{\\rm e}\\simeq 55$ keV, and the Thomson optical depth, $\\tau\\simeq 2$ \\citep{Magdziarz1998}. The average OSSE spectra of 11 Sy 1s and 8 Sy 2s have yielded similar parameters, $kT_{\\rm e}\\simeq 70$--80 keV, $\\tau\\simeq 1.7$ \\citep*{Zdziarski2000}. On the other hand, \\citet{Petrucci2000} obtained a much higher $kT_{\\rm e}\\simeq 250$ keV (and $\\tau\\simeq 0.2$--0.4) for NGC 5548 using \\sax\\/ data and a different anisotropic Comptonization model in the slab geometry. Then, \\citet{Petrucci2001a}, hereafter P01, found, with the same model, a similarly high values of $kT_{\\rm e}\\simeq 170$--320 keV (and $\\tau\\simeq 0.05$--0.20) for 6 Sy 1s observed by \\sax. We note, however, that such high $kT_{\\rm e}$ yield spectra being well above the OSSE fluxes at $\\ga 100$ keV for either NGC 5548, NGC 4151 or the average Seyferts. The nearby, $z = 0.0033$, Seyfert 1.5 galaxy NGC 4151 is the second (after Cen A) brightest persistent AGN in the 20--100 keV band. Its electromagnetic spectrum has been extensively studied, with its properties in the radio, infrared, optical, ultraviolet and X$\\gamma$ bands well established. Its black-hole mass has been estimated based on reverberation as $(4.6^{+0.6}_{-0.5}) \\times 10^{7}\\msun$ \\citep{Bentz2006}. Although this estimate is subject to a systematic uncertainty by a factor of $\\sim$3--4, it is consistent with the estimate based on scaling of the power spectrum to Cyg X-1 of \\citet{Czerny2001}. The overall picture of the region surrounding the nucleus, based on {\\it HST}/STIS, \\chandrae and \\xmme data, is that the complex absorption observed for NGC 4151 is due to massive outflows, presumably disc winds, forming several regions characterized by a range of column densities, $N_{\\rm H}$, and ionization levels \\citep{Schurch2004,Kraemer2006}. This explains earlier results from \\rosat, \\asca, \\ginga\\/ and \\xte\\/ satellites, with the medium-energy X-ray spectra fitted only after applying a complex absorber model, consisting of several components fully and partially covering the central source (e.g., \\citealt*{Zdziarski1996a,Zdziarski2002}, hereafter Z02). The absorber undergoes rapid daily changes of $N_{\\rm H}$ \\citep{Puccetti2007}. Also, a narrow Fe $K_{\\alpha}$ line with the equivalent width of 50--200 eV is observed (Z02; \\citealt{Schurch2003,DeRosa2007}). NGC 4151 has frequently been observed by \\gro\\/ and \\sax. The e-folded power law model yields $E_{\\rm c}\\simeq 50$--200 keV \\citep{Johnson1997,DeRosa2007}. Thermal Compton models yield $kT_{\\rm e}\\simeq 50$--80 keV, $\\tau\\simeq 1$--2 (\\citealt{Johnson1997}; Z02). The strength of Compton reflection is moderate, $R \\approx 0.4$ (e.g., \\citealt{DeRosa2007}). NGC 4151 was observed by \\integral\\/ 8 times as a primary or secondary target. The results of the first dedicated observation of 2003 May are given in \\citet{Beckmann2005}. The object was found in a spectral state similar to those observed with OSSE, but at the highest flux level ever noticed. There was no significant spectral variability despite $\\sim 50$ per cent variations of the flux. A thermal Compton model fitted to the summed spectrum yielded $kT_{\\rm e}\\simeq 90$ keV, $\\tau\\simeq 1.3$, $R\\simeq 0.7$, $N_{\\rm H}\\simeq 7 \\times 10^{22}$ cm$^{-2}$. During almost all later \\integral\\/ observations the source was found at moderate to very low flux levels. This has allowed us to study the spectral properties of the dim state of this AGN with unprecedented precision and compare them with the bright state of 2003 May. \\section[]{Observations and Data Reduction} \\label{data} We use all NGC 4151 data collected by \\integral\\/ as of 2010 January, see Table \\ref{obsint}. Good quality data come from the dedicated NGC 4151 observations made in 2003 May, 2007 January, May and December, and 2008 May. The remaining data are taken from observations of the Coma cluster, Mrk 273, NGC 4736, M51 and Mrk 421, when NGC 4151 was almost always seen at an off-axis angle of 9\\degr--15\\degr. We select the data with the off-axis angle $<15\\degr$ for ISGRI and SPI, and $<3\\degr$ for the JEM-X and OMC. The data have been reduced using the Offline Scientific Analysis ({\\sc osa}) 7.0 provided by the \\integral\\/ Science Data Centre \\citep{Courvoisier2003}, with the pipeline parameters set to the default values. Since the {\\sc osa} v.\\ 8.0 of 2009 August brought a major improvement for the JEM-X data analysis, we have used it for these data. The ISGRI and SPI spectra and light curves have been extracted with the standard spectral extraction software, including the catalogue sources NGC 4151, NGC 4051, NGC 4138, Mrk 766, NGC 4258, NGC 4395, Coma Cluster, NGC 5033, Mrk 421 and Mrk 268. The JEM-X spectra have been obtained from mosaic images stacked for a given observation. We use the standard response files for all instruments. The OMC magnitude was converted into flux using the calibration of \\citet{Johnson1966}, $F({\\rm V}\\!=\\!0)= 3.92\\times 10^{-9}$ erg cm$^{-2}$ s$^{-1}$ \\AA$^{-1}$. \\begin{table*} \\centering \\caption{The observation log of \\integral. The ISGRI exposure is effective, corresponding to fully coded observations, and the JEM-X one is summed over all pointings with the off-axis angle $<3\\degr$. The last column gives the spectral set as defined in Section \\ref{assumptions}; B = bright, D = dim, M = medium.} \\label{obsint} \\begin{tabular}{@{}cccccccc@{}} \\hline Revolution & Start time, UTC (MJD) & End time, UTC (MJD) & Eff. exposure [s] & JEM-X exposure [s] & Spectral set \\\\ \\hline 0036 & 2003-01-29 16:43 (52668.697) & 2003-01-31 08:18 (52670.346) & 11296 & -- & -- \\\\ 0071 & 2003-05-14 12:42 (52773.529) & 2003-05-16 11:39 (52775.485) & 9908 & -- & -- \\\\ 0072 & 2003-05-17 13:22 (52776.557) & 2003-05-18 19:56 (52777.831) & 5376 & -- & -- \\\\ 0073 & 2003-05-20 18:44 (52779.781) & 2003-05-22 10:23 (52781.433) & 20441 & -- & -- \\\\ 0074 & 2003-05-23 08:35 (52782.358) & 2003-05-25 11:42 (52784.488) & 117528 & 91633 & B \\\\ 0075 & 2003-05-25 21:03 (52784.877) & 2003-05-28 05:33 (52787.231) & 153481 & 153481 & B \\\\ 0076 & 2003-05-28 20:51 (52787.869) & 2003-05-29 12:13 (52788.509) & 39533 & 39533 & -- \\\\ 0274 & 2005-01-10 16:05 (53380.670) & 2005-01-12 17:36 (53382.733) & 9146 & -- & -- \\\\ 0275 & 2005-01-13 19:36 (53383.817) & 2005-01-15 04:20 (53385.181) & 11745 & -- & -- \\\\ 0310 & 2005-04-29 16:32 (53489.689) & 2005-04-30 03:46 (53490.157) & 1734 & -- & D \\\\ 0311 & 2005-05-01 08:54 (53491.371) & 2005-05-02 20:28 (53492.853) & 3290 & -- & D \\\\ 0312 & 2005-05-03 22:44 (53493.947) & 2005-05-04 09:57 (53494.416) & 1062 & -- & D \\\\ 0317 & 2005-05-19 18:04 (53509.753) & 2005-05-21 01:09 (53511.048) & 10144 & -- & D \\\\ 0318 & 2005-05-22 01:03 (53512.044) & 2005-05-24 06:37 (53514.276) & 7894 & -- & D \\\\ 0324 & 2005-06-09 17:55 (53530.747) & 2005-06-10 23:12 (53531.967) & 9748 & -- & D \\\\ 0448 & 2006-06-14 09:54 (53900.413) & 2006-06-17 00:03 (53903.002) & 29525 & -- & D \\\\ 0449 & 2006-06-17 10:19 (53903.430) & 2006-06-19 23:47 (53905.991) & 28497 & -- & D \\\\ 0450 & 2006-06-20 09:23 (53906.391) & 2006-06-22 23:29 (53908.978) & 26974 & -- & D \\\\ 0451 & 2006-06-23 09:06 (53909.379) & 2006-06-25 23:10 (53911.965) & 26844 & -- & D \\\\ 0521 & 2007-01-18 16:45 (54118.698) & 2007-01-20 23:28 (54120.978) & 122878 & 122878 & D \\\\ 0522 & 2007-01-21 16:35 (54121.691) & 2007-01-24 06:07 (54124.255) & 123751 & 105848 & D \\\\ 0561 & 2007-05-18 09:31 (54238.397) & 2007-05-20 16:11 (54240.674) & 110696 & 110696 & D \\\\ 0562 & 2007-05-22 00:44 (54242.031) & 2007-05-23 21:34 (54243.899) & 100729 & 78177 & D \\\\ 0563 & 2007-05-24 09:06 (54244.379) & 2007-05-25 06:31 (54245.272) & 46903 & 46903 & D \\\\ 0634 & 2007-12-22 17:53 (54456.745) & 2007-12-25 07:19 (54459.305) & 131811 & 131811 & M \\\\ 0636 & 2007-12-28 17:26 (54462.726) & 2007-12-31 01:53 (54465.078) & 119738 & 100028 & M \\\\ 0678 & 2008-05-02 21:17 (54588.887) & 2008-05-04 22:36 (54590.942) & 24805 & -- & M \\\\ 0679 & 2008-05-05 20:47 (54591.866) & 2008-05-07 07:09 (54593.298) & 17266 & -- & M \\\\ 0809 & 2009-05-31 11:31 (54982.480) & 2009-05-31 20:58 (54982.874) & 20019 & 2891 & -- \\\\ 0810 & 2009-06-01 05:49 (54983.242) & 2009-06-03 16:10 (54985.674) & 112086 & 46546 & -- \\\\ 0811 & 2009-06-04 05:37 (54986.234) & 2009-06-06 10:24 (54988.433) & 108501 & 40239 & -- \\\\ \\hline \\end{tabular} \\end{table*} To better constrain the spectra at low energies, we supplemented the \\integral\\/ data by all available X-ray observations of NGC 4151 taken since 2003 by \\xte, \\xmm\\/ and \\suzaku, see Table \\ref{obsother}. The \\xte\\/\\/ PCU2 light curves and spectra and HEXTE spectra were extracted with the {\\sc heasoft} 6.5.1, using standard selection criteria. The 5 \\xmm\\/ observations (denoted here X1--X5) were analyzed with {\\sc sas} 8.0.1. We used only the EPIC pn data, and excluded periods of high or unstable background. The spectra X1 and X2 have been found to be compatible with each other; we have therefore added them together. For the single \\suzaku\\/ observation, the data in the $3\\times 3$ mode were reduced with {\\sc heasoft} 6.6. The spectra of the front-illuminated CCD, XIS0 and XIS3, were added together, whereas the spectrum of the back-illuminated detector XIS1 was used separately. We also extracted the HXD/PIN spectrum using the standard procedure. Although there have been a number of \\chandra\\/ NGC 4151 observations, all those without gratings that were public at the time we started the analysis were from before 2000 March. Tables \\ref{obsint}--\\ref{obsother} identify the data used for our three main spectral sets, bright (B), medium (M), and dim (D), whose selection is based on ISGRI flux. The remaining \\integral\\/ data, not marked with a letter in Table \\ref{obsint}, were not used for spectral analysis. We have excluded them because either there were no corresponding high-quality spectra from X-ray satellites (e.g., Revs.\\ 0809--0811) and/or their flux level was outside the range assumed by us (e.g., beginning of Rev.\\ 0074 and Rev.\\ 0076 in the case of the B state). The data sets from the other satellites were not always simultaneous with the \\integral\\/ data. Thus, we assign them to one of the three sets based again on the flux, using the monitoring data from \\swift/BAT (see Sections \\ref{variability}, \\ref{assumptions} for details of the selection). \\begin{table*} \\centering \\caption{The observation log for the other X-ray satellites. The exposure times are given for the \\xte\\/ PCU2 (where the number of pointings added together is shown in parentheses), EPIC pn (\\xmm) and XIS (\\suzaku) detectors. The last column shows the data set as defined in Table \\ref{obsint}, and, in the case of \\xmm, its consecutive observation number.} \\label{obsother} \\begin{tabular}{@{}lcccccr@{}} \\hline Obs. ID & Start time, UTC (MJD) & End time, UTC (MJD) & Exposure & Spectral set \\\\ \\hline \\multicolumn{5}{c}{\\it RXTE} \\\\ 80416-01-01-(00--09) & 2003-05-24 05:49 (52783.242) & 2003-05-29 04:18 (52788.179) & 15760 (11) & B \\\\ 92113-08-(06--09) & 2006-05-13 02:28 (53868.103) & 2006-06-23 15:42 (53909.654) & 4240 (4) & D \\\\ 92113-08-(22--35) & 2006-12-23 12:51 (54092.535) & 2007-06-22 08:57 (54273.373) & 14272 (14) & D \\\\ \\hline \\multicolumn{5}{c}{\\it XMM-Newton} \\\\ 0143500101 & 2003-05-25 01:38 (52784.068) & 2003-05-25 06:54 (52784.288) & 5821 & X1, B \\\\ 0143500201 & 2003-05-26 20:35 (52785.858) & 2003-05-27 01:51 (52786.077) & 11389 & X2, B \\\\ 0143500301 & 2003-05-27 15:17 (52786.637) & 2003-05-27 20:33 (52786.856) & 12291 & X3, B \\\\ 0402660101 & 2006-05-16 06:22 (53871.265) & 2006-05-16 17:35 (53871.733) & 27980 & X4, D \\\\ 0402660201 & 2006-11-29 17:20 (54068.722) & 2006-11-30 08:00 (54069.333) & 21468 & X5, M \\\\ \\hline \\multicolumn{5}{c}{\\it Suzaku} \\\\ 701034010 & 2006-12-18 20:05 (54087.837) & 2006-12-21 09:14 (54090.385) & 124980 & D \\\\ \\hline \\end{tabular} \\end{table*} \\section[]{Variability} \\label{variability} NGC 4151 has been observed with X$\\gamma$ detectors for almost 40 years. The satellite and balloon results up to 1988 were compiled by \\citet{Perotti1991}. Then, it has been observed above 20 keV by \\ginga, \\granat, \\gro, \\sax, \\integral, \\swifte and \\suzaku. Fig.\\ \\ref{longest}(a) shows the 20--100 keV flux observed since 1972 October compared to the optical flux and the \\xte/ASM count rate. Because the spectra from earlier observations are not publicly available, the 20--100 keV fluxes were determined using the values of the flux at 35 keV and of the photon index presented in table 2 of \\citet{Perotti1991}. For this reason, we do not show their uncertainties. Similarly, the \\granat\\/ fluxes were determined using table 3a of \\citet{Finoguenov1995}, where the ART-P and SIGMA spectra are fitted by a power-law model. For all other observations, we use the spectra from {\\sc heasarc} (\\gro/OSSE, \\sax/PDS) or spectra extracted by us (\\integral/ISGRI, \\suzaku/PIN), where the 20--100 keV flux was computed from a power-law model fit (which was done in the 50--150 keV range for the OSSE data) and the uncertainty was determined from the relative error of the summed count rate in the fitted band. Fig.\\ \\ref{longest}(b) shows the 20--100 keV fluxes from the \\gro/BATSE in 7-d bins (together with those from OSSE). They have been obtained from the 20--70 keV fluxes given by \\citet{Parsons1998} by multiplying by 1.1 (assuming $\\Gamma=1.8$) and then dividing by the normalization factor with respect to OSSE of 1.42 \\citep{Parsons1998}. Fig.\\ \\ref{longest}(b) gives the 1.5--12 keV count rate of \\xte/ASM\\footnote{http://xte.mit.edu/asmlc/ASM.html} in 30-d bins. The blue curve in Fig.\\ \\ref{longest}(c) shows the optical data from the Crimean observatory in the V band (5500 \\AA, \\citealt{Czerny2003}), which, after MJD 52000, are supplemented by the \\integral/OMC data. Since the V band data do not overlap much in time with the ASM data, we also show the 5117 \\AA\\ fluxes from various observatories given by \\citet{Shapovalova2008}. \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_longest.eps} \\caption{Light curves of NGC 4151 since 1972. (a) The 20--100 keV flux from the compilation by \\citet{Perotti1991}, magenta diamonds, and from \\granat/SIGMA (cyan squares), \\gro/OSSE (red filled circles), \\sax/PDS (green triangles), \\integral/ISGRI (blue dots), and \\suzaku/PIN (the orange open circle). The dotted lines show the extrema of the 20--100 keV flux from ISGRI. (b) The 20--100 keV flux from \\gro/BATSE (green error bars) together with that from OSSE (red filled circles; same as in a). (c) The \\xte/ASM count rate in 30-d bins. (d) The 5500 \\AA\\ flux (\\citealt{Czerny2003}, blue lines), and from \\integral/OMC (vertical blue bars at MJD $> 52000$), and the 5117 \\AA\\ flux (\\citealt{Shapovalova2008}, black dots). } \\label{longest} \\end{figure} The 1.5--12 keV count rate appears well correlated with the optical fluxes, see Fig.\\ \\ref{longest}. As found by \\citet{Czerny2003}, the medium-energy X-ray flux of NGC 4151 varies on time scales $\\sim 5$--$10^3$ d, whereas the optical variability is also present on longer time scales. For hard X-rays, a correlation with the optical and the 1.5--12 keV fluxes is less clear. The main peak of the optical emission at MJD $\\simeq 49000$--51000 (1993--1998) is reflected in the 1.5--12 keV flux but not in the 20--100 keV one (including the BATSE data). The later data from \\sax, \\integral\\/ and \\suzaku\\/ show an overall agreement with the optical and softer X-ray fluxes in a sense that the minima at MJD $\\sim$52000 and MJD $\\sim$54000 and the maximum at MJD $\\sim$52800 appear for all these bands. However, the scarce hard X-ray coverage prevents unambiguous conclusions. \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_monitor.eps} \\caption{Medium and hard X-ray light curves from \\integral/ISGRI compared to those from \\xte\\/ and \\swift\\/ (shown with black circles and error bars). The 18--50 keV ISGRI flux is shown by open circles on all panels. The colours identify the three main states, bright (blue), medium (green), dim (red) and, additionally, data excluded from the spectral analysis, Revs.\\ 0076 and 0809--0811 (magenta), and the rest (cyan). (a) The \\xte/ASM count rate in 30-d bins (right axis). (b) The \\xte/PCU2 count rate (right axis). The vertical lines delineate the two periods for which the dim state PCA spectrum was obtained. (c) Comparison with the \\swift/BAT 15--50 keV flux in 7-d bins. The vertical lines show the times of the X4 (dim), X5 (medium) \\xmm\\/ observations and that of \\suzaku\\/ (dim). The horizontal dashed lines show the average flux levels of ISGRI bright, medium and dim states. } \\label{long} \\end{figure} The results shown in Fig.\\ \\ref{longest}(a--b) show that the hard X-ray emission of NGC 4151 has varied in well defined limits over the last 39 years. Almost all fluxes are within the extreme 20--100 keV fluxes from ISGRI (with the maximum in 2003 May and the minimum in 2007 May). There are four flux measurements from older (1975--77) observations above this maximum. However, they are relatively doubtful, corresponding to marginal ($\\approx 2$--$3\\sigma$) detections at $\\ga$ 40 keV. Thus, ISGRI 20--100 keV flux range appears very close to the corresponding overall actual range. This is also supported by the BATSE data, Fig.\\ \\ref{longest}(b), varying within this range. We also notice that NGC 4151 was rarely seen at low hard X-ray fluxes before the \\integral\\/ launch. Fig.\\ \\ref{long} compares ISGRI 18--50 keV fluxes with the contemporaneous data in the medium and hard X-ray bands from \\xte\\/ and \\swift. The ISGRI fluxes are well correlated with those from both \\xte\\/ and the \\swift/BAT\\footnote{http://swift.gsfc.nasa.gov/docs/swift/results/transients}. In particular, we see a good correlation with the 1.5--12 keV rate in spite of the variable absorption affecting a lower part of this band. Fig.\\ \\ref{long} also shows that a large fraction of the \\integral\\/ observations happened during periods with very low hard X-ray flux. Fig.\\ \\ref{long} also identifies the \\integral\\/ observations used for the spectral sets (Table \\ref{obsint}). \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_short.eps} \\caption{Illustration of short time-scale variability. The blue bars (left axis) and the red dots show the 18--40 keV ISGRI flux and the optical OMC flux, respectively, during the dedicated \\integral\\/ observations. The vertical lines and the green arrows in the top panel show the times of the bright-state X1--X3 \\xmm\\/ and 11 \\xte/PCA observations, respectively. The data used for the bright \\integral\\/ spectrum are those in the top panel simultaneous with the OMC observations except those around MJD 52788 (Rev.\\ 0076), which have markedly lower fluxes. } \\label{short} \\end{figure} The BAT light curve, Fig. \\ref{long}(c), shows that the hard X-ray flux varies by a factor of a few on a week-time scale. When the source is sufficiently bright, it is possible for ISGRI to monitor hour time-scale variability. This was the case for the bright state of 2003 May, and the source showed such variability, see Fig.\\ \\ref{short}. The shown ISGRI and OMC light curves were extracted with a time bin equal to a single pointing duration (10 m--2.5 h, typically $\\sim 1$ h). When observed in dimmer states, ISGRI flux remains constant within the measurement errors during a given continuous observation. The optical flux appears constant on hour--day time scales. Still, it varies on longer time scales, and it increases by $\\sim$30 per cent when the source is bright in X-rays (2003 May, 2009 June). We have extracted the bright (X1--X3) \\xmm/EPIC pn light curves in several energy bands in 100-s time bins, and compared with the OMC 100-s light curves. No variability is seen in either light curves on time scales $\\la 1$ h. The \\xmm\\/ light curves vary slowly in good agreement with the trends observed for ISGRI (and also JEM-X) fluxes. \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_correl.eps} \\caption{(a) The OMC 5500 \\AA\\ flux vs.\\ the ISGRI 18--40 keV flux. (b) The JEM-X 3--10 keV flux vs.\\ the ISGRI flux. Each cross corresponds to a single \\integral\\/ pointing. The colours identify the three main states, bright (blue), medium (green), dim (red), and, additionally, Revs.\\ 0076, 0809--0811 (magenta). } \\label{correl} \\end{figure} The \\integral\\/ data are suitable for correlation studies because they provide simultaneous data in the optical, medium and hard X-ray bands. Here, we use all the observations with the off-axis angle $<3\\degr$ for which OMC and JEM-X data are available. They include also part of the observations not used in our spectral analysis, namely those identified with magenta circles in Fig.\\ \\ref{long}. Fig.\\ \\ref{correl} shows the correlations between the 18--40 keV flux and those fluxes in the 3--10 keV and V bands. Both show strong correlations, but the correlation patterns are different. The optical fluxes fall into two separate regions, within each they are correlated with the hard X-rays. Significant rank-order probabilities ($>\\! 0.999$) of the ISGRI/OMC correlation are found separately for each of the regions, but also and for the entire data set. The observed bimodality may result from strong optical long-term variability between the X-ray observations. On the other hand, the 3--10 keV and 18--40 keV fluxes exhibit a single, extremely strong, linear correlation. Statistical uncertainty blurs somewhat the correlation for low fluxes. This result indicates that during all \\integral\\/ observations the slope of the X-ray spectrum of NGC 4151 below 40 keV remains approximately constant. \\section[]{Spectral analysis} \\label{spectral} \\subsection{Assumptions and selection of spectral sets} \\label{assumptions} Spectral fitting is performed with {\\sc xspec} 11.3 \\citep{Arnaud1996}. Errors are given for 90 per cent confidence level for a single parameter, $\\Delta \\chi^{2} = 2.7$. The luminosity and distance, $D=13.2$ Mpc, are for $H_{0} = 75$ km s$^{-1}$ Mpc$^{-1}$. We use the elemental abundances of \\citet{Anders1982} and the photoelectric absorption cross-sections from \\citet{Balucinska1992}. The Galactic column density is set to $N_{\\rm H}^{\\rm G}= 2.1\\times 10^{20}$ cm$^{-2}$. The assumed inclination angle of the reflector and the hot plasma (for anisotropic geometries) is $i=45\\degr$ \\citep{Das2005}. The reflector is assumed to be neutral, and relativistic broadening of reflected spectra is neglected. Still, in some cases we test for the effects of either a variable Fe abundance of the intrinsic absorber and the reflector, a variable $i$, or an ionization of the reflector. The main assumption used in our selection of spectral sets is that the state of the emitting hot plasma is determined by the hard X-ray flux. This is justified by our finding that the shape of ISGRI spectra for a given revolution is almost constant at a given hard X-ray flux. As an additional constraint, we do not split \\integral\\/ data from a single revolution into different spectral sets. Based on this criterion, we select three average ISGRI spectra, as specified in Table \\ref{obsint} and Fig.\\ \\ref{long}. These spectra are accompanied by spectra taken with the other instruments. The state is assigned to each of them on the basis of the light curves presented in Figs.\\ \\ref{long}--\\ref{short}. In particular, the high quality of the \\swift/BAT data allows us to determine the hard X-ray flux for periods with no \\integral\\/ observations, and thus to assign the correct state to the existing lower-energy data sets (Table \\ref{obsother}). On the other hand, the data from \\xte\\/ determine by themselves the hard X-ray state. The ISGRI bright (B) and dim (D) spectra are of relatively good quality, and correspond to the extrema of the measured hard X-ray flux. For the flux range in-between, we identify two flux ranges. However, we had no corresponding low-energy data sets for the upper of them. Thus, we use the lower of them, denoted as medium (M). The fitted data sets are as follows. The bright state, B. We use the data from 2003 May made at low off-axis angles, as shown in the top panel of Fig.\\ \\ref{short}. The spectrum includes a part of Rev.\\ 0074 (without its beginning) and Rev.\\ 0075. For these data, we have simultaneous \\xte\\/ and \\xmm\\/ observations, see Table \\ref{obsother}. The fitted spectral set consists of ISGRI spectrum, the X1+X2 and X3 \\xmm\\/ spectra, and the \\xte\\/ PCU2 spectrum summed over 11 observations. The medium state, M. We sum ISGRI data from Revs.\\ 0634, 0636, 0678--0679, and use the summed JEM-X 1 spectrum from Revs.\\ 0634 and 0636, and the X5 \\xmm\\/ spectrum. The dim state, D. The summed ISGRI spectrum is from Revs.\\ 0310--0563. We use the \\xte\\/ PCU2, the X4 \\xmm\\/ EPIC pn, and the \\suzaku\\/ XIS0,3 and XIS1 spectra. We have also investigated the effect of including other available spectra for each of the data set, whose resulting sets we denote 'extended'. The extended set B also includes the spectra from the \\integral/JEM-X 2 and SPI, and the \\xte/HEXTE clusters 0, 1. The extended set M also includes the SPI spectrum. The extended data set D also includes two \\integral\\/ JEM-X 1 summed spectra, from Revs.\\ 0521--0522 and 0561--0563, the \\suzaku\\/ PIN spectrum, the \\xte\\/ HEXTE cluster 1 spectrum, and the corresponding SPI spectrum. We discuss the effect of using the extended data sets in Section \\ref{Compton} below. In order to limit the complexity of the fitted models, we have decided not to use the X-ray spectra at $E< 2.5$ keV. A two-component partial absorption was then sufficient to model absorption of the continuum. Using the X-ray spectra in the 0.1--2.5 keV band would have required to add a number of spectral lines as well as a soft excess component. As we have tested, including that low-energy band would not affect our determination of the parameters of the hot plasma and of Compton reflection. Thus, the EPIC pn, XIS0+XIS3, and XIS1 spectra were used in the 2.5--11.3 keV, 2.5--9.5 keV and 2.5--9.0 keV bands, respectively. The PCA/PCU2 and JEM-X spectra were fitted in the 3--16 keV and 4--19 keV bands, respectively. The HXD/PIN spectrum was used in the 19--60 keV band because the $<$19-keV data were showing a strong excess. For ISGRI, data below 20 keV were always excluded, whereas the high-energy limit was in the 180--200 keV range, depending on our $3\\sigma$ detection threshold used to select good spectral channels. The ISGRI spectra in the staring mode from Revs.\\ 0074--75 show an excess of $\\sim$10 per cent below 23 keV, whose origin remains unknown. It could be related to a specific setting of the low-energy threshold in the early period of the mission. Thus, we used only the data at $>23$ keV in this case. The HEXTE spectra were used in the 13--160 keV band for the bright state and in 23--110 keV for the dim state. The SPI spectra were fitted in the 23--200 keV band. While fitting a given spectral set, the parameters of intrinsic continua are the same for all included spectral data, allowing only the normalization of the model to vary between them. This reflects our assumption that the hard X-ray flux determines the intrinsic state of the source, even for non-simultaneous data. On the other hand, absorption and the Fe K emission may vary quickly \\citep{Puccetti2007}, and we cannot expect their parameters to be the same for different low-energy spectra associated with the same spectral set. Thus, we have allowed them to vary between different spectra. Our assumed absorber consists of one fully covering the source with the column density of $N_{\\rm H}^{\\rm i}$, and the ionization parameter, $\\xi \\equiv 4\\upi F_{\\rm i}/n$, where $F_{\\rm i}$ is the 5 eV--20 keV irradiating flux and $n$ is the density of the reflector. For the low-resolution detectors, PCA and JEM-X, we assumed $\\xi=0$. For the high-resolution detectors, EPIC pn and XIS, we allowed $\\xi>0$, and, in addition, we included partial-covering neutral absorption, with the column density of $N_{\\rm H}^{p}$ covering a fraction of $f^{\\rm p}$ of the flux. For the ionized absorption, we use the model by \\citet{Done1992}. Although it is not applicable to highly ionized plasma, it is sufficiently accurate at low/moderate ionization, such as that found in our data. Absorption for ISGRI, HEXTE, SPI and PIN spectra is only marginally important, and we thus used only the neutral fully-covering $N_{\\rm H}^{\\rm i}$ at the values fitted to the corresponding PCA or JEM-X spectra. The Fe K$\\alpha$ emission is modelled by a Gaussian line, with the centre energy, width, photon flux and equivalent width of $E_{\\rm Fe}$, $\\sigma_{\\rm Fe}$, $I_{\\rm Fe}$ and $w_{\\rm Fe}$, respectively. Fitting the dim state spectra, we have found that adding an Fe K$\\beta$ line (at $\\approx$7.1 keV) was significantly improving the fit for the EPIC pn and XIS spectra. \\subsection{The cut-off power-law model} \\label{cutoff} We first fit an e-folded power-law model including a Compton reflection component. This allows us to compare our results to other published ones. We use the {\\sc pexrav} model \\citep{Magdziarz1995}, and our other assumptions are specified above. The results are shown in Table \\ref{pl_fits}, where we do not show the fitted absorber and line parameters since they are similar to those for our main model of thermal Comptonization, see Section \\ref{Compton} below. Our results regarding $\\Gamma$ and $R$ are similar to those of earlier works. However, a significant difference is found for the e-folding energy, $E_{\\rm c}$, which we find to be high, and compatible with no cut-off for two of our states, see Table \\ref{pl_fits}. A finite $E_{\\rm c}$ is found only for the bright state. Even then, we obtain values well above those reported before, which are 30--45 keV (P01), 80--180 keV \\citep{DeRosa2007} and $\\sim$210 keV (Z02). As discussed in \\citet{Zdziarski2003}, an exponential cut-off is much shallower than that characteristic of thermal Comptonization, and not sharp enough to model the spectral high-energy cut-offs observed in Seyferts, in particular in NGC 4151. Thus, our high values of $E_{\\rm c}$ result from the good quality of the data below the cut-off, which dominate the statistics and force an approximately straight power law to continue to just below the beginning of the cut-off. \\begin{table} \\setlength{\\tabcolsep}{1.2mm} \\begin{center}{ \\caption{Spectral fitting results for the e-folded power-law model including reflection. The {\\sc xspec} model is {\\sc constant*wabs*absori*zpcfabs(pexrav\\ +zgauss)}, and the normalization, $K$, is given by the photon flux at 1 keV in units of $10^{-2}$ keV$^{-1}$ cm$^{-2}$ s$^{-1}$. } \\small \\label{pl_fits} \\begin{tabular}{cccccc} \\hline State & $E_{\\rm c}$ [keV] & $\\Gamma$ & $R$ & $K$ & $\\chi ^{2}$/d.o.f. \\\\ \\hline Bright & 264$_{-26}^{+48}$ & 1.71$_{-0.01}^{+0.06}$ & 0.45$_{-0.05}^{+0.12}$ & 8.7$_{-0.3}^{+1.1}$ & 3179/3414 \\\\ Medium & $>$ 1025 & 1.81$_{-0.03}^{+0.05}$ & 0.6$_{-0.3}^{+0.3}$ & 4.6$_{-0.7}^{+1.4}$ & 1554/1599 \\\\ Dim & $>$ 1325 & 1.81$_{-0.01}^{+0.01}$ & 0.92$_{-0.05}^{+0.05}$ & 2.24$_{-0.02}^{+0.01}$ & 3338/3384 \\\\ \\hline \\end{tabular} }\\end{center} \\end{table} \\subsection{Comptonization model} \\label{Compton} \\begin{table*} \\begin{minipage}{172mm} \\begin{center} \\caption{{\\bf (a)} Spectral fitting results with thermal Comptonization as the main continuum. The {\\sc xspec} model is {\\sc constant*wabs*absori*zpcfabs*(compps\\ +zgauss)}. See Section \\ref{Compton} for the definitions of the parameters. $L_{\\rm X\\gamma}$ and $L_{\\rm X\\gamma}'$ are the bolometric model luminosity (in units of $10^{43}$ erg s$^{-1}$) with and without the reflection component, respectively, and only the values for the geometry 0 are given as the values for other geometries are similar. } \\small \\label{comp_results} \\begin{tabular}{lccccccccc} \\hline State & $kT_{\\rm e}$ [keV] & $y$ & $\\tau$ & $R$ & $K\\,[10^8]$ & $L_{\\rm X\\gamma}$ & $L_{\\rm X\\gamma}'$ &Geometry & $\\chi ^{2}$/d.o.f. \\\\ \\hline Bright & 54$_{-5}^{+7}$ & 1.10$_{-0.02}^{+0.02}$ & 2.6$_{-0.3}^{+0.3}$ & 0.40$_{-0.06}^{+0.07}$ & 5.2$_{-0.6}^{+1.4}$ & 5.18 & 4.61 & sphere (0) & 3165/3414 \\\\ & 62$_{-7}^{+7}$ & 0.60$_{-0.01}^{+0.01}$ & 1.3$_{-0.1}^{+0.1}$ & 0.55$_{-0.07}^{+0.08}$ & 7.3$_{-0.9}^{+1.9}$ & & & slab (1) & 3166/3414 \\\\ & 62$_{-7}^{+9}$ & 0.61$_{-0.01}^{+0.01}$ & 1.3$_{-0.2}^{+0.1}$ & 0.55$_{-0.06}^{+0.09}$ & 3.9$_{-0.7}^{+0.4}$ & & & slab ($-1$) & 3166/3414 \\\\ & 73$_{-9}^{+14}$ & 1.06$_{-0.02}^{+0.02}$ & 1.9$_{-0.2}^{+0.4}$ & 0.47$_{-0.07}^{+0.05}$ & 12.0$_{-1.6}^{+2.9}$ & & & cylinder (2) & 3168/3414 \\\\ & 73$_{-13}^{+13}$ & 1.06$_{-0.01}^{+0.03}$ & 1.9$_{-0.3}^{+0.3}$ & 0.51$_{-0.9}^{+0.5}$ & 5.4$_{-1.1}^{+1.4}$ & & & cylinder ($-2$) & 3168/3414 \\\\ & 73$_{-7}^{+16}$ & 1.21$_{-0.02}^{+0.02}$ & 2.1$_{-0.6}^{+0.4}$ & 0.48$_{-0.06}^{+0.06}$ & 11.9$_{-1.4}^{+3.2}$ & & & hemisphere (3) & 3168/3414 \\\\ & 73$_{-9}^{+15}$ & 1.21$_{-0.02}^{+0.03}$ & 2.1$_{-0.4}^{+0.3}$ & 0.50$_{-0.06}^{+0.08}$ & 4.7$_{-1.0}^{+1.5}$ & & & hemisphere ($-3$) & 3168/3414 \\\\ & 61$_{-5}^{+7}$ & 0.80$_{-0.02}^{+0.01}$ & 1.7$_{-0.2}^{+0.1}$ & 0.38$_{-0.05}^{+0.04}$ & 1.2$_{-0.1}^{+0.1}$ & & & sphere (4) & 3172/3414 \\\\ & 57$_{-7}^{+13}$ & 0.86$_{-0.01}^{+0.01}$ & 1.9$_{-0.4}^{+0.2}$ & 0.39$_{-0.05}^{+0.06}$ & 6.4$_{-0.6}^{+0.6}$ & & & sphere ($-4$) & 3169/3414 \\\\ & 57$_{-5}^{+10}$ & 0.86$_{-0.01}^{+0.01}$ & 1.9$_{-0.3}^{+0.2}$ & 0.38$_{-0.05}^{+0.06}$ & 4.9$_{-0.4}^{+1.5}$ & & & sphere ($-5$) & 3166/3414 \\\\ \\hline Medium & 128$_{-24}^{+74}$ & 1.02$_{-0.13}^{+0.08}$ & $1.0_{-0.3}^{+0.5}$ & $< 0.41$ & 3.9$_{-1.0}^{+0.4}$ &2.26 & 2.16 & sphere (0) & 1554/1597 \\\\ \\hline Dim & 190$_{-12}^{+13}$ & 0.98$_{-0.01}^{+0.01}$ & 0.66$_{-0.05}^{+0.04}$ & 0.75$_{-0.04}^{+0.04}$ & 1.54$_{-0.01}^{+0.01}$ &1.47 & 1.19 & sphere (0) & 3337/3384 \\\\ & 200$_{-8}^{+7}$ & 0.49$_{-0.01}^{+0.01}$ & 0.31$_{-0.01}^{+0.01}$ & 1.01$_{-0.05}^{+0.05}$ & 2.04$_{-0.02}^{+0.01}$ & & & slab (1) & 3335/3384 \\\\ & 188$_{-7}^{+6}$ & 0.48$_{-0.01}^{+0.01}$ & 0.33$_{-0.01}^{+0.01}$ & 1.04$_{-0.05}^{+0.05}$ & 1.20$_{-0.01}^{+0.01}$ & & & slab ($-1$) & 3350/3384 \\\\ & 210$_{-14}^{+12}$ & 1.05$_{-0.01}^{+0.01}$ & 0.64$_{-0.04}^{+0.04}$ & 0.79$_{-0.04}^{+0.04}$ & 2.85$_{-0.01}^{+0.02}$ & & & cylinder (2) & 3334/3384 \\\\ & 215$_{-13}^{+12}$ & 1.01$_{-0.01}^{+0.01}$ & 0.60$_{-0.03}^{+0.04}$ & 0.86$_{-0.05}^{+0.05}$ & 1.73$_{-0.01}^{+0.01}$ & & & cylinder ($-2$) & 3336/3384 \\\\ & 227$_{-14}^{+10}$ & 1.21$_{-0.02}^{+0.01}$ & 0.68$_{-0.03}^{+0.04}$ & 0.84$_{-0.04}^{+0.04}$ & 2.89$_{-0.02}^{+0.01}$ & & & hemisphere (3) & 3334/3384 \\\\ & 220$_{-11}^{+13}$ & 1.14$_{-0.01}^{+0.01}$ & 0.66$_{-0.04}^{+0.03}$ & 0.90$_{-0.05}^{+0.05}$ & 1.71$_{-0.01}^{+0.02}$ & & & hemisphere ($-3$) & 3336/3384 \\\\ & 186$_{-9}^{+11}$ & 0.66$_{-0.01}^{+0.01}$ & 0.45$_{-0.03}^{+0.02}$ & 0.77$_{-0.04}^{+0.04}$ & 0.866$_{-0.004}^{+0.004}$ & & & sphere (4) & 3341/3384 \\\\ & 191$_{-17}^{+13}$ & 0.86$_{-0.01}^{+0.01}$ & 0.58$_{-0.04}^{+0.05}$ & 0.77$_{-0.04}^{+0.04}$ & 1.56$_{-0.01}^{+0.01}$ & & & sphere ($-4$) & 3337/3384 \\\\ & 196$_{-13}^{+13}$ & 0.81$_{-0.01}^{+0.01}$ & 0.53$_{-0.04}^{+0.04}$ & 0.77$_{-0.04}^{+0.04}$ & 1.49$_{-0.01}^{+0.01}$ & & & sphere ($-5$) & 3337/3384 \\\\ \\hline \\end{tabular} \\end{center} \\end{minipage} \\end{table*} \\setcounter{table}{3} \\begin{table*} \\begin{minipage}{172mm} \\begin{center} \\caption{{\\bf (b)} Continuation of Table \\ref{comp_results}(a). Fit results with the thermal Comptonization model (for a sphere, the geometry parameter $=0$) regarding the absorber, line component, and the relative normalization, $C_{\\rm ISGRI}$, of a given X-ray spectrum with respect to ISGRI. See Section \\ref{assumptions} for the definitions of the other parameters. The units of $N_{\\rm H}$, $\\xi$, and $I_{\\rm Fe}$ are $10^{22}$ cm$^{-2}$, $10^{-2}$ erg cm s$^{-1}$, $10^{-4}$ cm$^{-2}$ s$^{-1}$, respectively, and `f' denotes a fixed parameter. } \\begin{tabular}{llccccccccc} \\hline State & X-ray spectrum & $C_{\\rm ISGRI}$ & $N_{\\rm H}^{\\rm i}$ & $\\xi$ & $N_{\\rm H}^{\\rm p}$ & $f^{\\rm p}$ & $E_{\\rm Fe}$ [keV] & $\\sigma_{\\rm Fe}$ [keV] & $I_{\\rm Fe}$ & $w_{\\rm Fe}$ [eV] \\\\ \\hline Bright & EPIC pn (X1+X2) & 0.92$_{-0.03}^{+0.03}$ & 4.9$_{-0.7}^{+0.9}$ & 10$_{-6}^{+79}$ & 15.8$_{-3.3}^{+3.9}$ & 0.39$_{-0.07}^{+0.09}$ & 6.400$_{-0.011}^{+0.011}$ & 0.07$_{-0.01}^{+0.02}$ & 3.0$_{-0.3}^{+0.3}$ & { 83$_{{{-9}}}^{{{+7}}}$ } \\\\ & EPIC pn (X3) & 1.08$_{-0.03}^{+0.03}$ & 3.0$_{-1.8}^{+1.5}$ & 64$_{-61}^{+47}$ & 9.5$_{-1.1}^{+4.0}$ & 0.55$_{-0.22}^{+0.18}$ & 6.393$_{-0.012}^{+0.013}$ & 0.07$_{-0.01}^{+0.01}$ & 2.8$_{-0.3}^{+0.3}$ & { 76$_{{{-9}}}^{{{+9}}}$ } \\\\ & PCA PCU2 (B) & 1.17$_{-0.03}^{+0.03}$ & 7.4$_{-0.4}^{+0.4}$ & 0f & 0f & -- & 6.06$_{-0.09}^{+0.08}$ & 0.36$_{-0.13}^{+0.15}$ & 6.8$_{-0.8}^{+1.3}$ & { 169$_{{{-20}}}^{{{+32}}}$ } \\\\ \\hline Medium & EPIC pn (X5) & 0.83$_{-0.06}^{+0.08}$ & 7.6$_{-0.4}^{+0.5}$ & 4$_{-3}^{+11}$ & 22.6$_{-1.5}^{+4.4}$ & 0.50$_{-0.04}^{+0.05}$ & 6.388$_{-0.007}^{+0.007}$ & 0.06$_{-0.01}^{+0.01}$ & 2.7$_{-0.3}^{+0.5}$ & { 149$_{{{-12}}}^{{{+27}}}$ } \\\\ & JEM-X (M) & 0.74$_{-0.06}^{+0.06}$ & 13.0$_{-3.3}^{+3.7}$ & 0f & 0f & -- & -- & -- & -- & -- \\\\ \\hline Dim & EPIC pn (X4) & 1.17$_{-0.01}^{+0.01}$ & 6.36$_{-0.15}^{+0.14}$ & 4$_{-3}^{+30}$ & 23.5$_{-0.7}^{+0.6}$ & 0.69$_{-0.01}^{+0.01}$ & 6.398$_{-0.005}^{+0.004}$ & 0.06$_{-0.01}^{+0.01}$ & 2.3$_{-0.1}^{+0.1}$ & { 271$_{{{-13}}}^{{{+13}}}$ } \\\\ & XIS0,3 (D) & 1.11$_{-0.01}^{+0.01}$ & 8.03$_{-0.16}^{+0.12}$ & 0.7$_{-0.4}^{+0.9}$ & 28.8$_{-1.0}^{+0.9}$ & 0.50$_{-0.01}^{+0.01}$ & 6.384$_{-0.003}^{+0.003}$ & 0.03$_{-0.01}^{+0.01}$ & 2.2$_{-0.1}^{+0.1}$ & { 291$_{{{-13}}}^{{{+13}}}$ } \\\\ & XIS1 (D) & 1.13$_{-0.01}^{+0.01}$ & \\multicolumn{4}{c}{fitted together with XIS0,3} & 6.412$_{-0.004}^{+0.005}$ & 0.04$_{-0.01}^{+0.01}$ & 2.3$_{-0.1}^{+0.1}$ & { 280$_{{{-13}}}^{{{+13}}}$ } \\\\ & PCA PCU2 (D) & 1.16$_{-0.01}^{+0.01}$ & 13.6$_{-0.5}^{+0.5}$ & 0f & 0f & -- & 6.4f & 0f & 2.1$_{-0.3}^{+0.3}$ & { 303$_{{{-39}}}^{{{+39}}}$ } \\\\ \\hline \\end{tabular} \\end{center} \\end{minipage} \\end{table*} To model thermal Comptonization, we use the {\\sc compps} model \\citep{Poutanen1996} in {\\sc xspec}. It models Compton scattering in a plasma cloud of given temperature, $T_{\\rm e}$, and Thomson optical depth, $\\tau$, for a number of geometries and locations of the seed photon sources. Since $T_{\\rm e}$ and $\\tau$ are strongly intrinsically anticorrelated, we also use the Compton parameter (e.g., \\citealt{Rybicki1979}), $y\\equiv 4(kT_{\\rm e}/m_{\\rm e} c^2)\\tau$ (where $m_{\\rm e}$ is the electron mass) as the second fitting parameter instead of $\\tau$. The advantage of this choice is that $T_{\\rm e}$ and $y$ are almost orthogonal; $y$ determines closely the power-law slope at low energies whereas $kT_{\\rm e}$ determines the position of the high-energy cut-off. The seed photons are assumed to have a disc blackbody distribution ({\\sc diskbb} in {\\sc xspec}, \\citealt{Mitsuda84}) with the maximum temperature of $kT_{\\rm bb}=10$ eV. The model normalization, $K$, is that of the seed disc blackbody model as defined in {\\sc xspec}, see Section \\ref{xray}. Among the geometries included in {\\sc compps}, we consider here four spherical cases; an approximate treatment of radiative transfer using escape probability (which is denoted in {\\sc compps} by the geometry parameter $=0$), a sphere with central soft photons (geometry 4), homogeneously distributed seed photons (geometry $-4$) and seed photons distributed according to the diffusion-equation eigen-function, $\\propto \\sin(\\upi\\tau'/\\tau)/ (\\upi\\tau'/\\tau)$, where $0\\leq \\tau'\\leq\\tau$ ($-5$). Then, we consider a slab with the seed photons either at its bottom (geometry 1) or distributed homogeneously (geometry $-1$). Also, we consider the hot plasma in the shape of either a cylinder with the height equal to its radius, or a hemisphere, with the seed photons being either at its bottom (2, 3, respectively) or homogeneously distributed ($-2$, $-3$, respectively). The spectral data of NGC 4151 studied in this paper are of high quality and represent the largest set ever collected for this source. This gives us a possibility to test how precise information about X$\\gamma$ emission from brightest Seyferts can be achieved with the current satellites. Therefore, we extensively test various variants of the Comptonization model for our two best-quality states, bright and dim. This could, in principle, give us some indications regarding the actual geometry of the source. Our results are presented in Tables \\ref{comp_results}(a) and (b), regarding the parameters of the Comptonizing plasma and the strength of reflection (which are the same for all used detectors within a given state), and of the absorber and the Fe line (which are specific to a given X-ray detector within each of the states), respectively. The results in Table \\ref{comp_results}(b) and the entries for the bolometric model luminosity, $L_{\\rm X\\gamma}$, in Table \\ref{comp_results}(a) are given only for the case of spherical geometry calculated using escape probability formalism (geometry 0). We find the thermal Comptonization model to provide very good fits to the data. The values of the reduced $\\chi^2$ are $\\la 1$, which, however, appears not to be due to an overly complex model. In particular, the high-energy continua are determined only by three parameters, $y$, $R$, $kT_{\\rm e}$. For the bright state, Comptonization provides a much better fit than the e-folded power law, with $\\Delta \\chi^2$ of up to $-14$. However, any trends seen for the geometry of the Comptonizing plasma are rather weak. In the case of the bright state, equally good models can be obtained with either a slab or a sphere. Also, models with seed photons being distributed within the source are of similar quality as those with localized seed photons. On the other hand, the differences between the models are slightly stronger in the case of the dim state. In particular, the slab model with seed photons at its bottom is somewhat better than that with the seed photons distributed throughout it. This can be a hint that the actual source has the seed photons external to the source rather than internal (which would be the case for the dominant synchrotron seed photons). We also note that the fitted values of $kT_{\\rm e}$ are rather insensitive to the assumed geometry in both the bright and dim states. Our results for the bright state are similar to those of previous Comptonization fits, in particular to those of Z02, who obtained $kT_{\\rm e}= 73^{+34}_{-29}$ keV, $y=0.88^{+0.12}_{-0.11}$, and $R=0.60^{+0.24}_{-0.21}$ at $L_{\\rm X\\gamma}\\simeq 3.9\\times 10^{43}$ erg s$^{-1}$ using a broad-band X-ray spectrum from \\asca\\/ and OSSE. In all geometries, we see in Table \\ref{comp_results}(a) that the Compton $y$ parameter remains approximately constant between the bright and dim states, with only a slight decrease in the dim state. For the medium state, Table \\ref{comp_results}(a) gives the results only for the model with escape probabilities, for which the best fit $y$ is almost the same as that in the dim state. These results are illustrated by the confidence contours shown in Fig.\\ \\ref{contours}. \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_contours.eps} \\caption{The 90 per cent confidence regions for the Compton $y$ parameter (left) and the optical depth, $\\tau$ (right), vs.\\ the temperature, $kT_{\\rm e}$ for three geometries of the Comptonizing plasma. The solid blue, solid red and green dashed (only for $y$) contours correspond to the bright, dim and medium state, respectively. The dashed curves for $\\tau$ vs.\\ $kT_{\\rm e}$ correspond to $\\tau\\propto kT_{\\rm e}^{-1}$, normalized to the bright-state results.} \\label{contours} \\end{figure} Apart from the results given in Table \\ref{comp_results}, a reduction of $\\chi^{2}$ appears when the Fe abundance (assumed equal for the absorber and the reflector) is a free parameter. We found a moderate overabundance, ${1.5\\pm 0.2}$ times that of \\citet{Anders1982} for the bright state, with $\\Delta\\chi^{2} = -6$, and ${1.3\\pm 0.1}$ for the dim state, $\\Delta \\chi^{2} = -2$. For the \\xmm\\/ observation in 2000 December, an Fe overabundance by 2--3 was claimed by \\citet{Schurch2003} but that fit was done with the \\xmm\\/ spectrum only and the reflection strength (affecting that determination) was not well constrained. We also find that both the bright and dim state spectra prefer an inclination angle $>45\\degr$, with $\\Delta \\chi^{2} = -3$ and $-6$, respectively, at $75\\degr$. Since this effect is observed for symmetric (sphere) and asymmetric (hemisphere) geometries, we conclude that its main cause is the changing shape of the reflection component. On the other hand, allowing the reflector to be ionized does not improve the fit, and $\\xi < 0.2$ erg cm s$^{-1}$ (at an assumed reflector temperature of $10^5$ K). The lack of the far-UV and very soft X-ray spectra does not allow us to fit the maximum temperature of the seed disc blackbody photons, $kT_{\\rm bb}$. We have found that changing it to either 5 or 20 eV yields no improvement of the fit. This is indeed expected given that the photon energies emitted by the disc are much below the fitted energy range, at which the shape of the Comptonization spectrum has already achieved a power-law shape independent of $kT_{\\rm bb}$. Table \\ref{comp_results}(b) also shows the normalization of a given X-ray spectrum relative to that from ISGRI. For the bright state, when the \\integral, \\xmm\\/ and \\xte\\/ observations were almost contemporary, $C_{\\rm ISGRI}$ is very close to 1 and, for both of the \\xmm\\/ spectra, it follows the flux changes seen in the top panel of Fig.\\ \\ref{short}. This shows that ISGRI, EPIC pn and PCA detectors are well cross-calibrated. In the case of the two other spectral sets, the observations are not simultaneous. Nevertheless, $C_{\\rm ISGRI}$ varies only within the range of 0.74--1.17, confirming the validity of our selection of the spectra for a given flux state. \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_levels.eps} \\caption{The Comptonization model spectra (solid curves; the geometry parameter $=0$) of the three states of the NGC 4151, shown together with ISGRI spectra. The bright, medium and dim spectra are shown by the blue crosses, green circles and red triangles, respectively. The dashed curves show the reflection model components for the bright (upper) and dim state. The fitted JEM-X, \\xte, \\xmm\\/ and \\suzaku\\/ spectra are not shown for clarity.} \\label{models} \\end{figure} \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_highenergy.eps} \\caption{The Comptonization models for the extreme states (solid curves, same as in Fig.\\ \\ref{models}) compared to the average OSSE spectrum (crosses and two $2\\sigma$ upper limits). The rightmost upper limit is the COMPTEL upper limit in the 0.75--1 MeV band \\citep{Maisack1995}. The histograms give the \\integral/PICsIT $3\\sigma$ upper limits for 1 Ms (upper) and 10 Ms (lower) effective exposure time. The PICsIT 1.34 Ms average 277--461 keV flux is shown by error bars with a diamond. The SPI spectrum obtained by \\citet{Bouchet2008} is shown by error bars with circles. Note that the 200--600 keV data point may be spurious. } \\label{high} \\end{figure} We have also tested our results using the extended data sets (see Section \\ref{assumptions}). However, the quality of the fits and the plasma parameter uncertainty are not improved when they are used. This confirms that our selection of the primary sets was valid. Thus, we do not present here those results. We note that even for such a bright AGN as NGC 4151, \\textbf{we have no data at energies $\\ga 200$ keV, as shown in Fig.\\ \\ref{models}. Still, for the dim state, we have found $kT_{\\rm e}\\simeq 200$ keV with a statistical uncertainty of $\\sim 7$ per cent or so. The physical reason for this low uncertainty is the observed lack of spectral curvature. In the process of thermal Comptonization, a power law slope at $E\\la kT_{\\rm e}$ appears due to the merging of many individual up-scattering profiles, each being curved. The cut-off at $E\\ga kT_{\\rm e}$ is due to the electron energies being comparable to the photon ones, in which case a photon no longer increases its energy in a scattering. Since we see no cut-off up to $\\sim$200 keV, models with a lower $kT_{\\rm e}$ are ruled out. On the other hand, models with a higher $kT_{\\rm e}$ correspond to an increase of the photon energy in a single scattering by a factor so large that the individual, curved, scattering profiles become clearly visible and the spectrum at $E\\la kT_{\\rm e}$ is no longer a power law. Since the observed spectral shape {\\it is\\/} a power law, this case is also excluded. We note, however, that our assumed model corresponds to a single scattering zone. A power law spectrum at $E\\la 200$ keV may also be produced by a superposition of emission regions with different temperatures (smoothing out the resulting spectrum), as, e.g., in a hot accretion flow. In this case, the maximum temperature of the flow could be higher than 200 keV. In any case, we do see a very clear difference in the characteristic Comptonization temperature between the bright and dim state.} We comment here on the very high value of $kT_{\\rm e}=315\\pm 15$ keV obtained by P01 for the 1999 January \\sax\\/ spectrum of NGC 4151. The fit of that model appears rather poor when compared with the curvature of the \\sax/PDS data above 50 keV, see fig.\\ 2(b) in P01. We also find that the PDS spectrum is very similar to the average OSSE spectrum \\citep{Johnson1997}, shown in Fig.\\ \\ref{high} (with only the normalization of the PDS spectrum being about 5 per cent lower). This indicates that the source was in a moderately bright state in 1999 January, for which state the temperature of 315 keV appears highly unlikely. We cannot explain this discrepancy; it may be that $kT_{\\rm e}=315\\pm 15$ keV represented a spurious local minimum. As we see in Fig.\\ \\ref{high}, our bright-state model spectrum agrees well in shape with the average OSSE spectrum \\citep{Johnson1997}. The latter also agrees with the first three points of the average NGC 4151 spectrum from \\integral/SPI \\citep{Bouchet2008}. However, the SPI 200--600 keV flux is implausibly high, which may reflect a problem with the background estimate. Indeed, the SPI analysis of \\citet{Petry2009} yields only upper limits at $E>150$ keV. We see in Fig.\\ \\ref{high} that the Comptonization models of the bright and dim states cross at $\\simeq 300$ keV. Unfortunately, it is impossible to verify it with our data. In particular, we cannot test if there is any non-thermal component at high energy. Even the average spectrum from OSSE, the most sensitive detector up to date at $E\\ga 150$ keV, gives only a weak detection of NGC 4151 at $E\\ga 250$ keV with an $>4$ Ms exposure time. From the shown upper limits (including systematic errors) of the \\integral/PICsIT, the current detector most sensitive at $E\\ga 250$ keV \\citep{Lubinski2009}, we see it would need $>10$ Ms exposure time to detect NGC 4151 in the 300--400 keV range. Merging all the suitable \\integral\\/ data (with the effective exposure time of 1.34 Ms taken without the bright state, for which the observation was in the staring mode, not suitable for the PICsIT), we extracted the PICsIT flux in the 277--461 keV band. Although it is not a detection (and the shown error bar does not include systematic errors), it provides a hint that the average NGC 4151 emission in this energy range is weak. \\section[]{Discussion} \\subsection{Reflection and Fe line} \\label{reflector} \\subsubsection{Comparison with previous work} \\label{comparison} A reflection component in the X-ray spectra of NGC 4151 was first found by \\citet{Zdziarski1996a}, using contemporaneous data from \\gro\\/ and \\ginga, yielding $R\\simeq 0.43$ (using a Comptonization model and assuming $i = 65\\degr$). Then, Z02 found $R\\simeq 0.6$ (at $i=17\\degr$) using \\gro\\/ and \\asca\\/ data, and P01 found $R\\simeq 0.2$--1.8 using \\sax\\/ data and a reflection model averaged over the viewing angle. A previous analysis of \\integral\\/ bright-state data from 2003 May yielded a larger reflection, $R = 0.72\\pm 0.14$ \\citep{Beckmann2005}, than that determined by us (their value of $kT_{\\rm e} = 94_{-10}^{+4}$ keV is also different from that found by us for the same data). By fitting the ISGRI and JEM-X bright-state spectra alone, we find results very similar to those with the \\xmme and \\xtee spectra included (Table \\ref{comp_results}). Therefore the differences appear to be a consequence of the substantial change of the \\integral\\/ calibration since 2005. Furthermore, \\citet{Beckmann2005} used the SPI spectra obtained in the staring observation mode, which can be affected by a large background uncertainty. Then, \\citet{DeRosa2007} found $R\\simeq 0.4$ for several \\sax\\/ spectra and $R\\simeq 0$ for the remaining ones, but using an e-folded power-law model rather than Comptonization. Then, \\citet{Schurch2002} found $R\\simeq 1.9$ using \\xmm\\/ data assuming $i=65\\degr$, but this result was obtained using the 2.5--12 keV band only and assuming a fixed $\\Gamma = 1.65$. Regarding Fe K$\\alpha$ line, \\citet{Zdziarski1996a} found a narrow line with $I_{\\rm Fe}\\simeq (0.6$--$1.0)\\times 10^{-4}$ cm$^{-2}$ s$^{-1}$ using \\ginga\\/ data, and Z02 found a narrow line at $I_{\\rm Fe}\\simeq 2.5\\times 10^{-4}$ cm$^{-2}$ s$^{-1}$ accompanied by a weaker broad relativistic line using \\asca\\/ data. The broad component has not been observed later, whereas the narrow line flux measured by \\sax\\/ \\citep{DeRosa2007}, \\chandra\\/ \\citep{Ogle2000} and \\xmm\\/ \\citep{Schurch2003} was found to be in the range (1.3--$4.2)\\times 10^{-4}$ cm$^{-2}$ s$^{-1}$. As seen in Tables \\ref{comp_results}(a--b), our results on $R$ and $I_{\\rm Fe}$ approximately agree with the previous measurements. \\subsubsection{Close and distant reflector} \\label{close_distant} We see in Table \\ref{comp_results}(a) that the relative strength of reflection, $R$, is close to being twice higher in the dim state than in the bright one. (In the medium state, the amount of data is very limited, and the apparent lack of detectable reflection may be not typical to that flux level.) The increase of $R$ with decreasing flux could be a real change of the solid angle subtended by the reflector between the states. On the other hand, it could be that it is due to the contribution of a distant reflector, in particular a molecular torus, as postulated in the AGN unified model. Reflection from distant media is indeed commonly seen in Seyferts 2, e.g., \\citet{Reynolds1994}. If we attribute the increase of $R$ in the dim state due to that component, we can calculate the fractional reflection strength from the close reflector, presumably an accretion disc, and the flux reflected from the torus, which, due to its large size, is averaged over a long time scale, and thus assumed to be the same in either state. The hypothesis of the stable geometry of the source is supported by the approximate constancy of the Compton $y$ parameter, see Table \\ref{comp_results}(a) and Fig.\\ \\ref{contours}, and of the photon index, $\\Gamma$, see Table \\ref{pl_fits}. Those parameters are closely related to the amplification factor of Comptonization (e.g., \\citealt{Beloborodov1999}), whose constancy can be most readily achieved if the system geometry is constant (e.g., \\citealt{Haardt1991,Zdziarski1999}). Thus, we consider an approximately constant relative reflection strength from the disc very likely. Observationally, we do see a strong $R$-$\\Gamma$ correlation in Seyferts and X-ray binaries, which can be interpreted by a geometrical feedback model \\citep{Zdziarski1999}, and which would require a constant disc $R$ for a given $\\Gamma$. The unabsorbed flux in a given state, $F_{\\rm S}$, where ${\\rm S}=$ either D (dim) or B (bright), is the sum of the (unabsorbed) incident (i) and reflected (r) fluxes, \\begin{equation} F_{\\rm S}=F_{\\rm i,S}+F_{\\rm r,S}, \\label{decomposition} \\end{equation} where $F_{\\rm S}$ and $F_{\\rm r,S}$ can be calculated from the results of our spectral fitting in Section \\ref{Compton}. We assume that the observed (unabsorbed) reflected flux in each state is the sum of the close disc (d) state-dependent component and the constant distant torus (t) component, \\begin{equation} F_{\\rm r,S} = F_{\\rm d,S} + F_{\\rm t}. \\label{sum} \\end{equation} Then, assuming the constancy of the solid angle subtended by the disc reflector, or, equivalently, its reflection strength, $R_{\\rm d}$, the disc-reflected fluxes are, \\begin{equation} F_{\\rm d,S} =F_{\\rm r,S} (R_{\\rm d}/ R_{\\rm S}), \\label{disc_fluxes} \\end{equation} where $R_{\\rm S}$ is the reflection strength for a given state. We have four equations (\\ref{sum}--\\ref{disc_fluxes}) for four unknowns, $F_{\\rm d,B}$, $F_{\\rm d,D}$, $R_{\\rm d}$ and $F_{\\rm t}$, which can be readily solved. We note that they can be formulated without involving $F_{\\rm i,S}$, which implies that the solution does not depend on the choice of the energy interval in which the fluxes are measured as long as it includes the entire reflected spectrum. The solutions for $R_d$ and $F_{\\rm t}$ are, \\begin{equation} R_{\\rm d} = \\frac{(F_{\\rm r,B}-F_{\\rm r,D})R_{\\rm B}R_{\\rm D}}{F_{\\rm r,B} R_{\\rm D}-F_{\\rm r,D} R_{\\rm B}}, \\qquad F_{\\rm t} = \\frac{(R_{\\rm D}-R_{\\rm B})F_{\\rm r,B}F_{\\rm r,D} } {F_{\\rm r,B} R_{\\rm D}-F_{\\rm r,D} R_{\\rm B}}. \\label{solution} \\end{equation} Then, the fractional reflection strength of the torus can be approximately estimated as, \\begin{equation} R_{\\rm t}\\simeq \\langle R\\rangle (F_{\\rm t}/ \\langle F_{\\rm r}\\rangle), \\label{torus} \\end{equation} where $\\langle R\\rangle$ and $\\langle F_{\\rm r}\\rangle$ are the average observed reflection strength and the average reflected flux, respectively. Using the values obtained for thermal Comptonization using escape probability formalism (geometry parameter $=0$), we obtain $R_{\\rm d} = 0.27\\pm 0.07$, $F_{\\rm t}$ = { (8.2$\\pm$0.6)}$\\times 10^{-11}$ erg cm$^{-2}$ s$^{-1}$, and $R_{\\rm t}$ = { 0.24$\\pm$0.04} using either arithmetic or geometric averages in equation (\\ref{torus}). Similar values are obtained for other Comptonization geometries considered in Section \\ref{Compton}. We can check the robustness of our estimates by allowing the reflection strength to depend on the accretion rate, or flux. If we assume that $R_{\\rm d}$ in the dim state is a half of that in the bright state [which requires an appropriate change of equation (\\ref{disc_fluxes})], we obtain relatively similar values of the disc reflection in the bright state of $R_{\\rm d,B}$ = { 0.23$\\pm$0.12}, $F_{\\rm t}$ = { (1.1$\\pm$0.1)}$\\times 10^{-10}$ erg cm$^{-2}$ s$^{-1}$, and $R_{\\rm t}$ = { 0.32$\\pm$0.04}. Similar reasoning can be applied to the Fe K$\\alpha$ line. The observed line photon flux in the state S, $I_{\\rm S}$, is assumed to be the sum of contributions from the disc, $I_{\\rm d,S}$ and the torus, $I_{\\rm t}$. The latter is assumed constant, and the former proportional to the ionizing flux incident on the disc. As a simplification, we assume the ionizing flux ($\\geq 7.1$ keV for neutral Fe) to be proportional to the differential continuum photon flux, $F(E)/E$, at the line centroid energy, which we denote $N_{\\rm S}$, and which is equal to $I_{\\rm S}/w_{\\rm S}$, where $w_{\\rm S}$ is the observed equivalent width in a given state. This then implies a constant disc line equivalent width, $w_{\\rm d}$. Above, for the sake of simplicity, we have dropped the indices 'Fe', used in Section \\ref{Compton} and Table \\ref{comp_results}(b). The equations are, \\begin{equation} I_{\\rm S} = I_{\\rm d,S}+I_{\\rm t}, \\qquad I_{\\rm d,S} = w_{\\rm d} N_{\\rm S}, \\label{line} \\end{equation} which can be solved for \\begin{equation} w_{\\rm d} = {I_{\\rm B}-I_{\\rm D}\\over N_{\\rm B}-N_{\\rm D}}, \\qquad I_{\\rm t} = {I_{\\rm D} N_{\\rm B}-I_{\\rm B}N_{\\rm D}\\over N_{\\rm B}-N_{\\rm D}}. \\label{line_solution} \\end{equation} The equivalent width of the torus line flux with respect to the average incident photon flux, $ \\langle N\\rangle$, can be estimated as, \\begin{equation} w_{\\rm t}={I_{\\rm t}/ \\langle N\\rangle}. \\label{wt} \\end{equation} For the numerical values, we use here averages of the fit results for both EPIC pn spectra in the bright state, and for the EPIC pn and XIS spectra in the dim state, see Table \\ref{comp_results}(b). We find $I_{\\rm B}$ = { (2.9$\\pm$0.2)} $\\times 10^{-4}$ cm$^{-2}$ s$^{-1}$, $I_{\\rm D}$ = { (2.3$\\pm$0.1)}$\\times 10^{-4}$ cm$^{-2}$ s$^{-1}$, $N_{\\rm B}$ = { (3.6$\\pm$0.1)}$\\times 10^{-3}$ keV$^{-1}$ cm$^{-2}$ s$^{-1}$ and $N_{\\rm D}$ = { (8.1$\\pm$0.1)}$\\times 10^{-4}$ keV$^{-1}$ cm$^{-2}$ s$^{-1}$. This yields $w_{\\rm d}$ = { 23$\\pm$11} eV, $I_{\\rm t}$ = { (2.1$\\pm$0.9)}$\\times 10^{-4}$ cm$^{-2}$ s$^{-1}$, and $w_{\\rm t}$ = { 93$\\pm$41} eV, { 120$\\pm$53} eV using the arithmetic or geometric average, respectively. The equivalent width of the Fe line at $i=45\\degr$ by an isotropically illuminated cold disc for $\\Gamma = 1.75$ is $\\simeq 140$ eV \\citep{George1991}. The value of $R_{\\rm d}\\simeq 0.27$ found based on the observed reflection strength then implies $w_{\\rm d}\\sim 38$ eV, somewhat higher than the estimate based on the line fluxes, but still in an approximate agreement taking into account measurement errors and a number of assumptions we have made. On the other hand, $R_{\\rm t}\\simeq 0.24$ found above would explain only $w_{\\rm t}\\sim 30$ eV. However, we note that we have neglected the local absorber, which obviously also gives rise to an Fe K$\\alpha$ line component. Its characteristic $N_{\\rm H}$ of $\\sim 10^{23}$ cm$^{-2}$ (Table \\ref{comp_results}b) can readily explain the excess equivalent width of $\\sim 70$ eV (e.g., \\citealt{Makishima1986,Awaki1991}) of the constant line component with respect to that expected from the torus. We note a number of caveats for our results. The standard accretion disc is flared, not flat \\citep{Shakura1973}. This, however, would have a relatively minor effect, changing somehow the distribution of the inclination angles. Our results indicate disc reflection with $R_{\\rm d}$ substantially less than unity, which implies the X$\\gamma$ source is not entirely above the disc. A likely geometry explaining it is a hot inner flow surrounded by a disc (e.g. \\citealt{Abramowicz1995,Narayan1995,Yuan2001}), in which case the incident radiation will have much larger incident angles (measured with respect to the axis of symmetry) than those assumed in the used model \\citep{Magdziarz1995}. The disc may be warped (e.g., \\citealt{Wijers1999}), which again would change the distribution of the incident angles. Furthermore, Compton scattering in the hot plasma above the disc reduces the observed reflection strength \\citep{Petrucci2001b}. Then, we also used the slab geometry for the torus reflection. Thus, the obtained solid angle does not correspond to the actual angle subtended by the torus from the X-ray source. For example, \\citet{Murphy2009} considered a torus with a circular cross section subtending (as seen from the centre) a $2\\upi$ solid angle. They found that the reflection component in this case is several times weaker than that corresponding to a slab subtending the same angle. One obvious effect here is that, unlike the case of a slab, an observer sees only a fraction of the reflecting surface, without parts obscured by the torus itself. If we take this into account, our value of $R_{\\rm t}\\sim 0.2$ appears consistent with a torus subtending a $\\sim 2\\upi$ solid angle. We note, however, that even if the torus solid angle is formally $2\\upi$ or so, a large part of it will be shielded from the X-ray source by the accretion disc and the black hole itself, so the actual irradiated solid angle may be substantially lower. This effect was not taken into account in the geometrical model of \\citet{Murphy2009}. Furthermore, the cross section of the torus may be substantially different from circular, which may increase the observed reflection, as well as it may be clumpy \\citep*{Krolik1988,Nenkova2002}, which will decrease it. Still, our value of $w_{\\rm t}\\sim 100$ eV agrees with that from a torus with the column density of $N_{\\rm H}\\sim 10^{24}$ cm$^{-2}$ for $\\Gamma\\simeq 1.75$ \\citep{Murphy2009}, which provides an explanation for that equivalent width alternative, or additional, to that as being due to the line emission of the absorber (discussed above). We have assumed that the torus reflection component is constant on the time scale of years. On the other hand, \\citet{Minezaki2004} found that the dusty torus inner boundary is at $\\simeq 0.04$ pc. Thus, a fraction of the torus reflection may vary on the corresponding time scale of $\\sim 50$ d. Furthermore, 0.04 pc corresponds to $\\simeq 2\\times 10^4 R_{\\rm g}$ (where $R_{\\rm g}\\equiv GM/c^2$ is the gravitational radius), where the accretion disc may still be present and join onto the torus. On the other hand, \\citet{Radomski2003} have constrained the torus outer boundary to $\\la 35$ pc, so the bulk of the torus reflection may still be constant over a time scale of years. \\subsection{Absorber properties} \\label{absorber} We compare our results with those based on \\sax, which also provided broad-band spectra, allowing to simultaneously determine absorber properties and the continuum. They were studied by \\citet{Puccetti2007} and \\citet{DeRosa2007}, who used the same absorber model as in this work. We find their results to be compatible with ours. The fully covering absorber has $N_{\\rm H} \\simeq (0.9$--$9.4)\\times 10^{22}$ cm$^{-2}$ for \\sax\\/ and $\\simeq (3.0$--$8.1)\\times 10^{22}$ cm$^{-2}$ in our case, see Table \\ref{comp_results}(b). For the partially covering absorber, $N_{\\rm H}^{\\rm p}\\simeq (3.5$--$30.3)\\times 10^{22}$ cm$^{-2}$ (\\sax) and $\\simeq (9.5$--$30.2)\\times 10^{22}$ cm$^{-2}$ (this work). The covering fractions are also similar, $f^{\\rm p}\\simeq 0.34$--0.71 (\\sax) and 0.36--0.71 (this work). The only exception is the long-exposure \\sax\\/ observation of NGC 4151 in December 2001 showing a very low $N_{\\rm H}$ \\citep{DeRosa2007}. We find an anticorrelation between the $N_{\\rm H}$ of both absorber components and the hard X-ray flux. We define the total column, $N_{\\rm H}=N_{\\rm H}^{i}+ f^{p} N_{\\rm H}^{p}$. For the bright state, we have $N_{\\rm H}/10^{22}$ cm$^{-2}$ = { 11.1$\\pm$3.5} (\\xmm\\/ X1+X2) and { 8.2$\\pm$4.9} (X3), for the medium state, { 18.0$\\pm$2.8} (\\xmm\\/ X5), and for the dim state, { 23.1$\\pm$0.8} (\\xmm\\/ X4) and { 23.2$\\pm$0.9} (\\suzaku). Also, the covering fraction is anti-correlated with the hard X-ray flux, see Table \\ref{comp_results}(b). A similar trend is seen for the results of \\citet{Puccetti2007}, who used the 6--10 keV for the X-ray flux. The variable part of the absorber needs to be relatively close to the X-ray source to be able to follow its flux on the time scale of days. The absorber in NGC 4151 was identified with massive outflows from the accretion disc \\citep{Piro2005}, a broad-line region \\citep{Puccetti2007} or the surface or wind of the torus \\citep{Schurch2002}. Only the disc wind is close enough to the X-ray source to follow the change of the nuclear emission and to produce any correlation. The physical explanation for the correlation remains unclear; possibly the wind rate in NGC 4151 decreases with increasing accretion rate. \\subsection{The nature of the X-ray source} \\label{xray} The geometry and parameters of the source are constrained by a number of our findings. We find (i) that both the Compton parameter and the X-ray spectral index are approximately the same in both bright and and dim states (Sections \\ref{cutoff}--\\ref{Compton}). This implies an approximately constant amplification ratio of the Comptonization process, i.e., the ratio of the power emitted by the plasma to that supplied to it by seed photons (e.g., \\citealt{Beloborodov1999}). The amplification factors obtained with the Comptonization model are indeed similar, $A$ = { 17$\\pm$2, 15$\\pm$2} for the bright and dim state, respectively (assuming $kT_{\\rm bb}=10$ eV). A similar value of $A\\simeq 13$ was obtained for a Comptonization model fitted to the 1991 data from \\rosat, \\ginga\\/ and OSSE \\citep{Zdziarski1996a}. If the seed photons are supplied by an accretion disc surrounding the hot plasma, this implies an approximately constant disc inner radius. Based on this, we have inferred (ii) a relatively weak reflection from the disc, $R_{\\rm d}\\simeq 0.3$ (Section \\ref{close_distant}). This qualitatively agrees with $A\\gg 1$ and both findings rule out the static disc corona geometry; the (outer) disc subtends a small solid angle as seen from the X-ray source, and the X-ray source subtends a small solid angle as seen from inner parts of the disc \\citep{Zdziarski1999}. An implication of the latter is that the modelled disc blackbody emission (which provides seed photons for Comptonization) should be much weaker than that observed (which corresponds to the entire disc emission). This seems to be indeed the case as shown in Fig.\\ \\ref{sed}, where we see that the bright-state UV flux inferred from the Comptonization model is about an order of magnitude below the shown maximum observed far UV (1350 \\AA; 9.2 eV) flux. On the other hand, the model dim-state far UV flux is close to the historical minimum observed. However, that minimum represented a single isolated dip in the light curve \\citep{Kraemer2006}, and the actual far UV flux corresponding to the dim state is likely to be significantly higher. In choosing the shown range of the far UV flux we used its strong correlation with the optical flux, which can be seen by comparing Fig.\\ \\ref{longest}(d) with fig.\\ 1 in \\citet{Kraemer2006}. The IR, optical and UV fluxes shown in Fig. \\ref{sed} appear relatively weakly affected by the host galaxy emission. In particular, the dominance of the AGN in the U band is shown by its strong variability \\citep{Czerny2003}. In the optical range, there can be some non-negligible fraction of emission from the broad and narrow line regions, but in the UV we expect that the disc dominates. Given that the measured UV emission is, furthermore, absorbed by the host galaxy, the conclusion above that only a small fraction of the disc emission undergoes Comptonization appears secure. \\begin{figure} \\includegraphics[width=\\columnwidth]{ngc4151_sedis5eV.eps} \\caption{The broad-band spectrum of NGC 4151. The radio, IR, optical and UV data (black dots) were taken from the NED database. Additional data are shown in circles in radio (green, \\citealt{Ulvestad2005}), mid and near IR (brown, \\citealt{Radomski2003,Ruiz2003}, optical (magenta, OMC, this work) and UV (purple, \\swift/UVOT, this work; and cyan, \\hst/STIS, the flux extrema from fig.\\ 3 of \\citealt{Kraemer2006}). The \\xmm\\/ EPIC pn and ISGRI spectra of the bright and dim states are shown by the blue and red dots, respectively, the corresponding unabsorbed Comptonization models below 10 keV are shown by dashes, and the disc blackbody (assuming geometry parameter $=0$ and $kT_{\\rm bb}=5$ eV) incident on the hot plasma are shown by the dotted curves. The upper limits in the 0.75--1--3 MeV energy ranges are from COMPTEL \\citep{Maisack1995}.} \\label{sed} \\end{figure} Another result is (iii) the connection of the normalization of the fitted disc blackbody seed emission to the disc inner radius, $R_{\\rm in}$. Based on the definition from {\\sc xspec}, we can express it as \\begin{equation} K={10^8 r\\cos i\\over f_{\\rm col}^4} {(R_{\\rm in}/10^{12}\\,{\\rm cm})^2\\over (D/10\\,{\\rm Mpc})^2}, \\label{norm} \\end{equation} where $f_{\\rm col}\\sim 1.7$ is the colour correction to the blackbody temperature \\citep{Shimura1995}, and $r$ is the ratio of the blackbody emission incident on the plasma to that emitted by the disc. It implies (at $D=13.2$ Mpc and $i=45\\degr$), \\begin{equation} R_{\\rm in}\\simeq 4.5\\times 10^{13} \\left(K \\over 10^{9}\\right)^{1/2} \\left( r\\over 0.1\\right)^{-1/2} \\left(f_{\\rm col}\\over 1.7\\right)^2\\, {\\rm cm}. \\end{equation} At $M=4.6\\times 10^7\\msun$ \\citep{Bentz2006}, $R_{\\rm g}\\simeq 6.8\\times 10^{12}$ cm. Thus, the values of $K$ in Table \\ref{comp_results}(a) for the bright state (at $T_{\\rm bb}=10$ eV) imply $R_{\\rm in}\\sim 6 R_{\\rm g}$, the last stable orbit for a non-rotating black hole. This is in conflict with the findings above, according to which the inner part of the flow is occupied by the hot plasma and not by the disc, unless the black hole is rotating fast. Furthermore, the {\\sc diskbb} model used for the seed photons becomes invalid when $R_{\\rm in}$ is close to the last stable orbit with the standard zero-stress boundary condition \\citep{Gierlinski1999}. We note, however, the disc flux is $\\propto R_{\\rm in}^2 T_{\\rm bb}^4$, and approximately $R_{\\rm in}\\propto T_{\\rm bb}^{-2}$ for a given X$\\gamma$ spectrum. In particular, $K=7.2\\times 10^9$ for the bright state at $kT_{\\rm bb}=5$ eV (consistent with the UV data) and the geometry parameter $=0$, confirming the above scaling. This yields $R_{\\rm in} \\simeq 18 R_{\\rm g}$, allowing the existence of a hot inner flow. Given that agreement, we used $kT_{\\rm bb}=5$ eV in the models shown in Fig.\\ \\ref{sed}. We note that if $R_{\\rm in}\\simeq$ constant, $kT_{\\rm bb}$ has to be somewhat lower in the dim state. For comparison, the radius, $R_{\\rm eq}$, at which the integrated gravitational energy dissipation at $R>R_{\\rm eq}$ equals that at $R$10$^{7}$ cm$^{-3}$). To account for the broad-line emission in these four objects, \\citet{I07} considered various physical mechanisms such as Wolf-Rayet (WR) stars, stellar winds from Ofp or luminous blue variable stars, single or multiple supernova (SN) remnants propagating in the interstellar medium, and SN bubbles. While these mechanisms may be able to produce $L_{br}$ $\\sim$ 10$^{36}$ to 10$^{40}$ erg s$^{-1}$, they cannot generate yet higher luminosities, which are more likely associated with SN shocks or AGNs. \\citet{I07} considered type IIn SNe because their H$\\alpha$ luminosities are higher ($\\sim$10$^{38}$--10$^{41}$ erg~s$^{-1}$) than those of the other SN types and they decrease less rapidly. \\citet{IT08} found no significant temporal evolution of broad H$\\alpha$ in all four galaxies over a period of 3--7 years. Therefore, the IIn SNe mechanism may be excluded, leaving only the AGN mechanism capable of accounting for the high luminosity of the broad H$\\alpha$ emission. However, we also have difficulty with this mechanism. In particular, all four galaxies are present in neither the ROSAT catalogue of the X-ray sources nor the NVSS catalogue of radio sources. High-ionisation emission lines such as He {\\sc ii} $\\lambda$4686 or [Ne {\\sc v}] $\\lambda$3426 are weak or not detected in optical spectra. Based on the observational evidence, \\citet{IT08} concluded that all four studied galaxies most likely belong to the very rare type of low-metallicity AGNs in which non-thermal ionising radiation is strongly diluted by the radiation of a young massive stellar population. The fifth galaxy of this type, Tol 2240--384, was first spectroscopically studied by \\citet{T91} and \\citet{M94}. However, in those low-resolution spectra, some important emission lines, such as [O~{\\sc ii}] $\\lambda$3727, [Ne~{\\sc iii}] $\\lambda$3868, and H$\\alpha$ $\\lambda$6563, are missing. This, in particular, precludes abundance determination and the detection of broad hydrogen emission. \\citet{K04,K06} studied Tol 2240--384 spectroscopically, and \\citet{K06} derived the oxygen abundance of this galaxy, 12+logO/H = 7.77 $\\pm$ 0.08. We note that no broad emission was reported by \\citet{T91}, \\citet{M94}, and \\citet{K06}. In this paper, we present 8.2m Very Large Telescope (VLT) spectroscopic observations and 3.5 ESO New Technology Telescope (NTT) photometric observations of this emission-line galaxy. Its optical spectrum shows the very broad components of hydrogen emission lines and is similar to those found previously by \\citet{I07} and \\citet{IT08} for the four other galaxies. We describe observations in Sect.2. The morphology of the galaxy is discussed in Sect.3 and its location in the emission-line diagnostic diagram is discussed in Sect.4. Element abundances are derived in Sect.5. The kinematics of the ionised gas from narrow emission lines is discussed in Sect.6. We discuss in Sect.7 the properties of the broad emission and derive the mass of the central black hole assuming an AGN mechanism for the origin of the broad line emission. Our conclusions are summarized in Sect.8. \\begin{figure*}[t] \\hspace*{0.5cm}\\psfig{figure=14390f1.ps,angle=-90,width=16.cm,clip=} \\caption{Flux-calibrated VLT/UVES spectrum of Tol 2240--384, obtained on 23 August 2009, corrected for the redshift of $z$ = 0.07595 [ESO program 383.B-0271(A)] (upper spectrum). The lower spectrum is the upper spectrum downscaled by a factor of 100. The scale of the ordinate is that for the upper spectrum. Note the broad emission in the hydrogen line H$\\alpha$ $\\lambda$6563. No appreciable broad emission is detected in other strong permitted and forbidden lines, which is indicative of the rapid motions of relatively dense ionised gas with an electron number density $N_e$ $\\geq$ 10$^7$ cm$^{-3}$.} \\label{fig1} \\end{figure*} \\begin{figure*}[t] \\hspace*{1.0cm}\\psfig{figure=14390f2a.ps,angle=-90,width=7.5cm}% \\hspace*{1.0cm}\\psfig{figure=14390f2b.ps,angle=-90,width=7.5cm,clip=} \\caption{Flux-calibrated and redshift-corrected archival VLT/FORS1 medium-resolution (left) and low-resolution (right) spectra of Tol 2240--384 obtained on 12 September 2002 [ESO program 69.C-0203(A)] (upper spectra). The lower spectra are the upper spectra downscaled by a factor of 100. The scale of the ordinate is that for the upper spectra. Note the strong broad emission in the hydrogen line H$\\alpha$ $\\lambda$6563 and much weaker broad emission in the hydrogen line H$\\beta$ $\\lambda$4861. No appreciable broad emission is detected in strong forbidden lines.} \\label{fig2} \\end{figure*} ", "conclusions": " 1. Image deconvolution reveals two high-surface brightness regions in Tol~2240--384 separated by 2.4 kpc and differing in their luminosity by a factor of $\\sim$10. The brightest southwestern region is surrounded by intense ionised gas emission, which strongly affects the observed $B-R$ colour on a spatial scale of $\\sim$5 kpc. This high-excitation H~{\\sc ii} region is associated with broad H$\\alpha$ and H$\\beta$ emission. Surface photometry does not indicate, in agreement with the results of image deconvolution and unsharp masking, the presence of a bulge in Tol~2240--384. 2. We derived the oxygen abundance 12+logO/H = 7.85$\\pm$0.01 in Tol 2240--384, which is consistent within the errors with the value of 7.77$\\pm$0.08 derived earlier by \\citet{K06}. 3. The emission line profiles in the high resolution UVES spectrum reveal the presence of two narrow components with a radial velocity difference of $\\sim$ 78 km s$^{-1}$. Furthermore, the full widths at half maximum (FWHMs) of the narrow lines differ. Strong forbidden nebular lines [O~{\\sc iii}] $\\lambda$4959, 5007, [O~{\\sc ii}] $\\lambda$3726, 3729, and [Ne~{\\sc iii}] $\\lambda$3868 have FWHMs of 73 -- 80 km s$^{-1}$. The FWHMs of hydrogen lines are larger, $\\sim$ 85 km s$^{-1}$ and decrease from H$\\alpha$ to H$\\delta$ emission lines. The largest FWHM of $\\sim$ 92 km s$^{-1}$ is found for the He {\\sc i} $\\lambda$5876 emission line. This data suggest that narrow permitted hydrogen and helium lines probe the denser inner parts of the emitting regions compared to the forbidden lines. 4. Both UVES and FORS1 spectra reveal the presence of very broad hydrogen lines with FWHMs greater than 2000 km s$^{-1}$. The steep Balmer decrement of the broad hydrogen lines and the very high luminosity of the broad H$\\alpha$ line 3$\\times$10$^{41}$ erg s$^{-1}$ suggest that the broad emission arises from very dense and high luminosity regions such as those associated with supernovae of type IIn or with accretion discs around black holes. However, the presence of the broad H$\\alpha$ emission over a period of 7 years rules out the SN mechanism. Thus, the emission of broad hydrogen lines in Tol 2240--384 is most likely associated with an accretion disc around a black hole. 5. There is no obvious spectroscopic evidence of a source of non-thermal hard ionising radiation in Tol 2240--384. However, none is expected if, as we argue, the density of the broad line region is 5$\\times$10$^{6}$ -- 5$\\times$10$^{8}$ cm$^{-3}$. 6. Assuming that the broad emission in Tol 2240--384 is powered by an AGN, we have estimated a mass for the central black hole of $M_{\\rm BH}$ $\\sim$ 10$^7$ $M_\\odot$." }, "1005/1005.4699_arXiv.txt": { "abstract": "We present an analysis of an 8 arcminute diameter map of the area around the galaxy cluster Abell 1835 from jiggle map observations at a wavelength of 1.1 mm using the Bolometric Camera (Bolocam) mounted on the Caltech Submillimeter Observatory (CSO). The data is well described by a model including an extended Sunyaev-Zel'dovich (SZ) signal from the cluster gas plus emission from two bright background submm galaxies magnified by the gravitational lensing of the cluster. The best-fit values for the central Compton value for the cluster and the fluxes of the two main point sources in the field: SMM J140104+0252, and SMM J14009+0252 are found to be $y_{0}=(4.34\\pm0.52\\pm0.69)\\times10^{-4}$, 6.5$\\pm{2.0}\\pm0.7$ mJy and 11.3$\\pm{1.9}\\pm1.1$ mJy, where the first error represents the statistical measurement error and the second error represents the estimated systematic error in the result. This measurement assumes the presence of dust emission from the cluster's central cD galaxy of $1.8\\pm0.5$~mJy, based on higher frequency observations of Abell 1835. The cluster image represents one of the highest-significance SZ detections of a cluster in the positive region of the thermal SZ spectrum to date. The inferred central intensity is compared to other SZ measurements of Abell 1835 and this collection of results is used to obtain values for $y_{0} = (3.60\\pm0.24)\\times10^{-4}$ and the cluster peculiar velocity $v_{z} = -226\\pm275$~km/s. ", "introduction": "The Sunyaev Zel'dovich (SZ) effect \\citep{Sunyaev70} is the redistribution of energy in the Cosmic Microwave Background (CMB) spectrum due to interations between CMB photons and hot electrons along the line of sight between the surface of last scattering and an observer. The main source for the SZ effect is from the hot gas that exists in the intra-cluster medium (ICM) of massive galaxy clusters \\citep{Birkinshaw99, Carlstrom02, Rephaeli06}. The SZ actually comprises two effects. The thermal effect consists of a dimming or decrement in the apparent brightness of the CMB towards a galaxy cluster at low frequencies and a corresponding brightening or increment at high frequencies with the null crossover point at approximately 215 GHz. The kinematic effect has the spectral dependence of a standard temperature shift in the CMB which has the same sign at all frequencies. The two effects can therefore be distinguished from each other with measurements at multiple frequencies. Measurements of the amplitude of the SZ thermal distortion towards a cluster can be combined with measurements of the X-ray emission to determine the angular diameter distance, $d_{A}$, whose value depends on cosmology \\citep{Birkinshaw91}. Estimates of the Hubble constant have been made using this technique for a number of clusters (e.g. \\cite{Jones95, Grainge96, Holzapfel97, Tsuboi98, Mauskopf00, Reese03, Battistelli03, Udomprasert04}) and can be used to constrain cosmological models. In addition, because the SZ surface brightness for a cluster with a given mass is almost independent of redshift, SZ surveys can give information about the evolution of the number counts of clusters vs. redshift. This depends strongly on the evolution of the so-called dark energy or cosmological constant and therefore SZ surveys have been identified as one of the key probes of the nature of dark energy (e.g. \\cite{Diego02, Weller02, DeDeo05, Albrecht06, Bhattacharya07}). A number of dedicated SZ surveys are already producing results, in particular the Atacama Cosmology Telescope (ACT) and the South Polar Telescope (SPT) \\citep{Plagge09, Hincks09, High10}. Most of the SZ detections reported to date have been made at low frequencies corresponding to the SZ decrement. Follow-up photometric or spectroscopic measurements of known clusters at higher frequencies would significantly improve the precision in the measurement of the kinematic SZ effect as well as constraining possible contamination in the low frequency data. Measurements at high frequencies corresponding to the SZ increment suffer from confusion from emission from dusty galaxies, including background high redshift galaxies amplified by the gravitational lensing of the cluster as well as increased atmospheric contamination from ground-based telescopes. Accurate measurement of the SZ increment requires a combination of angular resolution sufficient to resolve and remove point sources combined with high sensitivity and control of systematics necessary to detect the more diffuse SZ signal. This paper presents analysis of observations of the galaxy cluster Abell 1835 using the Bolometric Camera (Bolocam) mounted on the Caltech Submillimeter Observatory (CSO), situated on the summit of Mauna Kea, Hawaii. We also compare the results of this analysis with other sets of data taken at different wavelengths. Abell 1835 is one of the most luminous clusters observed in the ROSAT catalogue and is well-known as a cooling core cluster. It is also known to contain two lensed sub-mm point sources, SMM J14009+0252 and SMM J140104+0252 (see e.g. \\citet{Ivinson00, Zemcov07}). The paper is organized as follows: Section 2 describes the observations with the Bolocam instrument; Section 3 describes the analysis pipeline developed for processing the data; Section 4 describes the modelling of the data and the determination of the characteristic parameters of the model (as well as the errors in their values), and Section 5 discusses the results of the analysis of the Bolocam data and their combination with other literature results. ", "conclusions": "The observation and analysis of 1.1 mm Bolocam jiggle-map data of Abell 1835, including details of the process of converting the raw data to maps of the cluster, have been discussed. A parameter search has been used to determine the best fit values of the central Compton parameter for the cluster and the fluxes for the point sources SMM J140104+0252 and SMM J14009+0252. These values are found to be $y_{0}=(4.68\\pm0.48\\pm0.82)\\times 10^{-4}$, 6.5$\\pm{2.0}$ mJy and 11.3$\\pm{1.9}$ mJy, respectively with a statistical signal-to-noise of 10.5, 3.3 and 6.0 respectively. If the model is refined further to include a point source, representing dust emission from the cD galaxy in Abell 1835, the Compton parameter for the cluster is found to be $y_{0}=(4.34\\pm0.52\\pm0.69)\\times 10^{-4}$ The value for $y_{0}$ was compared to other literature results and found to agree well. The SZ spectrum for Abell 1835, based upon these values for $y_{0}$ was fit with the full SZ form, including both thermal and kinetic effects and relativistic corrections up to 7th order. The values of $y_{0}$ and $v_z$ which optimized this fit were found to be $(3.60\\pm0.24)\\times 10^{-4}$ and $-226\\pm 275$~km/s, which are both in agreement with previous literature results. It is also concluded that, in order to better evaluate the contamination from dust and accurately characterize the spectrum of point sources in the same field as SZ clusters more results need to be obtained at shorter wavelengths in order to obtain more a detailed spectral energy distribution." }, "1005/1005.3587_arXiv.txt": { "abstract": "Because of an old quasar APM $08279+5255$ at $z=3.91$, some dark energy models face the challenge of the cosmic age problem. It has been shown by Wei and Zhang [Phys. Rev. D {\\bf 76}, 063003 (2007)] that the holographic dark energy model is also troubled with such a cosmic age problem. In order to accommodate this old quasar and solve the age problem, we propose in this Letter to consider the interacting holographic dark energy in a non-flat universe. We show that the cosmic age problem can be eliminated when the interaction and spatial curvature are both involved in the holographic dark energy model. ", "introduction": "\\label{sec1} The fact that our universe is undergoing accelerated expansion has been confirmed by lots of astronomical observations such as type Ia supernovae (SNIa)~\\cite{Riess98}, large scale structure (LSS)~\\cite{Tegmark04} and cosmic microwave background (CMB) anisotropy~\\cite{Spergel03}. It is the most accepted idea that this cosmic acceleration is caused by some kind of negative-pressure matter known as dark energy whose energy density has been dominative in the universe. The combined analysis of cosmological observations indicates that the universe today consists of about 70\\% dark energy, 30\\% dust matter (cold dark matter plus baryons), and negligible radiation. The famous cosmological constant $\\lambda$ introduced first by Einstein is the simplest candidate for dark energy. However, the cosmological constant scenario has to face the so-called ``fine-tuning problem'' and ``cosmic coincidence problem''~\\cite{dereview}. Many dark energy models have been proposed, while the nature of dark energy is still obscure. Besides quintessence~\\cite{quintessence}, a wide variety of scalar-field dark energy models have been studied including $k$-essence~\\cite{kessence}, hessence~\\cite{hessence}, phantom~\\cite{phantom}, tachyon~\\cite{tachyon}, quintom~\\cite{quintom}, ghost condensate~\\cite{ghost}, etc. In addition, there are other proposals on dark energy such as interacting dark energy models~\\cite{intde}, brane world models~\\cite{brane}, Chaplygin gas models~\\cite{cg}, Yang-Mills condensate models~\\cite{YMC}, and so on. The dark energy problem is essentially an issue of quantum gravity, owing to the concern of the vacuum expectation value of some quantum fields in a universe governed by gravity. However, by far, we have no a complete theory of quantum gravity yet. So, it seems that we have to consider the effects of gravity in some effective quantum field theory in which some fundamental principles of quantum gravity could be taken into account. It is commonly believed that the holographic principle \\cite{hp} is just a fundamental principle of quantum gravity. Based on the effective quantum field theory, Cohen et al. \\cite{r16} pointed out that the quantum zero-point energy of a system with size $L$ should not exceed the mass of a black hole with the same size, i.e., $L^3\\Lambda^4\\leq LM_{Pl}^2$, where $\\Lambda$ is the ultraviolet (UV) cutoff of the effective quantum field theory, which is closely related to the quantum zero-point energy density, and $M_{Pl}\\equiv 1/\\sqrt{8\\pi G}$ is the reduced Planck mass. This observation relates the UV cutoff of a system to its infrared (IR) cutoff. When we take the whole universe into account, the vacuum energy related to this holographic principle can be viewed as dark energy (its energy density is denoted as $\\rho_{\\Lambda}$ hereafter). The largest IR cutoff $L$ is chosen by saturating the inequality, so that we get the holographic dark energy density \\begin{equation} \\label{eq1}\\rho_{\\Lambda}=3c^2M_{Pl}^2L^{-2} \\end{equation} where $c$ is a numerical constant characterizing all of the uncertainties of the theory, and its value can only be determined by observations. If we take $L$ as the size of the current universe, say, the Hubble radius $H^{-1}$, then the dark energy density will be close to the observational result. However, Hsu \\cite{r17} pointed out that this yields a wrong equation of state for dark energy. Subsequently, Li \\cite{Li04} suggested to choose the future event horizon of the universe as the IR cutoff of this theory. This choice not only gives a reasonable value for dark energy density, but also leads to an accelerated universe. Moreover, the cosmic coincidence problem can also be explained successfully in this model, provided that the inflation lasts for more than 60 $e$-folds. Most recently, a calculation of the Casimir energy of the photon field in a de Sitter space is performed \\cite{Li:2009pm}, and it is a surprising result that the Casimir energy is indeed proportional to the size of the horizon (the usual Casimir energy in a cavity is inversely proportional to the size of the cavity), in agreement with the holographic dark energy model. Up to now, the holographic dark energy model has been tested by various observational data including SNIa~\\cite{r38}, SNIa+BAO+CMB~\\cite{holodata,Li:2009zs}, X-ray gas mass fraction of galaxy clusters~\\cite{r40}, differential ages of passively evolving galaxies~\\cite{r41}, Sandage-Leob test~\\cite{r42}, and so on \\cite{Shen:2004ck}. These analyses show that the holographic dark energy model is consistent with the observational data. However, Wei and Zhang~\\cite{hao} used some old high redshift objects (OHROs) to test the holographic dark energy model and found that the original holographic dark energy model can be ruled out unless a lower Hubble constant (e.g., $h=0.56$) is taken. So, according to Ref.~\\cite{hao}, there is a cosmic age crisis in the holographic dark energy model. In fact, many dark energy models are in the face of such a cosmic age problem. In history, the cosmic age problem has been focused in cosmology for several times. At present, the cosmic age crisis coming from some OHROs appears again in cosmological models, even though dark energy is involved in the models. In cosmology there is a very basic principle that the universe cannot be younger than its constituents. So, if the age of some astronomical object (at some redshift) is measured accurately, then it can be used to test cosmological models according to this simple age principle. Now, there are some OHROs discovered, for example, the $3.5$ Gyr old galaxy LBDS 53W091 at redshift $z=1.55$~\\cite{r46} and the $4.0$ Gyr old galaxy LBDS 53W069 at redshift $z=1.43$~\\cite{r48}. In particular, the old quasar APM $08279+5255$ at redshift $z=3.91$ is an important one, which has been used as a ``cosmic clock'' to constrain cosmological models. Its age is estimated to be $2.0-3.0$ Gyr \\cite{r51}. These three OHROs at $z=1.43$, 1.55 and 3.91 have been used to test many dark energy models, including the $\\Lambda$CDM model~\\cite{r44}, the general EoS dark energy model~\\cite{r57}, the scalar-tensor quintessence model~\\cite{r58}, the $f(R)=\\sqrt{R^2-R_0^2}$ model~\\cite{r59}, the DGP braneworld model~\\cite{r60}, the power-law parameterized quintessence model~\\cite{r62}, the Yang-Mills condensate model~\\cite{Tong:2009mu}, the holographic dark energy model~\\cite{hao}, the agegraphic dark energy model~\\cite{Zhang:2007ps}, and so on. These investigations show that the two OHROs at $z=1.43$ and 1.55 can be easily accommodated in most dark energy models, whereas the OHRO at $z=3.91$ cannot, even in the $\\Lambda$CDM model~\\cite{r44} and the holographic dark energy model ~\\cite{hao}. In this Letter, we revisit the cosmic age problem in the holographic dark energy model. We consider an interacting holographic dark energy model in a non-flat universe. We will show that the age crisis in the original holographic dark energy model can be avoided when the interaction and the spatial curvature are involved in the holographic dark energy model. ", "conclusions": "\\label{sec4} In this Letter, we have revisited the cosmic age problem in the holographic dark energy model. The cosmic age problem brought by the old quasar APM $08279+5255$ has caused trouble to many cosmological models, and the holographic dark energy model is not an exception either \\cite{hao}. In order to accommodate the old quasar APM $08279+5255$ in the holographic dark energy model, we propose to consider the interaction between dark energy and matter in the model. We have shown that the quasar indeed can be accommodated in the holographic dark energy model when an appropriate interaction strength is chosen. Taking the current observational constraints \\cite{Li:2009zs} into account, we have demonstrated that both interaction and spatial curvature should be simultaneously involved in the holographic dark energy model. It has been shown that if such a sophisticated case is considered the quasar APM $08279+5255$ can be accommodated and the cosmic age problem can thus be avoided in the holographic dark energy model. The price of solving the age problem in this way is also apparent, i.e., the model involves too many free parameters, which may weaken the plausibility of the model, to some extent. It is well known that the consideration of interaction in the holographic dark energy can be used to avoid the future big-rip singularity caused by $c<1$ \\cite{Li:2009zs,Li:2008zq}. In this Letter we have provided another advantage for the consideration of interaction in the holographic dark energy, i.e., the interaction between dark energy and matter can also be used to avoid the age problem caused by the old quasar. So, our result can be viewed as a further support to the interacting holographic dark energy model. Of course, we have to confess that the age problem would still exist if some extreme cases are taken into account, say, a much larger possible age of the quasar $t_{obj}$ with a larger $h$. It is remarkable that the age of the old quasar APM $08279+5255$ has not been measured accurately yet, and the age problem caused by this quasar has troubled many dark energy models (including the $\\Lambda$CDM model). It is expected that the future accurate measurement on the age of this old quasar would eliminate the cosmic age crisis in dark energy models." }, "1005/1005.5642_arXiv.txt": { "abstract": "In a pilot project to study the relationship between star formation and molecular gas properties in nearby normal early-type galaxies, we have obtained observations of dense molecular gas tracers in the four galaxies of the {\\tt SAURON} sample with the strongest $^{12}$CO emission. We used the Institut de Radio Astronomie Millimetrique (IRAM) 30m telescope $3$ and $1$~mm heterodyne receivers to observe $^{13}$CO(J=$1$--$0$), $^{13}$CO(J=$2$--$1$), HCN(J=$1$--$0$) and HCO$^+$(J=$1$--$0$). We report the detection of $^{13}$CO emission in all four {\\tt SAURON} sources and HCN emission in three sources, while no HCO$^+$ emission was found to our detection limits in any of the four galaxies. We find that the $^{13}$CO/$^{12}$CO ratios of three {\\tt SAURON} galaxies are somewhat higher than those in galaxies of different Hubble types. The HCN/$^{12}$CO and HCN/$^{13}$CO ratios of all four {\\tt SAURON} galaxies resemble those of nearby Seyfert and dwarf galaxies with normal star formation rates, rather than those of starburst galaxies. The HCN/HCO$^+$ ratio is found to be relatively high (i.e., $>$1) in the three {\\tt SAURON} galaxies with detected HCN emission, mimicking the behaviour in other star-forming galaxies but being higher than in starburst galaxies. When compared to most galaxies, it thus appears that $^{13}$CO is enhanced (relative to $^{12}$CO) in three out of four {\\tt SAURON} galaxies and HCO$^+$ is weak (relative to HCN) in three out of three galaxies. All three galaxies detected in HCN follow the standard HCN--infrared luminosity and dense gas fraction--star formation efficiency correlations. As already suggested by $^{12}$CO observations, when traced by infrared radiation, star formation in the three {\\tt SAURON} galaxies thus appears to follow the same physical laws as in galaxies of different Hubble types. The star formation rate and fraction of dense molecular gas however do not reach the high values found in nearby starburst galaxies, but rather resemble those of nearby normal star-forming galaxies. ", "introduction": "\\label{sec:intro} \\begin{figure*} \\begin{center} \\resizebox{\\hsize}{!}{\\rotatebox{-90}{\\includegraphics{sauron-spec4.eps}}} \\caption{Emission line spectra of the four {\\tt SAURON} early-type galaxies observed. From top to bottom in each panel: HCO$^+$(J=1--0), HCN(J=1--0), $^{13}$CO(J=1--0) and $^{13}$CO(J=2--1). From left to right: NGC~3032, NGC~4150, NGC~4459 and NGC~4526. The $^{13}$CO emission is overlaid with the spectra of the $^{12}$CO emission ({\\it blue}) from \\citet{com07}, scaled down to match the $^{13}$CO line emission; the scaling factors are indicated in the respective spectra. The (single, dual, triple) Gaussian fits to the line spectra are plotted in red.} \\label{fig:fig1} \\end{center} \\end{figure*} Over the last twenty years, growing evidence has accumulated that nearby early-type galaxies (E/S0s) are not always devoid of molecular gas, and can even harbour a substantial amount of it \\citep*[e.g.,][]{lee91,wik95,bet03,sag07,com07,you09}, often settled in a regularly rotating disc \\citep*[e.g.,][]{you02,you05,you08,cro08,cro09,cro10}. Interestingly, not all early-type galaxies with a substantial amount of molecular gas show obvious signs of (current) star formation \\citep[SF; e.g.,][]{cro08,cro09,cro10,you08}. However, the converse is generally true \\citep[e.g.,][]{jeo09}, and the $\\ga30$~per cent fraction of SF objects within the nearby early-type galaxy population \\citep[e.g.,][]{yi05} appears to be superficially consistent with their $\\approx25$~per cent CO detection rate \\citep[e.g.,][]{sag06,sag07,com07,you09}. This fact raises many as yet unanswered questions: how is the (current) SF ignited in early-type galaxies? Is SF coupled to the molecular discs, i.e.\\ do all molecular gas discs form stars? Does SF only take place in the densest regions, perhaps ill-traced by the $^{12}$CO lines used in most surveys? How, if at all, does the morphology of the molecular discs and the physical characteristics of the molecular gas correlate with the ages of the stars? In order to answer some of these questions, it is essential to study the spatially- and kinematically-resolved properties of the stellar populations, as well as those of the molecular gas in early-type galaxies (E/S0). CO emission can also easily trace signs of gravitational interactions or other dynamical perturbations that can help to understand how star formation is induced. Observations of a sample of 48 nearby E/S0 galaxies have been carried out with a custom-designed panoramic optical integral-field spectrograph \\citep{bac01} within the framework of the {\\tt SAURON} project. A description of the sample and illustrative results can be found in \\citet{zee02}. Results include the detection of kinematic misalignments, twists and decoupled cores as well as central discs in roughly half of the sources \\citep{ems04,kra08}. Information on the ionized gas (distribution, kinematics and ionisation; \\citealt{sar06,sar10}) and the age, metallicity and alpha-element enhancement of the stellar populations \\citep[][]{kun06,kun10} are also available. The specific angular momentum of the galaxies appears to be a dominant factor in their evolution \\citep{ems07,cap07}, and all CO detections are found in fast-rotating systems \\citep{you09}. In order to study star formation in the {\\tt SAURON} sources and understand the various phenomena observed in the stellar and ionised gas data, observations looking for central molecular gas in the nearby {\\tt SAURON} E/S0 galaxies were obtained by \\citet{com07}. $^{12}$CO emission was detected in $12$ of $43$ galaxies observed with the Institut de Radio Astronomie Millimetrique (IRAM) 30m telescope, suggesting typical gas masses in a range between a few 10$^{6}$ to 10$^{8}$ M$_\\odot$. \\citet{com07} further found that the amount of molecular gas in the galaxies correlates with their far-infrared (FIR) luminosities, an often-used indicator of star formation. The gas-rich sources show the most pronounced star formation, as for normal star-forming spiral galaxies. However, the lack of correlations between the molecular gas and most stellar properties suggests that the molecular gas is largely unrelated to the old, pre-existing stellar population, and might have been externally accreted, as a result for instance of a merger or interaction event \\citep{com07}. \\begin{table} \\caption{Basic parameters of the four {\\tt SAURON} galaxies.} \\label{tab:tab1} \\centering \\begin{tabular}{lrrrrc} \\hline NGC & RA(J2000) & DEC(J2000) & $V_\\odot$ & D & Activity \\\\ & (hh:mm:ss.s) & (dd:mm:ss) & (km~s$^{-1}$) & (Mpc) & Type \\\\ \\hline 3032 & 09:52:08.2 & $+$29:14:10 & 1533 & 21.7 & H~{\\small II} \\\\ 4150 & 12:10:33.6 & $+$30:24:06 & 226 & 13.7 & none \\\\ 4459 & 12:29:00.0 & $+$13:58:43 & 1210 & 16.3 & H~{\\small II}+L \\\\ 4526 & 12:34:03.0 & $+$07:41:57 & 550 & 16.3 & none \\\\ \\hline \\end{tabular} Notes: Coordinates, distances, velocities and activity types are taken from the NASA/IPAC Extragalactic Database (NED). Galaxy types: L=LINER, H~{\\small II}=star formation bursts. \\end{table} \\begin{table*} \\caption{Properties of the molecular line spectra.} \\label{tab:tab2} \\centering \\begin{tabular}{lrrrrrr} \\hline \\\\[-0.2cm] Line & $^{12}$CO(J=1--0) & $^{12}$CO(J=2--1) & $^{13}$CO(J=1--0) & $^{13}$CO(J=2--1) & HCN(J=1--0) & HCO$^+$(J=1--0)$^b$ \\\\ Parameters$^a$ & (CYB07) & (CYB07) & (this paper) & (this paper) & (this paper) & (this paper)\\\\[0.2cm] \\hline \\\\[-0.2cm] NGC3032: & \\\\ \\hskip 0.4cm $I$ (K~km~s$^{-1}$) & 9.3$\\pm$0.3 & 7.9$\\pm$0.2 & 0.9$\\pm$0.1 & 1.4$\\pm$0.2 & 0.32$\\pm$0.09 & $<$0.2\\\\ \\hskip 0.4cm $$ (km~s$^{-1}$) & 1549$\\pm$2 & 1537$\\pm$2 & 1571$\\pm$10 & 1561$\\pm$10 & 1571$\\pm$13 & \\\\ \\hskip 0.4cm $\\Delta v$ (km~s$^{-1}$) & 129$\\pm$4 & 103$\\pm$3 & 130$\\pm$14 & 104$\\pm$10 & 160$\\pm$30 & \\\\ \\\\[-0.2cm] NGC4150: & \\\\ \\hskip 0.4cm $I$ (K~km~s$^{-1}$) & 6.1$\\pm$0.5 & 13.2$\\pm$0.5 & 0.4$\\pm$0.1 & 1.1$\\pm$0.2 & $<$0.2 & $<$0.2 \\\\ \\hskip 0.4cm $$ (km~s$^{-1}$) & 204$\\pm$7 & 200$\\pm$3 & 221$\\pm$14 & 226$\\pm$12 & & \\\\ \\hskip 0.4cm $\\Delta v$ (km~s$^{-1}$) & 158$\\pm$14 & 189$\\pm$8 & 150$\\pm$30 & 160$\\pm$30 & & \\\\ \\\\[-0.2cm] NGC4459: & \\\\ \\hskip 0.4cm $I$ (K~km~s$^{-1}$) & 10.9$\\pm$0.5& 14.3$\\pm$0.5 & 3.2$\\pm$0.2 & 4.9$\\pm$0.4 & 0.8$\\pm$0.1 & $<$0.7\\\\ \\hskip 0.4cm $$ (km~s$^{-1}$) & 1169$\\pm$9 & 1176$\\pm$8 & 1194$\\pm$22 & 1179$\\pm$40 & 1140$\\pm$20 & \\\\ \\hskip 0.4cm $\\Delta v$ (km~s$^{-1}$) & 477$\\pm$16 & 405$\\pm$15 & 390$\\pm$20 & 380$\\pm$20 & 440$\\pm$70 & \\\\ \\\\[-0.2cm] NGC4526: & \\\\ \\hskip 0.4cm $I$ (K~km~s$^{-1}$) & 23.8$\\pm$0.8& 37.4$\\pm$0.7 & 6.0$\\pm$0.2 & 11$\\pm$0.4 & 1.6$\\pm$0.1 & $<$0.7\\\\ \\hskip 0.4cm $$ (km~s$^{-1}$) & 697$\\pm$2 & 695$\\pm$5 & 605$\\pm$30 & 600$\\pm$30 & 606$\\pm$40 & \\\\ \\hskip 0.4cm $\\Delta v$ (km~s$^{-1}$) & 650$\\pm$23 & 533$\\pm$11 & 687$\\pm$26 & 677$\\pm$27 & 760$\\pm$50 & \\\\ \\\\[-0.2cm] \\hline \\end{tabular} Notes: $^a$Line properties have been obtained by fitting a single Gaussian to the data from \\citet{com07} and this paper; $I\\equiv\\int~T_{\\rm mb}~dv$ is the velocity-integrated line intensity; $$ is the mean (central) heliocentric velocity; $\\Delta v$ is the velocity width (FWHM). $^b$ Determined from 3$\\sigma$ upper limits from spectrum at $\\sim$100~km~s$^{-1}$ spectral resolution assuming the same FWHM as found for the HCN(J=1--0) or $^{13}$CO(J=1--0) line. \\end{table*} However, as $^{12}$CO emission has been found to be a rather unreliable tracer of the dense gas where star formation takes place \\citep[e.g.,][]{gao04a,gao04b}, as a pilot study we focus here on the volume and/or column density tracers HCN, HCO$^+$ and $^{13}$CO in the four {\\tt SAURON} galaxies with the brightest $^{12}$CO emission. Basic parameters of these four sources are given in Table~\\ref{tab:tab1}. They form an interesting subset of the {\\tt SAURON} sample as they nicely populate the two main evolutionary groups: i) those with a more 'troubled' recent past, likely involving accretion and/or minor interactions with other galaxies (NGC~3032 and NGC~4150); and ii) those with a more peaceful immediate past dominated by secular evolution (NGC~4459 and NGC~4526). The first group appears to harbor a large fraction of widely distributed young stars while the second group possesses a lower young stellar fraction that is spatially limited to the center \\citep[][]{kun10}. Additionally, NGC~3032 shows signs that its young stars have a lower metallicity than its bulk stellar population. Nevertheless, only NGC~3032 and NGC~4459 show signs of circumnuclear activity both in the form of H~{\\small II} regions, indicative of star formation bursts, and, in the case of NGC~4459 also in the form of a low-ionization nuclear emission region (LINER) nucleus. The paper is organised as follows. Section~\\ref{sec:obs} presents the observations. Section~\\ref{sec:results} consists of the results and a discussion of the data and their trends. Finally, section~\\ref{sec:conclusions} summarizes our findings and conclusions. \\begin{table*} \\caption{Molecular line ratios for a representative set of nearby galaxies covering the Hubble sequence and different activity types.} \\label{tab:tab3} \\begin{center} \\begin{tabular}{lcccccccc} \\hline Source & Activity Type$^a$ & $\\frac{^{13}{\\rm CO(J=1-0)}}{^{12}{\\rm CO(J=1-0)}}$ & $\\frac{^{13}{\\rm CO(J=2-1)}}{^{12}{\\rm CO(J=2-1)}}$ & $\\frac{{\\rm HCN(J=1-0)}}{^{13}{\\rm CO(J=1-0)}}$ & $\\frac{{\\rm HCN(J=1-0)}}{^{12}{\\rm CO(J=1-0)}}$ & $\\frac{{\\rm HCN(J=1-0)}}{{\\rm HCO^+(J=1-0)}}$ & Ref$^b$ & Region$^c$ \\\\%[0.15cm] \\hline \\multicolumn{9}{c}{Ellipticals: $-$6 $\\leq$ T $\\leq$ $-$4}\\\\[0.1cm] Cen~A & Sy2 & $\\sim$0.07 & $\\sim$0.08 & $\\sim$0.9 & $\\sim$0.06 & $\\sim$0.6$^d$ & (1) & Centre \\\\ & & $\\sim$0.1 & $\\sim$0.09 & $\\sim$0.2 & $\\sim$0.02 & -- & (1) & Dust Lane \\\\[0.2cm] \\multicolumn{9}{c}{Sauron (S0): $-$4 $<$ T $\\leq$ 0}\\\\[0.1cm] NGC~3032 & H~{\\small II} & 0.09$\\pm$0.01 & 0.20$\\pm$0.02 & 0.36$\\pm$0.07 & 0.03$\\pm$0.01 & $>$1.7 & (2) & Centre \\\\ NGC~4150 & NONE & 0.07$\\pm$0.01 & 0.08$\\pm$0.01 & $<$0.5 & $<$0.03 & -- & (2) & Centre \\\\ NGC~4459 & H~{\\small II}+L & 0.30$\\pm$0.02 & 0.34$\\pm$0.02 & 0.23$\\pm$0.03 & 0.07$\\pm$0.01 & $>$1.1 & (2) & Centre \\\\ NGC~4526 & NONE & 0.26$\\pm$0.01 & $\\sim$0.3 & 0.35$\\pm$0.02 & 0.09$\\pm$0.01 & $>$2.3 & (2) & Centre \\\\[0.15cm] \\multicolumn{9}{c}{Seyferts: 0 $<$ T $<$ +8}\\\\[0.1cm] NGC~1068$^\\star$ & Sy2 & 0.09$\\pm$0.02 & 0.06$\\pm$0.01 & 1.6$\\pm$0.6 & 0.13$\\pm$0.02 & 1.7$\\pm$0.2 & (3,6) & Centre\\\\ & & $\\sim$0.02 & 0.062$\\pm$0.002 & $\\sim$80 & $\\sim$1.4 & $\\sim$1.4 & (4,19) & CND$^d$\\\\ NGC~2237 & ? & $\\sim$0.1 & -- & 0.41$\\pm$0.07 & 0.04$\\pm$0.01 & -- & (4,5) & Centre \\\\ NGC~3079 & Sy2/L & 0.06$\\pm$0.01 & 0.07$\\pm$0.01 & 0.6$\\pm$0.2 & 0.039$\\pm$0.09 & $>$5 & (6,23) & Centre \\\\ NGC~3627 & Sy2/L & $\\sim$0.06 & -- & $\\sim$1.1 & $\\sim$0.07 & 1.0$\\pm$0.1 & (3,8) & Centre \\\\ NGC~3628 & L+H~{\\small II} & 0.12$\\pm$0.02 & 0.10$\\pm$0.02 & 0.34$\\pm$0.09 & 0.027$\\pm$0.007 & 1.6$\\pm$0.5 & (22,26) & Centre \\\\ NGC~3982 & Sy1.9+H~{\\small II} & 0.07$\\pm$0.01 & -- & 0.32$\\pm$0.08 & 0.022$\\pm$0.005 & -- & (4) & Centre \\\\ NGC~4051 & Sy1/NL & 0.06$\\pm$0.01 & -- & 0.32$\\pm$0.08 & 0.019$\\pm$0.004 & -- & (4) & Centre \\\\ NGC~4258 & Sy1.9/L & 0.10$\\pm$0.01 & -- & 0.29$\\pm$0.03 & 0.029$\\pm$0.003 & -- & (4) & Centre \\\\ NGC~4388 & Sy2 & 0.03$\\pm$0.01 & -- & 0.3$\\pm$0.1 & 0.010$\\pm$0.003 & -- & (4) & Centre \\\\ NGC~4826 & Sy2+H~{\\small II} & 0.12$\\pm$0.02 & 0.13$\\pm$0.02 & 0.40$\\pm$0.04 & 0.06$\\pm$0.01 & 1.7$\\pm$0.1 & (3,6) & Centre\\\\ NGC~4945 & Sy2 & 0.06$\\pm$0.01 & 0.11$\\pm$0.01 & 0.75$\\pm$0.02 & 0.05$\\pm$0.01 & $\\sim$1 & (27,28) & Centre \\\\ NGC~5033$^\\star$ & Sy1.8 & 0.12$\\pm$0.01 & -- & 0.36$\\pm$0.03 & 0.044$\\pm$0.005 & $\\sim$1.9 & (4) & CND$^d$ \\\\ NGC~5194$^\\star$ & Sy2 & $\\sim$0.1-0.2 & $\\sim$0.1 & $\\sim$2 & $\\sim$0.5 & 1.4$\\pm$0.1 & (3,7,8) & CND$^d$ \\\\ NGC~6951$^\\star$ & Sy2/L & -- & 0.13$\\pm$0.02 & $\\gtrsim$2 & $\\gtrsim$2 & 1.4$\\pm$0.1 & (3,5) & CND$^d$ \\\\ NGC~7172 & Sy2+H~{\\small II} & $\\sim$0.1-0.2 & -- & 0.33$\\pm$0.07 & $\\sim$0.03 & -- & (4) & Centre \\\\ NGC~7314 & Sy1.9 & $\\sim$0.1-0.2 & -- & $<$0.4 & $\\sim$0.05 & -- & (4) & Centre \\\\ NGC~7331 & L & $\\sim$0.14 & $\\sim$0.17 & $\\sim$0.03 & $\\sim$0.04 & $>$1.7 & (22,24) & Center \\\\ NGC~7582 & Sy2 & -- & -- & -- & 0.03$\\pm$0.01 & 1.3$\\pm$0.6 & (29) & Centre \\\\ NGC~7469$^\\star$ & Sy1.2 & $\\sim$0.06 & $\\sim$0.05 & 0.95$\\pm$0.1 & $\\sim$0.1 & $\\sim$1.4 & (4,6,7) & Centre \\\\ Mrk~231$^{\\dag\\dag}$ & Sy1 & -- & $<$0.03 & -- & $\\sim$0.2 & 0.72$\\pm$0.09 & (29,30) & Centre \\\\ Mrk~331$^\\dag$ & Sy2+H~{\\small II} & -- & -- & -- & $\\sim$0.2 & $\\sim$1.3 & (29) & Centre \\\\ IRAS~05414+5840$^\\dag$ & Sy2 & -- & -- & -- & 0.040$\\pm$0.006 & 0.6$\\pm$0.2 & (29) & Centre\\\\[0.15cm] \\multicolumn{9}{c}{Starbursts: 0 $<$ T $<$ +8}\\\\[0.1cm] M82 & SB+H~{\\small II} & $\\sim$0.03-0.1 & $\\sim$0.1 & $\\sim$1 & $\\sim$0.2 & 0.7$\\pm$0.1 & (3,8,9) & Centre \\\\ M83 & SB+H~{\\small II} & 0.10$\\pm$0.01 & $\\sim$0.10$\\pm$0.03 & $\\sim$1 & $\\sim$0.1 & 1.3$\\pm$0.1 & (23,25) & Centre \\\\ NGC~253 & SB+Sy2+H~{\\small II} & $\\sim$0.1 & $\\sim$0.1-0.3 & $\\sim$0.2-1.0 & $\\sim$0.05-0.3 & 0.8$\\pm$0.1 & (10,11) & Centre \\\\ IRAS~210293 & SB? & 0.06$\\pm$0.01 & 0.11$\\pm$0.02 & --\t & --\t & -- & (20) & Centre \\\\ NGC~660 & SB?+Sy2/L+H~{\\small II} & 0.07$\\pm$0.01 & 0.05$\\pm$0.01 & 0.7$\\pm$0.3 & 0.05$\\pm$0.02 & 1.0$\\pm$0.3 & (20,21) & Centre \\\\ NGC~891\t & H~{\\small II} & $\\sim$0.2 & -- & -- & 0.020$\\pm$0.005& 1.1$\\pm$0.5 & (29,31) & Centre \\\\ NGC~986 & SB+H~{\\small II} & 0.10$\\pm$0.01 & 0.06$\\pm$0.01 & 0.9$\\pm$0.2 & 0.09$\\pm$0.02 & - & (20) & Centre \\\\ NGC~1808 & Sy2 & 0.06$\\pm$0.01 & 0.07$\\pm$0.01 & 1.0$\\pm$0.2 & 0.06$\\pm$0.01 & 0.56$\\pm$0.04 & (20,21) & Centre \\\\ NGC~2146 & SB+H~{\\small II} & 0.08$\\pm$0.01 & 0.10$\\pm$0.01 & 0.8$\\pm$0.2 & 0.06$\\pm$0.01 & 0.77$\\pm$0.05 & (4,20,21) & Centre \\\\ NGC~2369$^\\dag$ & SB? & -- & -- & -- & 0.04$\\pm$0.01 & 1.0$\\pm$0.4 & (29) & Centre \\\\ NGC~2903\t & SB?+H~{\\small II} & -- & -- & -- & 0.020$\\pm$0.005& 0.4$\\pm$0.1 & (29) & Centre \\\\ NGC~3256 & SB+H~{\\small II} & 0.03$\\pm$0.01 & 0.10$\\pm$0.04 & 2.0$\\pm$0.4 & 0.06$\\pm$0.01 & -- & (20) & Centre \\\\ NGC~4355\t & Sy2 & -- & -- & -- & 0.8$\\pm$0.2 & 0.08$\\pm$0.01 & (29) & Centre \\\\ NGC~6946 & SB+Sy2+H~{\\small II} & $\\sim$0.05-0.1 & -- & $\\sim$1 & $\\sim$0.2 & 1.2$\\pm$0.1 & (3,4,8) & Centre \\\\ NGC~7552 & SB+L+H~{\\small II} & 0.07$\\pm$0.01 & 0.11$\\pm$0.02 & 1.1$\\pm$0.2 & 0.08$\\pm$0.01 & 0.9$\\pm$0.2 & (20,21) & Centre \\\\ \\hline \\end{tabular} \\end{center} Notes: T is the numerical Hubble type. Sources marked with a ``$\\star$'' show a particular gas chemistry with unusually high HCN/$^{12}$CO, HCN/$^{13}$CO and/or HCN/HCO$^+$ ratios; sources marked with a ``\\dag'' or ``\\dag\\dag'' are LIRGs and ULIRGs respectively. Values marked with a ``$\\sim$'' or given as a range indicate either values averaged over (or a range of values for) several positions/observations or values for which an accurate estimate of the uncertainty was not possible. $^a$ Sy=Seyfert; L=LINER, SB= starburst, H~{\\small II}=star formation bursts/H~{\\small II} regions, classifications taken from NED; $^b$ References: (1) \\citet*{wil00}, \\citet{wild97} and \\citet{wik97}; (2) this paper; (3) \\citet{kri08}; (4) Krips et al.\\ (2010, in prep.); (5) \\citet{pet03}; (6) \\citet{isr09a}; (7) \\citet*{isr06}; (8) \\citet{pag01}; (9) \\citet{mat00}; (10) \\citet{sak06}; (11) \\citet{sor01}; (12) \\citet{wils97}; (13) \\citet{tos07}; (14) \\citet{bro05}; (15) \\citet{pet98}; (16) \\citet{bol05}; (17) \\citet{hei99}; (18) \\citet{chin97}; (19) \\citet{gar08}; (20) \\citet{aalto95}; (21) \\citet{baan08} (and references therein); (22) \\citet{mei08}; (23) \\citet{ngrieu92}; (24) \\citet{isr99}; (25) \\citet{isr01}; (26) \\citet{isr09b}; (27) \\citet{hen94}; (28) \\citet{wan04}; (29) \\citet{baan08}; (30) \\citet{glen01}; (31) \\citet{saka97}; (32) \\citet{wil08}. $^c$ Regions in which the line ratios where determined: Centre = central $\\la10$~kpc of a galaxy; CND = circumnuclear disc, i.e.\\ a radius $<1$~kpc; GMC = average over several giant molecular clouds; SA = spiral arms. $^d$ Ratio determined from absorption lines. $^e$Ratios determined from interferometric observations. \\end{table*} \\begin{table*} \\contcaption{} \\begin{center} \\begin{tabular}{lcccccccc} \\hline Source & Activity Type$^a$ & $\\frac{^{13}{\\rm CO(J=1-0)}}{^{12}{\\rm CO(J=1-0)}}$ & $\\frac{^{13}{\\rm CO(J=2-1)}}{^{12}{\\rm CO(J=2-1)}}$ & $\\frac{{\\rm HCN(J=1-0)}}{^{13}{\\rm CO(J=1-0)}}$ & $\\frac{{\\rm HCN(J=1-0)}}{^{12}{\\rm CO(J=1-0)}}$ & $\\frac{{\\rm HCN(J=1-0)}}{{\\rm HCO^+(J=1-0)}}$ & Ref$^a$ & Region$^b$ \\\\%[0.15cm] \\hline \\multicolumn{9}{c}{Starbursts: 0 $<$ T $<$ +8}\\\\[0.1cm] NGC~7771$^\\dag$ & SB+H~{\\small II} & -- & -- & -- & $\\sim$0.05 & $\\sim$1.0 & (29) & Centre \\\\ UGC~2855 & ? & 0.08$\\pm$0.02 & 0.11$\\pm$0.01 & -- & -- & -- & (20) & Centre \\\\ IC~860$^\\dag$ & H~{\\small II} & -- & -- & -- & 0.08$\\pm$0.01 & 0.5$\\pm$0.2 & (29) & Centre \\\\ Arp~220$^{\\dag\\dag}$ & SB+Sy2/L+H~{\\small II} & $<$0.05 & $\\sim$0.05 & -- & $\\sim$0.08 & 0.46$\\pm$0.09 & (29) & Centre \\\\ IRAS~22025+4205$^\\dag$ & SB? & -- & -- & -- & 0.15$\\pm$0.03 & 0.42$\\pm$ 0.2 & (29) & Centre \\\\[0.15cm] \\multicolumn{9}{c}{SF-spirals: 0 $<$ T $<$ +8} \\\\[0.1cm] M33 & H~{\\small II} & $\\sim$0.10 & $\\sim$0.14 & -- & -- & -- & (12) & GMC in SA\\\\ M31 & L? & $\\sim$0.13 & -- & $\\sim$0.15 & $\\sim$0.02 & $\\sim$0.9 & (13,14) & GMC in SA\\\\ Maffei~2 & NONE & 0.02-0.1 & -- & $\\sim$0.4-1.3 & $\\sim$0.02-0.13 & 2.4$\\pm$0.4 & (22,23) & GMC in SA\\\\[0.15cm] \\multicolumn{9}{c}{Dwarf galaxies: +8 $\\leq$ T $\\leq$ +10 }\\\\[0.1cm] IC10 & SB & -- & $\\sim$0.1 & -- & -- & -- & (15) & GMC \\\\ LMC & NONE & $\\sim$0.1 & $\\sim$0.2 & $\\sim$0.2-0.4 & $\\sim$0.03-0.06 & $\\sim$0.3-0.7 & (16,17,18) & GMC \\\\ SMC & NONE & $\\sim$0.1 & $\\sim$0.1-0.2 & $\\sim$0.3-0.4 & $\\sim$0.02 & $\\sim$0.4 & (16,17,18) & GMC \\\\ [0.15cm] \\multicolumn{9}{c}{Peculiar galaxies and/or mergers}\\\\[0.1cm] NGC~6240$^\\dag$ & Sy2/L & $\\sim$0.02 & $\\sim$0.01 & -- & $\\sim$0.09 & $\\sim$0.8 & (29) & Centre \\\\ Mrk~273$^{\\dag\\dag}$ & Sy2/L & -- & -- & -- & $\\sim$0.2 & $\\sim$ 0.4 & (29) & Centre \\\\[0.2cm] NGC~3620 & SB? & -- & -- & -- & 0.08$\\pm$0.03 & $<$0.4 & (29) & Centre \\\\ IC~1623$^\\dag$ & ? & -- & -- & -- & 0.040$\\pm$0.006 & 1.2$\\pm$0.2 & (29) & Centre \\\\ Arp~55$^\\dag$ & L+H~{\\small II} & -- & $<$0.04 & -- & $\\sim$0.1 & $\\sim$0.8 & (29,32) & Centre \\\\ Arp~193$^\\dag$ & L+H~{\\small II} & -- & -- & -- & $\\sim$0.03 & $\\sim$1.9 & (29) & Centre \\\\ Arp~299A$^\\dag$ & H~{\\small II} & $\\sim$0.04 & $\\sim$0.03 & $\\sim$1 & 0.040$\\pm$0.006 & 1.6$\\pm$0.3 & (29) & Centre \\\\ Arp~299B$^\\dag$ & H~{\\small II} & $\\sim$0.04 & $\\sim$0.03 & $\\sim$1 & 0.050$\\pm$0.004 & $\\sim$0.8 & (29) & Centre \\\\ IRAS~12112+0305$^\\dag$& L+H~{\\small II} & -- & -- & -- & $\\sim$0.07 & $\\sim$0.6 & (29) & Centre \\\\ IRAS~15107+0724 & H~{\\small II} & -- & -- & -- & 0.11$\\pm$0.03 & 0.5$\\pm$0.2 & (29) & Centre \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} ", "conclusions": "\\label{sec:conclusions} We detected significant $^{13}$CO in four out of four and HCN emission in three out of four {\\tt SAURON} early-type galaxies, while no HCO$^+$ emission was found in any of the four sources. We find some pronounced differences in the line ratios of the {\\tt SAURON} galaxies when compared to other nearby galaxies of different Hubble and activity types. In particular, $^{13}$CO/$^{12}$CO appears slightly enhanced in three {\\tt SAURON} galaxies compared to other galaxy types. This may indicate different molecular gas excitation conditions and/or chemistry in these sources. The closest resemblance to our four nearby {\\tt SAURON} galaxies is found in nearby star-forming galaxies, including Seyferts and dwarfs. This and the pronounced differences with starburst galaxies suggest that the four {\\tt SAURON} galaxies exhibit star-formation rates and efficiencies more similar to quiescent, normal star-forming galaxies than to starburst galaxies, and that they do not reach the high dense molecular gas fraction found in starbursts. Also, according to the high ($>$1) HCN/HCO$^+$ ratios, PDRs and/or supernovae explosions do not seem to play an important role in the chemistry of the molecular gas in our four targets. The three {\\tt SAURON} galaxies observed in HCN nicely follow the same physical laws concerning star formation and dense molecular gas as other galaxies, at least when infrared radiation is used as the star formation tracer. Although the four {\\tt SAURON} galaxies studied in this paper form a nice subset of the {\\tt SAURON} sample with respect to their general characteristics, they are not equally representative with respect to their molecular gas properties. They have been chosen for their strong $^{12}$CO emission and could thus in principal be extreme cases within the sample. Follow-up observations of the $^{13}$CO, HCN and HCO$^+$ emission in a much larger sample of nearby early-type galaxies are currently underway based on the success of this pilot study. These will eventually allow us to confirm (or discard) the trends found in this pilot project and put them on a sounder statistical basis." }, "1005/1005.3702_arXiv.txt": { "abstract": "{Observations of very high energy $\\gamma$-rays from blazars provide information about acceleration mechanisms occurring in their innermost regions. Studies of variability in these objects allow a better understanding of the mechanisms at play. } {To investigate the spectral and temporal variability of VHE ($>100\\,\\mathrm{GeV}$) $\\gamma$-rays of the well-known high-frequency-peaked BL\\,Lac object \\pks\\ with the H.E.S.S. imaging atmospheric Cherenkov telescopes over a wide range of flux states.} {Data collected from 2005 to 2007 are analyzed. Spectra are derived on time scales ranging from 3 years to 4 minutes. Light curve variability is studied through doubling timescales and structure functions, and is compared with red noise process simulations.} {The source is found to be in a low state from 2005 to 2007, except for a set of exceptional flares which occurred in July 2006. The quiescent state of the source is characterized by an associated mean flux level of $(4.32\\pm0.09_\\mathrm{stat}\\pm0.86_\\mathrm{syst}) \\times 10^{-11}\\,{\\rm cm^{-2}}\\,{\\rm s^{-1}}$ above $200\\,{\\rm GeV}$, or approximately $15\\%$ of the Crab Nebula, and a power law photon index of $\\Gamma=3.53\\pm0.06_\\mathrm{stat}\\pm0.10_\\mathrm{syst}$. During the flares of July 2006, doubling timescales of $\\sim 2\\,{\\rm min}$ are found. The spectral index variation is examined over two orders of magnitude in flux, yielding different behaviour at low and high fluxes, which is a new phenomenon in VHE $\\gamma$-ray emitting blazars. The variability amplitude characterized by the fractional r.m.s. $F_{\\rm var}$ is strongly energy-dependent and is $\\propto E^{0.19\\pm0.01}$. The light curve r.m.s. correlates with the flux. This is the signature of a multiplicative process which can be accounted for as a red noise with a Fourier index of $\\sim 2$.} {This unique data set shows evidence for a low level $\\gamma$-ray emission state from \\pks, which possibly has a different origin than the outbursts. The discovery of the light curve lognormal behaviour might be an indicator of the origin of aperiodic variability in blazars.} \\offprints{santiago.pita@apc.univ-paris7.fr and francesca.volpe@mpi-hd.mpg.de/volpe@llr.in2p3.fr} ", "introduction": "\\label{intro} The BL Lacertae (BL Lac) category of Active Galactic Nuclei (AGN) represents the vast majority of the population of energetic and extremely variable extragalactic very high energy $\\gamma$-ray emitters. Their luminosity varies in unpredictable, highly irregular ways, by orders of magnitude, and at all wavelengths across the electromagnetic spectrum. The very high energy (VHE, $E\\geq100\\,{\\rm GeV}$) $\\gamma$-ray fluxes vary often on the shortest timescales that can be seen in this type of object, with large amplitudes which can dominate the overall output. It hence indicates that the understanding of this energy domain is the most important one for understanding the underlying fundamental variability and emission mechanisms at play in high flux states. It has been, however, difficult to ascertain whether $\\gamma$-ray emission is present only during high flux states or also when the source is in a more stable or quiescent state but with a flux which is below the instrumental limits. The advent of the current generation of atmospheric Cherenkov telescopes with unprecedented sensitivity in the VHE regime gives new insights into these questions. The high frequency peaked BL Lac object (HBL) \\pks, located at a redshift $z=0.117$, initially discovered as a VHE $\\gamma$-ray emitter by the Mark 6 telescope (\\cite{cha99}), has been detected by the first H.E.S.S. telescope in 2002-2003 (\\cite{HESS2155_2003}~2005b). It has been frequently observed by the full array of four telescopes since 2004, either sparsely during the H.E.S.S. monitoring program, or intensely during dedicated campaigns such as that described in \\cite{HESS2155_MWL}~(2005c), showing mean flux levels of $\\sim20\\%$ of the Crab Nebula flux for energies above $200\\,{\\rm GeV}$. During the summer of 2006, \\pks\\ exhibited unprecedented flux levels accompanied by strong variability (\\cite{HESS2155_BigFlare}~2007a), making temporal and spectral variability studies possible on timescales of the order of a few minutes. The VHE $\\gamma$-ray emission is usually thought to originate from a relativistic jet, emanating from the vicinity of a Supermassive Black Hole (SMBH). The physical processes at play are still poorly understood, but the analysis of the $\\gamma$-ray flux spectral and temporal characteristics is well suited to provide better insights. For this goal the data set of H.E.S.S. observations of \\pks\\ between 2005 and 2007 is used. After describing the observations and the analysis chain in Section~\\ref{ObsAna}, the emission from the ``quiescent'', i.e. nonflaring, state of the source will be characterized in Section~\\ref{QuiescentState}. Section~\\ref{SpecVar} details spectral variability related to the source intensity. Section~\\ref{temporal_variab} will focus on the description of the temporal variability during the highly active state of the source, and its possible energy dependence. Section~\\ref{log_normal_process} will illustrate a description of the observed variability phenomenon by a random stationary process, characterized by a simple power density spectrum. Section~\\ref{characteristic_time} will show how limits on the characteristic time of the source can be derived. The multi-wavelength aspects from the high flux state will be presented in a second paper. ", "conclusions": "This data set, which exhibits unique features and results, is the outcome of a long-term monitoring program and dedicated, dense, observations. One of the main results here is the evidence for a VHE $\\gamma$-ray quiescent-state emission, where the variations in the flux are found to have a lognormal distribution. The existence of such a state was postulated by \\cite{ste96} in order to explain the extragalactic $\\gamma$-ray background at 0.03--$100\\,{\\rm GeV}$ detected by EGRET (\\cite{fic96,sre98}) as coming from quiescent-state unresolved blazars. Such a background has not yet been detected in the VHE range, as it is technically difficult with the atmospheric Cherenkov technique to find an isotropic extragalactic emission and even more to distinguish it from the cosmic-ray electron flux (\\cite{egb08}). In addition, the EBL attenuation limits the distance from which $\\sim$TeV $\\gamma$-rays can propagate to $\\sim 1\\,{\\rm Gpc}$ (\\cite{HESSEBL}). As pointed out by \\cite{che00}, emission mechanisms might be simpler to understand during quiescent states in blazars, and they are also the most likely state to be found observationally. In the X-ray band, the existence of a steady underlying emission has also been invoked for two other VHE emitting blazars (Mrk~421, \\cite{fos00}, and 1ES~1959+650, \\cite{gie02}). Being able to separate, and detect, flaring and nonflaring states in VHE $\\gamma$-rays is hence important for such studies. The observation of the spectacular outbursts of \\pks\\ in July 2006 represents one the most extreme examples of AGN variability in the TeV domain, and allows spectral and timing properties to be probed over two orders of magnitude in flux. Whereas for the flaring states with fluxes above a few $10^{-10}\\,\\percmsqrs$ a clear hardening of the spectrum with increasing flux is observed, familiar also from the blazars Mrk 421 and Mrk 501, for the quiescent state in contrast an indication of a softening is noted. If confirmed, this is a new and intruiging observation in the VHE regime of blazars. The blazar PKS~0208$-$512 (of the FSRQ class) also shows such initial softening and subsequent hardening with flux in the MeV range, but no general trend could be found for $\\gamma$-ray blazars (\\cite{nan07}). In the framework of synchrotron self-Compton scenarios, VHE spectral softening with increasing flux can be associated with, for example, an increase in magnetic field intensity, emission region size, or the power law index of the underlying electron distribution, keeping all other parameters constant. A spectral hardening can equally be obtained by increasing the maximal Lorentz factor of the electron distribution or the Doppler factor (see e.g. Fig.~11.7 in \\cite{kat99}). A better understanding of the mechanisms at play would require multi-wavelength observations of similar time span and sampling density as the data set presented here. It is shown that the variability time scale $t_{\\rm var}$ of a few minutes are only upper limits for the intrinsic lowest characteristic time scale. Doppler factors of $\\delta \\geq 100$ of the emission region are derived by \\cite{HESS2155_BigFlare}~(2007a) using the $\\sim$~$10^9~M_\\odot$ black hole (BH) Schwarzschild radius light crossing time as a limit, while \\cite{begelman} argue that such fast time scales cannot be linked to the size of the BH and must occur in regions of smaller scales, such as ``needles'' of matter moving faster than average within a larger jet (\\cite{ghi08}), small components in the jet dominating at TeV energies (\\cite{kat08}), or jet ``stratification'' (\\cite{bou08}). \\cite{levinson} attributes the variability to dissipation in the jet coming from radiative deceleration of shells with high Lorentz factors. The flaring period allowed the study of light curves in separated energy bands and the derivation of a power law dependence of $F_{\\rm var}$ with the energy ($F_{\\rm var}\\propto E^{\\sim 0.2}$). This dependence is comparable to that reported in \\cite{berrie}~(2007), \\cite{integr_mrk}~(2008), \\cite{xmm}~(2002), where $F_{\\rm var}(E)\\propto E^{\\sim0.2}$ between the optical and X-ray energy bands was found for Mrk\\,421 and \\pks, respectively. An increase with the energy of the flux variability has been found for Mrk\\,501 (\\cite{magic_mrk501}~2007) in VHE $\\gamma$-rays on timescales comparable to those observed here. The flaring period showed for the first time that the intrinsic variability of \\pks\\ increases with the flux, which can itself be described by a lognormal process, indicating that the aperiodic variability of \\pks\\ could be produced by a multiplicative process. The flux in the ``quiescent regime'', which is on average 50 times lower than in the flaring period and has a 3 times lower $F_{\\rm var}$, also follows a lognormal distribution, suggesting similarities between these two regimes. It has been possible to characterize a power spectral density of the flaring period in the frequency range $10^{-4}$--$10^{-2}\\,{\\rm Hz}$, resulting in a power law of index $\\alpha=2.06\\pm0.21$ valid for frequencies down to $\\sim 1/{\\rm day}$. The description of the rapid variability of a TeV blazar as a random stationary process must be taken into account by time-dependent blazar models. For \\pks~the evidence of this log-normality has been found very recently in X-rays (\\cite{pks_xrays}) and as previously mentioned, X-ray binaries and Seyfert galaxies also show lognormal variability, which is thought to originate from the accretion disk (\\cite{mchardy04,lyubarskii,arevalo}), suggesting a connection between the disk and the jet. This variability behaviour should therefore be searched for in existing blazar light curves, independently of the observed wavelength." }, "1005/1005.1707_arXiv.txt": { "abstract": "The evolution of primordial collapsing clouds and formation of proto-Population III stars are investigated using three-dimensional ideal MHD simulations. We calculated the collapse of magnetized primordial clouds from the prestellar stage until the epoch after the proto-Population III star formation, spatially resolving both parsec-scale clouds and sub-AU scale protostars. The formation process of proto-population III star is characterized by the ratio of rotational to magnetic energy of the natal cloud. When the rotational energy is larger than the magnetic energy, fragmentation occurs in the collapsing primordial cloud before the proto-Population III star formation and binary or multiple system appears. Instead, when the magnetic energy is larger than the rotational energy, strong jet with $>100\\,{\\rm km\\,s}^{-1}$ is driven by circumstellar disk around the proto-population III star without fragmentation. Thus, even in the early universe, the magnetic field plays an important role in the star formation process. ", "introduction": "Magnetic fields is a key ingredient in present-day star formation. For example, protostellar jets, which are ubiquitous in star-forming regions, are considered to be driven from protostars by the Lorentz force. Protostellar jets influence gas accretion onto protostars and disturb the ambient medium. In addition, the angular momentum of the cloud is removed by magnetic braking and protostellar jets. The removal of angular momentum makes protostar formation possible in a parent cloud that has a much larger specific angular momentum than the protostar. So far, magnetic effects in primordial gas clouds have been ignored in many studies because magnetic fields in the early universe are supposed to be extremely weak. However, recent studies indicate magnetic fields of moderate strength can exist even in the early universe. Cosmological fluctuations produced magnetic fields before the epoch of recombination \\citep{ichiki06}. These fields were sufficiently large to seed the magnetic fields in galaxies. A generation mechanism for magnetic fields at the epoch of reionization are also proposed \\citep{langer03}, in which magnetic fields in intergalactic matter are amplified up to $\\sim 10^{-11}\\ {\\rm G}$. These fields can therefore increase up to $\\sim 10^{-7}-10^{-8}\\ {\\rm G}$ in the first collapsed object having number density of $n \\sim 10^3 \\cm$. These fields may influence the evolution of primordial gas clouds and formation of Population III stars. Under spherical symmetry including hydrodynamical radiative transfer, many authors have carefully investigated both the present-day \\citep[e.g.,][]{masunaga00} and primordial \\citep[e.g.,][]{omukai98} star formation processes. A significant difference between present-day and primordial star formation exists in the thermal evolution of the collapsing gas cloud because of differences in the abundance of dust grains and metals. In present-day star formation, the gas temperature in molecular clouds is $\\sim10$\\,K. These clouds collapse isothermally for $\\nc \\lesssim 10^{11}\\cm$. Then the gas becomes adiabatic at $\\nc \\simeq 10^{11}\\cm$, and an adiabatic core (or the first core) forms. After the dissociation of molecular hydrogen ($\\nc \\simeq 10^{16}\\cm$), the protostar forms at $\\nc \\simeq 10^{21}\\cm$. On the other hand, primordial gas clouds have temperatures of $\\sim200-300$\\,K at $\\nc \\simeq 10^3\\cm$ \\citep{omukai05,bromm02}. These clouds collapse keeping polytropic index $\\gamma\\simeq 1.1$ for a long range of $10^4\\cm \\lesssim \\nc \\lesssim 10^{16}\\cm$. Thus, the first core does not appear in the primordial collapsing cloud. After the central density reaches $\\nc \\simeq 10^{16}\\cm$, the thermal evolution of the primordial collapsing cloud begins to coincide with that of a present-day cloud \\citep{omukai05}. The difference in thermal evolution between present-day and primordial clouds arises when $\\nc \\lesssim 10^{16}\\cm$. Another major difference between present-day and primordial star formation exists in their magnetic evolution. In present-day star formation, neutral gas is well-coupled with ions for $\\nc \\lesssim 10^{12}\\cm$ and $\\nc \\gtrsim 10^{15}\\cm$, while the magnetic field dissipates by Ohmic dissipation in the range of $10^{12}\\cm \\lesssim \\nc \\lesssim 10^{15}\\cm$ \\citep{nakano02}. In the collapsing cloud, $\\sim99$\\% of the magnetic field is dissipated for $10^{12}\\cm \\lesssim \\nc \\lesssim 10^{15}\\cm$ \\citep{machida07}. On the other hand, in a primordial gas cloud, the magnetic field couples strongly with the primordial gas during all phases of the star formation, as long as the initial field strength is weaker than $B_0\\lesssim 10^{-5}(n/10^3\\cm)^{0.55}$\\,G \\citep{maki04,maki07}. In summary, the magnetic field is largely dissipated by Ohmic dissipation before protostar formation in present-day clouds, while the magnetic field can continue to be amplified without dissipation in primordial clouds. The magnetic field in the collapsing cloud is closely related to the fragmentation or formation of binary or multiple stellar systems. In present-day star formation, the magnetic field strongly suppresses rotation-driven fragmentation \\citep{machida08b}. For primordial clouds, magnetic effects on fragmentation are still unknown. In this study, we investigate the evolution of weakly magnetized primordial clouds and the formation of Population III stars using three-dimensional simulations. ", "conclusions": "In this study, we calculated cloud evolution from the stage of $n_c = 10^3\\cm$ until the protostar is formed ($\\simeq 10^{22}\\cm$) for 36 models, parameterizing the initial magnetic field strength and rotation, to investigate effects of magnetic fields in collapsing primordial clouds. Our calculations showed that fragmentation occurs but no jet appears when $\\beta_0 > \\gamma_0$, and jet appears after the protostar formation without fragmentation when $\\beta_0 < \\gamma_0$. Thus, in the collapsing primordial cloud, the cloud evolution is mainly controlled by the centrifugal force than the Lorentz force when $\\beta_0 > \\gamma_0$, while the Lorenz force is more dominant than the centrifugal force when $\\gamma_0 > \\beta_0$. A jet is driven when the initial cloud has magnetic field of \\begin{equation} B_0 \\gtrsim 10^{-9} \\left( \\dfrac{\\nc}{10^3\\cm} \\right)^{2/3} \\,G \\label{eq:bcrit} \\end{equation} if the cloud rotates slowly as $\\Omega \\lesssim 4\\times 10^{-17}(\\nc/10^3\\cm)^{2/3}$\\,s$^{-1}$. The power of a jet, e.g., a mass ejection rate, is considered to be controlled by the accretion rate as indicated in present-day star formation; the mass ejection rate of a jet is 1/10 of the mass accretion onto the protostar. The accretion rate of primordial star formation is expected to be considerably larger than that at present day, and it produces a stronger jet. The life time of the jet also seems to be controlled by accretion in the present-day; a jet stops when mass accretion stops. For Population III stars, the gas accretion does not halt within their lifetimes \\citep{omukai01,omukai03}. Therefore, a jet also may continue during the all lifetime of the protostar, and the strong jet propagates to disturb a surrounding medium significantly. The disturbance of the medium could trigger the subsequent star formation as frequently observed in present-day star formation. Assuming the power law growth of $B_0 \\propto n_c^{2/3}$, the critical strength of the magnetic field, $B_0 = 10^{-9}$\\,G at $n_c = 10^{3}\\cm$ corresponds to $B_0 = 5\\times 10^{-13}$\\,G at $n_c = 0.01\\cm$ [see, eq.~(\\ref{eq:bcrit})], which is much stronger than the background magnetic field derived by \\citet{ichiki06}. However, when the magnetic field is amplified to $B \\sim 10^{-9}(\\nc/10^3\\cm)^{2/3}$\\,G by some mechanisms \\citep[e.g.,][]{schleicher10}, the magnetic field can affect the collapse of the primordial cloud. Even if a cloud has a magnetic field weaker than the critical strength $B_0 = 10^{-9}$\\,G, the magnetic field may play an important role after the protostar formation. In analytical study of the evolution of accretion disks around the first stars \\citep{tan04} , it is suggested that magnetic fields amplified in the circumstellar disk eventually give rise to protostellar jets during the protostellar accretion phase. Rotation promotes fragmentation when the first collapsed objects has the angular velocity of $\\Omega_0 \\gtrsim 10^{-17}(\\nc/10^3\\cm)^{2/3}$\\,s$^{-1}$. The fragmentation is expected to produces binary or multiple stellar system. When a multiple stellar system is formed, some stars can be ejected by close encounters. At the protostar formation epoch, the protostar has a mass of $M\\simeq 10^{-3}\\msun$. The ejected proto-Population III stars may evolve to metal-free brown dwarfs or low-mass stars. When a binary component in a multiple stellar system is ejected from the parent cloud by protostellar interaction, a low-mass metal free binary may also appear in the early universe. It is considered that the extremely metal-poor ([Fe/H]$<$-5) stars \\citep{christlieb01,frebel05} are formed as binary members from metal-free gas, and then have been polluted by the companion stars during the stellar evolution \\citep{suda04}. In addition, a binary frequency in Population III star may be comparable to or larger than that at present day \\citep{komiya06,machida08c,machida08d,machida09a,machida09b}. In order to confirm the ejection scenario, the further long-term calculations are required." }, "1005/1005.1641_arXiv.txt": { "abstract": "In the context of the VLA-COSMOS Deep project additional VLA A array observations at 1.4\\,GHz were obtained for the central degree of the COSMOS field and combined with the existing data from the VLA-COSMOS Large project. A newly constructed Deep mosaic with a resolution of 2.5$''$ was used to search for sources down to 4$\\sigma$ with 1$\\sigma \\approx 12\\,\\mu$Jy/beam in the central 50$'\\times$50$'$. This new catalog is combined with the catalog from the Large project (obtained at 1.5$''$$\\times$1.4$''$ resolution) to construct a new Joint catalog. All sources listed in the new Joint catalog have peak flux densities of $\\ge$5$\\sigma$ at 1.5$''$ and/or 2.5$''$ resolution to account for the fact that a significant fraction of sources at these low flux levels are expected to be slighty resolved at 1.5$''$ resolution. All properties listed in the Joint catalog such as peak flux density, integrated flux density and source size are determined in the 2.5$''$ resolution Deep image. In addition, the Joint catalog contains 43 newly identified multi-component sources. ", "introduction": "In recent years, several cosmological deep fields have been imaged at 20\\,cm \\citep[e.g., ][]{ric00,bon03,con03,hop03,sey04,nor05,huy05,fom06,sim06,ivi07,sch07,mil08,owe08} providing a few thousand radio sources down to flux limits of a few 10\\,$\\mu$Jy. These deep radio imaging data are sensitive enough to detect star forming galaxies with star formation rates of several 10 to 100 $\\solmy$ out to and beyond a redshift of $z$\\,$\\sim$\\,1. Similarly, radio galaxies can be seen out to redshifts of $z$\\,$\\sim$\\,5 and the most luminous ones even well into the epoch of reionization. Thus, deep radio images in conjunction with deep imaging data at X-ray, optical and infrared wavelengths are ideal to investigate the dust-unbiased star formation, the evolution of radio(-loud) AGN, as well as the population mix of radio sources in the first place. In order to study the cosmological evolution of galaxies and black holes, it is not only important to overcome the effect of cosmic variance (e.g. by studying a large enough area) but also to understand the effect of large scale structure on the evolution (e.g. by covering a large contiguous area). To address the second effect in particular, the Cosmic Evolution Survey (COSMOS)\\footnote{~\\tt http://cosmos.astro.caltech.edu} collaboration has conducted panchromatic imaging and spectroscopy of an equatorial field with a size of 2\\,$\\rm deg^2$ \\citep[for an overview, see ][]{sco07a} ranging from X-ray XMM-Newton and Chandra \\citep{has07,elv09}, UV GALEX \\citep{zam07}, optical and near-infrared ground-based \\citep{tan07,cap07a}, optical HST \\citep{sco07b,koe07}, mid- to far-infrared Spitzer \\citep{san07}, millimeter \\citep{ber07,sco08} and radio VLA \\citep{sch04,sch07} imaging to extensive optical spectroscopy using the VLT/VIMOS and Magellan/IMACS instruments \\citep{lil07,tru07}. Most of these datasets are now publicly available from the COSMOS archive at IPAC/IRSA\\footnote{~\\tt http://irsa.ipac.caltech.edu/Missions/cosmos.html}. The VLA-COSMOS survey at 20\\,cm is part of the overall imaging effort and its scientific goals and motivation have been described in detail by \\cite{sch07}. Initial observations from a pilot project testing the mosaicking strategy and giving a first source catalog are presented by \\cite{sch04}. As a large NRAO/VLA program, the VLA was used in A and C configuration to cover the entire COSMOS field resulting in an image with uniform noise properties in the central 1$\\times$1\\,deg$^2$ and an average rms of 10.5\\,$\\mu$Jy. \\cite{sch07} provide a detailed description of the survey set-up, the data reduction, as well as the testing and construction of the final VLA-COSMOS Large project catalog (hereafter: Large catalog). Subsequently, \\cite{bon08} derived the completeness of the Large catalog and also analyzed the effect of bandwidth smearing on the derived source flux densities to obtain the source counts. Although the VLA-COSMOS Large Project dataset has been used for several scientific results on, e.g. the faint radio population, the radio-derived star formation rate density, the radio AGN population and stacking of high-z galaxy populations \\citep{smo08,smo09a,smo09b,car07,car08}, the need for deeper radio imaging data became apparent during the search for radio counterparts to millimeter sources from the COSMOS MAMBO mapping data \\citep{ber07}; the Large project only provided counterparts for about half of the mm-sources. Thus, the Deep project was initiated with the aim of doubling the integration time for the central seven pointings which fully cover the MAMBO 1.2\\,mm map of the COSMOS field. These new observations are described here. The paper is organized as follows: After a description of the new observations and the data reduction (\\S \\ref{sec:observations}), the revision of the Large catalog (leading to a new version of v2.0) and the construction of the new Deep catalog are outlined in \\S \\ref{sec:newLarge} and \\S \\ref{sec:deepcat}. In \\S \\ref{sec:identify} we explain the construction of the final Joint catalog which is described in detail in \\S \\ref{sec:joint_cat} where we also present all the tests and corrections involved. A summary is given in \\S \\ref{sec:summary}. ", "conclusions": "\\label{sec:summary} Continued analysis of the VLA-COSMOS catalog presented in \\cite{sch07} and the completion of the VLA-COSMOS Deep project motivated the compilation of a new radio catalog for the COSMOS field. The VLA-COSMOS Joint catalog was generated by combining the catalogs of the VLA-COSMOS Large Project with a newly created source catalog (Deep catalog) from the 2.5$''$ resolution Deep mosaic. This catalog is already available for download by the public at the COSMOS archive at IPAC/IRSA\\footnote{~~\\texttt{http://irsa.ipac.caltech.edu/data/COSMOS/tables/vla}}. A comparison of the depth and areal coverage for a representative sample of deep field radio surveys at 1.4\\,GHz shows that the VLA-COSMOS covers the largest area at its depth and angular resolution (Fig. \\ref{fig:radio_surveys}). Thus it should be well suited to also study effects of the Large Scale Structure on the presence/absence of radio emission. The reduction and analysis of the deeper 20\\,cm observations of the central 7 pointings of the VLA-COSMOS projects using the VLA in A configuration have been described in detail (also referred to as VLA-COSMOS Deep project). In order to minimize the effect of bandwidth smearing the Deep mosaic has a resolution of 2.5$''$ (compared to 1.5$''$ for the Large mosaic) and it was used to create a corresponding source catalog using the task \\texttt{SAD}. An input list for the new Joint catalog was compiled by combining the revised Large catalog (v2.0) and the new Deep catalog. The criteria were set such that no particular bias against slightly extended radio sources was present when selecting the sources. All properties of the radio sources listed in the Joint catalog have been derived in the 2.5$''$ resolution Deep mosaic. The construction of the Joint catalog was motivated by the desire to provide a catalog of bona-fide radio sources in the COSMOS field for distinct science applications that are interested in the radio properties of certain populations of galaxies. On the other hand the revised Large catalog (v2.0) is flux-limited (in radio), has a fairly uniform sensitivity coverage and its completeness is well characterized \\citep[see][]{bon08}, thus it is well suited for, e.g., studies of the faint radio population \\citep[such as, e.g.,][]{smo08,smo09a,smo09b}. \\appendix" }, "1005/1005.4855_arXiv.txt": { "abstract": "In this paper we present in detail the methodology and the first results of a ground-based program to determine the absolute proper motion of the Fornax dwarf spheroidal galaxy. The proper motion was determined using bona-fide Fornax star members measured with respect to a fiducial at-rest background spectroscopically confirmed Quasar, \\qso. Our homogeneous measurements, based on this one Quasar gives a value of (\\mua,\\mud)$ = (0.64 \\pm 0.08, -0.01 \\pm 0.11)$~\\masy. There are only two other (astrometric) determinations for the transverse motion of Fornax: one based on a combination of plates and HST data, and another (of higher internal precision) based on HST data. We show that our proper motion errors are similar to those derived from HST measurements on individual QSOs. We provide evidence that, as far as we can determine it, our motion is not affected by magnitude, color, or other potential systematic effects. Last epoch measurements and reductions are underway for other four Quasar fields of this galaxy, which, when combined, should yield proper motions with a weighted mean error of $\\sim50\\,\\mu$as~y$^{-1}$, allowing us to place important constraints on the orbit of Fornax. ", "introduction": "The proper motions (PMs) of the satellites of the Milky Way (MW), when combined with existing radial velocities, allow us to determine the space velocity vectors of these satellites of our galaxy, which in turn place important constraints on their orbits (see, e.g., \\citet{bes07}, \\citet{pia07}). This knowledge is crucial to determine if these galaxies are gravitationally bound to the Galaxy, and to our understanding of the evolution and origin of its satellite system (\\citet{byr94}). The PMs of the satellites of the MW are necessary to understand: a) the origin of the MW satellite system and its relationship with the formation and evolution of the galactic halo (\\citet{din00}, \\citet{pal02}, \\citet{car09}), b) the nature and origin of the streams that seem to align different subgroups of these galaxies (\\citet{lyn82}, \\citet{lyn95}, \\citet{pia05}), and c) the role of tidal interactions in the evolution and star formation history of low mass galaxies (\\citet{zar04}, \\citet{pia05}, \\citet{noe09}, \\citet{may10}). A comprehensive study of the MW's satellite system can lead to a greater general understanding of galaxy evolution and the physical processes governing star formation in galaxies. From another perspective, reliable space motions of Local Group galaxies is a key ingredient to populate the phase-space components for flow models which predict the dynamical evolution of the local universe. Indeed, current center-of-mass locations and motions from the present distribution of galaxies can be used as boundary conditions to ``trace them back in time'' \\citep{pee89,pee94} and to test the paradigm that galaxy clusters grew by gravitational instabilities from an originally smoother medium, with small random motions. These earlier simulations have not been repeated too often due to the growing realization that more complex effects such as tides, satellite interactions, accretion and mergers, influencing the growth of galaxies with time (specially at epochs earlier than 7-8~Gyr from now), can greatly complicate this approach. Nevertheless, the motions of local group galaxies can be used as boundary conditions in a first aproximation to carry out the above analysis. This goal has become one of the important drivers behind future astrometric space missions, such as SIM (\\citet{unw08}, specially their Figure~13), which expect to measure PM for Local Group galaxies with a precision ten times better than what can be achieved with present techniques. With the above motivations in mind, in the year 2000 we started a ground-based program aimed at determining, the absolute PM of three southern dwarf Spheroidal (dSph) galaxies, Carina, Fornax, and Sculptor, with respect to known background Quasars (QSOs) that can be used as inertial reference points. Three epochs, over a period of eight years were obtained using a single telescope+detector set up: Ours is then the first entirely optical CCD/ground-based proper motion study of an external galaxy other than the Magellanic Clouds. In this paper we report on the first results from this program, based on one QSO field in Fornax, for which we have data of good enough quality to allow us to asses the expected precision of our measurements, and to describe our methodology in detail. Last epoch measurements and reductions are underway for other four Quasar fields of Fornax, as well as for a similar number of QSO fields in Carina and Sculptor. The results for these will be presented in forthcoming papers. In section~\\ref{obma} we describe our observational material and data acquisition strategies, in section~\\ref{rest} we describe our methodology for deriving the PMs, in section~\\ref{anal} we present the analysis for our QSO field, and in section~\\ref{comp} we present our main conclusions as well as a comparison to other results. \\newpage ", "conclusions": "\\label{comp} There have been only two astrometric determinations of the PM for the Fornax dSph galaxy, namely that by \\citet{din04}, based on a combination of ground-based plates and Hubble-WFPC data, and that based exclusively on HST data (\\citet{pia07}, which gives revised values to those reported earlier in \\citet{pia02}). \\citet{din04}, based on an independent determination using 48 galaxies and 8 QSOs give a weighted mean of (\\mua,\\mud)=($0.59 \\pm 0.16, -0.15 \\pm 0.16$)~\\masy. On the other hand, \\citet{pia07}, based on 4 QSOs (their Table~3), gives a weighted mean of (\\mua,\\mud)=($0.476 \\pm 0.046, -0.360 \\pm 0.041$)~\\masy. In order to compare our result (based on a single QSO) with their measurements it makes more sense to compare our value with their {\\it individual} PM values. This allows us to estimate the expected uncertainty on our final weighted mean when we incorporate our other 4 QSO fields. In Figure~\\ref{pmcomp} we plot the individual measurements from these previous works, along with our measurement. We note however that, while \\citet{pia07} report individual measurements, \\citet{din04} only give the mean values with respect to 8 QSOs and 48 galaxies, and these are the results plotted in Figure~\\ref{pmcomp}. \\citet{din04} {\\it do} give the {\\it individual} PMs for images A and B of the same QSO reported in this paper. These individual values are also plotted in Figure~\\ref{pmcomp}. By looking at the open and filled squares we can see the improvement achieved by ground-based astrometry when using CCDs and a homogeneous data set in comparison with the results that combine non-linear plates and heterogeneous data. Nevertheless, it is somewhat surprising how close our (single) value is to the mean PM from galaxies as derived by \\cite{din04} (differences smaller than $1\\sigma$ of {\\it our} smaller error), while there is a larger discrepancy (although still smaller than $2\\sigma$ of {\\it their} (larger) error) when compared to the PM using QSOs. In Dinescu's study they have a large sample of brighter better-measured galaxies (see their Figure~2) while the (much fewer) QSOs are fainter and possibly have larger positional errors (although, their individual PM errors are smaller than those of galaxies for a given magnitude all the way down to V$\\sim 21$, possibly due to the extended, more diffuse nature of galaxies). Overall, their mean PM with respect to galaxies has an error $\\sim50$\\% smaller than that derived from the QSOs. In comparison with the individual HST measurements, we see a rather large discrepancy, specially in DEC: While our value lies between 2.3 and 1.2$\\sigma$ away from HST individual PMs in \\mua, this difference becomes between 4.4 and 2.4$\\sigma$ in \\mud. We note that the largest \\& smallest PM values from HST for {\\it both} \\mua\\ and \\mud\\ (see Table~3 on \\citet{pia07}) come precisely from images A and B of \\qso. \\citet{din04} also point the rather large discrepancy in the PMs derived from components A and B (plotted in our Figure~\\ref{pmcomp}). As was described in section~\\ref{centroids} the A component of QJ~0240-3434 {\\it is} indeed affected by two close companions (see Figure~\\ref{qsoab}). It is not unlikely that tiny changes in the photocenter of the images due to other close, unresolved and fainter, companions, or even slightly extended structure(s) in the wings of the QSO (from the underlying QSO galactic disk) could introduce an extra source of noise in the PMs that is not necessarily accounted for in the final PM error budget. Obviously, whether or not our own measurements are subject to some (as yet unknown) systematic effect related to what we just mentioned, is an issue that could be addressed once we incorporate the other QSO fields. In Table~\\ref{pmtab} we summarize the measurements plotted in Figure~\\ref{pmcomp} (the ``perspective'' motions are explained further below). It is interesting to ascertain to what type of final (Heliocentric) velocity errors correspond our individual PM measurement errors. For a given distance and distance error $r \\pm \\sigma_r$ and PM and its error in component ``x'' (in this case either RA or DEC), $\\mu_x \\pm \\sigma_{\\mu_x}$, the corresponding velocity and its error is given by: \\begin{eqnarray} v_x = K \\, r \\, \\mu_x \\label{vel} \\\\ \\nonumber \\\\ \\frac{\\sigma_{v_x}}{v_x} = \\sqrt{ \\left( \\frac{\\sigma_r}{r} \\right)^2 + \\left( \\frac{\\sigma_{\\mu_x}}{\\mu_x} \\right)^2 } \\label{velerr} \\end{eqnarray} where, if $r$ is in kpc, and $\\mu_x$ is in \\masy, then $K=4.74$ and $v_x$ and $\\sigma_{v_x}$ are in km~s$^{-1}$. According to \\citet{mat98}, the distance to Fornax is $138 \\pm 8$~kpc (i.e., 6\\% formal error). Therefore in equation~(\\ref{velerr}) the distance error is negligible in comparison with our PM errors in either RA or DEC. For our PM reference value (see Table~\\ref{pmtab}) we obtain a Heliocentric velocity and error of $v_{\\alpha}=419 \\pm 58$~km~s$^{-1}$ and $v_{\\delta}= -7 \\pm 72$~km~s$^{-1}$. The measured Heliocentric radial velocity from \\citep{mat98} is $v_r = 53 \\pm 3$~km~s$^{-1}$. It is thus clear that the derived Heliocentric Fornax motion is dominated by uncertainties in the PM from our one-QSO measurement, not by the distance or radial velocity uncertainty. Given our measurement errors, once we incorporate our other four QSO fields, we expect to achieve velocity errors for the weighted mean on the order of 30~km~s$^{-1}$ per component, totally compatible with the (weighted mean) HST results. Table~\\ref{pmtab} also shows that our proper motion value exhibits the largest tangential velocity for Fornax after the Dinescu et. al value from Galaxies. At this stage we howevere refrain from making a full analysis of the implications of this result, until we have collected the data from the other four QSO fields which will allow us to have a more robust weighted mean. We must note that, when comparing PMs values derived for an extended object, such as Fornax, one must actually compare ``center-of-mass'' (COM) motions to avoid any projection effects and internal galaxy motions (e.g., galactic rotation) that might alter the observed motions. For computing galactic orbits of these galaxies we also need COM velocities. Our measurements are performed, however, on fields away from the COM, and we have to apply corrections to account for this situation. As explained in \\citet{cos09}, in the case of the LMC our PMs had to be corrected not only for projection effects (see below), but also by the effect on our measured PMs induced by the (differential) rotation of the plane of the LMC. To apply these corrections, we followed the prescriptions described by \\cite{jon94} and \\cite{van02}, which assumed that our fields lay in the plane of the LMC (or SMC) and shared the motion of its disk. The situation for the dSphs is quite different however, they do not exhibit any hint of large-scale rotation, nor of the existence of a disk or other well-defined structure; rather they show a smooth spheroidal distribution and the motion of its stars is dominated by their velocity dispersion, which is typically a few~km~s$^{-1}$ \\citep{van99}. This velocity dispersion translates into a PM dispersion of a few $\\mu$as~y$^{-1}$ which is not measurable by current astrometric techniques (although it will be measured in the future by SIM). Therefore, if we assume that there are no large-scale streaming motions in Fornax, our PMs are not affected, as far as we can measure it, by internal kinematic effects. We do, however, need to correct for purely geometrical projection effects, this is done as described in the following paragraph. If $(v_r,v_\\alpha,v_\\delta)$ are the observed (measured) Heliocentric radial velocity, velocity in RA and velocity in DEC respectively, for a field at position $(\\alpha,\\delta)$ and Heliocentric distance $r$, the velocity for the COM $(v^*_r,v^*_\\alpha,v^*_\\delta)$ at position $(\\alpha^*,\\delta^*)$ and distance $r^*$ is given by: \\begin{equation} \\left( \\begin{array}{c} v^*_r \\\\ v^*_\\alpha \\\\ v^*_\\delta \\end{array} \\right) = M^{-1} (\\alpha^*,\\delta^*) \\cdot M(\\alpha,\\delta) \\cdot \\left( \\begin{array}{c} v_r \\\\ v_\\alpha \\\\ v_\\delta \\end{array} \\right) \\label{proj} \\end{equation} where $M(\\alpha,\\delta)$ is a rotation matrix whose components are: \\begin{equation} M(\\alpha,\\delta) = \\left( \\begin{array}{ccc} \\cos \\delta \\cos \\alpha & -\\sin \\alpha & -\\sin \\delta \\cos \\alpha \\\\ \\cos \\delta \\sin \\alpha & \\cos \\alpha & -\\sin \\delta \\sin \\alpha \\\\ \\sin \\delta & 0 & \\cos \\delta \\end{array} \\right) \\label{matrix} \\end{equation} and which satisfies that $M^{-1}(\\alpha,\\delta)=M^t(\\alpha,\\delta)$. While $r$ and $r^*$ are not explicitely written in equation~(\\ref{proj}), they are implicitly used to go from PMs to tangential velocities (or vice versa) through equation~(\\ref{vel}). Of course, $r^* = 138 \\pm 8$~kpc. In the case of the LMC (and SMC) \\citet{cos09} assumed that our fields were located in a disk-like structure with known orientation in the sky (inclination and line of nodes). We do not have however such structures in the featureless dSphs. If we assume, e.g., as a first crude approximation that the galaxy lies in a plane in the sky perpendicular to the line-of-sight to the COM. In this case one can show that $r$ is given by: \\begin{equation} r = \\frac{r^*}{\\cos \\delta \\cos \\alpha \\cos \\delta^* \\cos \\alpha^* + \\cos \\delta \\sin \\alpha \\cos \\delta^* \\sin \\alpha^* + \\sin \\delta \\sin \\delta^*} \\end{equation} We note that, in equation~(\\ref{proj}) $v_r$ is {\\it not} known, but we do know $v^*_r = 53 \\pm 3$~km~s$^{-1}$ from the literature. Therefore, starting from our measured $v_\\alpha$ and $v_\\delta$ values we iterate on the $v_r$ values until we reproduce the expected $v^*_r$ for the COM, in a procedure similar to that adopted by \\citet{cos09}. Also, equations (\\ref{proj}) and (\\ref{matrix}) easily allow us to fully propagate errors on all the measured quantities (radial velocities, proper motions, distances) from the observed values to the sought-for COM values. For our measured PM, the COM distance and radial velocity indicated previously, and the Fornax COM position given by \\citet{mat98}, we obtain $v^*_{\\alpha}=419 \\pm 52$~km~s$^{-1}$ and $v^*_{\\delta}= -7 \\pm 72$~km~s$^{-1}$, i.e., a value similar to that computed without {\\it any} projection correction. This is due to the fact that, in this particular case, these corrections are actually quite small, the PM corrections are 0.05~$\\mu$as in RA and by 0.45~$\\mu$as in DEC. These corrections are, of course, much smaller than the measurement uncertainties involved (this was not case of the LMC and SMC fields reported in \\citet{cos09}). Note that the field reported here is quite close to the Fornax COM, for some of the other more distant QSO fields in our program, these corrections might be slightly larger, but can be readily computed in each case. None of the values reported by other authors in Table~\\ref{pmtab} have been corrected for any projection effect, but since these corrections are tiny, one can readily compare them without further corrections. For completeness, in Table~\\ref{pmtab} (and in Figure~\\ref{pmcomp}), we have included the recent determination of the Fornax PM by \\citet{wal08}, using what they call the ``perspective rotation'' method (in a way, this method is the reverse of the ``astrometric radial velocities'', described by, e.g., \\citet{lin00}). As it can be easily seen from Equations~\\ref{proj} and~\\ref{matrix}, one could write an equivalent equation for $(v_r, v_\\alpha, v_\\delta)$ as a function of $(v^*_r, v^*_\\alpha, v^*_\\delta)$ by a simple matrix inversion. The first row of that equation would give the observed radial velocity $v_r$ as a function of a combination of $(v^*_r, v^*_\\alpha, v^*_\\delta)$ (which, for a given galaxy is of course a fixed quantity - independent of the field observed). Therefore, if one measures the $v_r$ of samples of stars at different locations $(\\alpha,\\delta)$ across a galaxy, one can solve through some minimization algorithm for the unknown $(v^*_r, v^*_\\alpha, v^*_\\delta)$. \\citet{wal08} have used precisely this approach to determine the ``perspective'' PMs for 4 dSphs, including Fornax, using radial velocities exclusively. As it can be seen from Table~\\ref{pmtab}, the method has errors compatible with the best purely astrometric determinations, and can thus become a useful complement to them. \\newpage" }, "1005/1005.3472_arXiv.txt": { "abstract": "{We present far-infrared spectroscopic observations, taken with the Photodetector Array Camera and Spectrometer (PACS) on the \\emph{Herschel Space Observatory}, of the protoplanetary disk around the pre--main--sequence star HD\\,100546. These observations are the first within the DIGIT \\emph{Herschel} key program, which aims to follow the evolution of dust, ice, and gas from young stellar objects still embedded in their parental molecular cloud core, through the final pre--main--sequence phases when the circumstellar disks are dissipated.} {Our aim is to improve the constraints on temperature and chemical composition of the crystalline olivines in the disk of HD\\,100546 and to give an inventory of the gas lines present in its far-infrared spectrum. } {The 69\\,$\\mu$m feature is analyzed in terms of position and shape to derive the dust temperature and composition. Furthermore, we detected 32 emission lines from five gaseous species and measured their line fluxes. } {The 69\\,$\\mu$m emission comes either from dust grains with $\\sim$70~K at radii larger than 50\\,AU, as suggested by blackbody fitting, or it arises from $\\sim$200\\,K dust at $\\sim$13\\,AU, close to the midplane, as supported by radiative transfer models. We also conclude that the forsterite crystals have few defects and contain at most a few percent iron by mass. Forbidden line emission from [C~{\\sc ii}] at 157\\,$\\mu$m and [O~{\\sc i}] at 63 and 145\\,$\\mu$m, most likely due to photodissociation by stellar photons, is detected. Furthermore, five H$_2$O and several OH lines are detected. We also found high-J rotational transition lines of CO, with rotational temperatures of $\\sim$300\\,K for the transitions up to $J=22-21$ and $T\\,\\sim$800\\,K for higher transitions. } {} ", "introduction": "Circumstellar disks around young stars are the birthplaces of planetary systems. To understand planet formation, it is vital to study the processes that govern the evolution of gas and dust in these disks. PACS provides unique information in this field through observations of far-infrared (IR) solid-state features which are particularly sensitive to temperature and elemental composition. Moreover, PACS is well suited to study the warm gas of a few 100\\,K, complementary to the hot gas probed in near-IR and the cold gas observed at millimeter wavelengths. An intensively studied pre-main-sequence star is the Herbig B9.5Vne star HD\\,100546. While signs of ongoing accretion indicate the star's infancy \\citep{Deleuil2004}, the estimated age of 10\\,Myr makes it unusually old for a star with a disk \\citep{vdAncker1997}. HD\\,100546 is nearby (103 pc) and optical/near-IR scattered light imaging has revealed a wealth of structures in the disk \\citep[e.g.,][]{Grady2001, Augereau2001, Ardila2007}. Based on the mid-- to near-IR excess ratio, \\citet{Bouwman2003} suggested an inner cavity in the disk. Later observations confirmed this gap \\citep[e.g.,][]{Grady2005, Benisty2010}. This remarkable feature and the longevity of the HD\\,100546 disk may point to a young planet in the inner 10 AU, making it a prime target for detailed studies. The system of HD\\,100546 has a rich 2.4 -- 180~$\\mu$m spectrum as observed with ISO \\citep{Malfait1998}, showing a striking similarity with that of the comet Hale-Bopp \\citep{Crovisier1997}. \\citet{Juhasz2010} investigated the \\emph{Spitzer} spectrum of the source. Strong forsterite emission is observed in both ISO and \\emph{Spitzer} data. A continuum fit indicated crystalline dust at 210\\,K and 50\\,K \\citep{Malfait1998}. The ISO spectrum shows a strong line of [O~{\\sc i}] 63\\,$\\mu$m and a weaker [C~{\\sc ii}] 158\\,$\\mu$m line. In the near-IR, \\citet{Brittain2009} and \\citet{vdPlas2009} observed CO ro--vibrational lines with a rotational temperature of $\\sim$1000\\,K and found no CO gas in the inner dust cavity. Observations of optical [O {\\sc i}] and Balmer line emission, on the other hand, demonstrate that the inner disk is not completely devoid of gas \\citep{AckevdAncker2006}. At submillimetre wavelengths, \\citet{Panic2010} have detected pure rotational lines of CO up to $J=7-6$ probing a warm (60--70\\,K) layer in the outer disk ($\\sim$\\,100\\,AU). Information on the intermediate temperatures is still lacking. Here we present an analysis of narrow features (about 1\\,$\\mu$m) in the far-IR spectrum of HD\\,100546 over the full PACS range that was obtained within the \\emph{Herschel} key program `Dust, Ice and Gas in Time' (DIGIT). ", "conclusions": "\\label{sec:discussion} The PACS data provide the strongest constraints yet on the composition of olivines in a protoplanetary disk. We found the olivines to be extremely iron-poor (less than 3--4\\% iron). Our observations can be modeled with dust at $\\sim$70\\,K, consistent with ISO results, but requiring at least 2\\% iron. On the other hand the determination of the temperature of the forsterite based on intrinsic features of the 69\\,$\\mu$m band instead of using a continuum analysis offers a second option: The emission may emerge from pure forsterite at 200\\,K and 13\\,AU close to the midplane. This region is optically thick for shorter wavelengths. The far-infrared spectrum of HD\\,100546 contains a wealth of molecular gas lines including CO, H$_2$O, and OH. We found CO transitions up to $J$=31--30 which arise from gas in the temperature range of 300 to 800 K and appear to sample gas in the surface layers of the disk at temperatures and surface densities intermediate to those probed by submillimeter and near-infrared CO measurements." }, "1005/1005.2762.txt": { "abstract": "\\noindent It has been recognized that the turbulent cross helicity (correlation between the velocity and magnetic-field fluctuations) can play an important role in several magnetohydrodynamic (MHD) plasma phenomena such as the global magnetic-field generation, turbulence suppression, etc. Despite its relevance to the cross-helicity evolution, little attention has been paid to the dissipation rate of the turbulent cross helicity, $\\varepsilon_W$. In this paper, we consider the model expression for the dissipation rate of the turbulent cross helicity. In addition to the algebraic model, an evolution equation of $\\varepsilon_W$ is proposed on the basis of the statistical analytical theory of inhomogeneous turbulence. A turbulence model with the modeling of $\\varepsilon_W$ is applied to the solar-wind turbulence. Numerical results on the large-scale evolution of the cross helicity is compared with the satellite observations. It is shown that, as far as the solar-wind application is concerned, the simplest possible algebraic model for $\\varepsilon_W$ is sufficient for elucidating the large-scale spatial evolution of the solar-wind turbulence. Dependence of the cross-helicity evolution on the large-scale velocity structures such as velocity shear and flow expansion is also discussed. \\vspace{6pt} \\noindent {{\\bf{Keywords:}} Magnetohydrodynamic turbulence; turbulence model; cross helicity; dissipation rate; solar wind; }\\bigskip ", "introduction": "\\label{sec:level1} In the magnetohydrodynamic (MHD) turbulent flow at high magnetic Reynolds number ($Rm \\gg 1$), magnetic fields are considered to be frozen in plasmas, and move with the flow.\\cite{alf1950} In such a flow, the induced magnetic field is often much larger than the originally imposed field. Besides, MHD waves such as the Alfv\\'{e}n wave are considered to exist ubiquitously. The cross helicity, defined by the correlation between the velocity $\\bf{u}$ and magnetic field $\\bf{b}$, is a possible describer of such MHD turbulence properties. Actually, the magnetic-field generation due to the turbulent cross helicity has been investigated.\\cite{yos1990,yos1993,yok1996,yos1998,yos1999,yok1999,yos2000,yos2004} As is well known, the total amount of cross helicity $\\int_V {{\\bf{u}} \\cdot {\\bf{b}}} dV$, as well as that of the MHD energy $\\int_V {({\\bf{u}}^2 + {\\bf{b}}^2) / 2}\\ dV$, is an inviscid invariant of the MHD equations. Because of this conservative property, the turbulent densities of the MHD energy and cross-helicity, $K \\equiv \\langle { {\\bf{u}}' {}^2 + {\\bf{b}}' {}^2 } \\rangle / 2$ and $W \\equiv \\langle { {\\bf{u}}' \\cdot {\\bf{b}}' } \\rangle$, may serve themselves as a good measure for characterizing the statistical properties of MHD turbulence (${\\bf{u}}'$: velocity fluctuation, ${\\bf{b}}'$: magnetic-field fluctuation, $\\langle \\cdots \\rangle$: ensemble average). The evolution equations of $K$ and $W$ are similar in form, and their mathematical structures are quite simple. The evolution of $K$ and $W$ are determined by three constitutes: the production, dissipation, and transport rates. Firstly, the production rate is expressed by the correlation of turbulence fields coupled with the mean-field inhomogeneity. We note that the production rate of turbulence quantities such as $\\langle {{\\bf{u}}' \\cdot {\\bf{b}}'} \\rangle$ can be expressed exactly in the same as the counterpart of mean-field quantities such as ${\\bf{U}} \\cdot {\\bf{B}}$, but with the opposite sign ($\\bf{U}$: the mean velocity, $\\bf{B}$: the mean magnetic field). This makes our interpretation possible that the drain of mean-field quantity gives rise to the generation of turbulence counterpart. So, the production rate represents how a quantity is supplied to turbulence by way of its cascading process (See Appendix~\\ref{sec:appendixA}). Secondly, the dissipation rates of the turbulent MHD energy and cross helicity, $\\varepsilon$ and $\\varepsilon_W$, whose definitions will be given shortly in Section~\\ref{sec:level2}, represent the effects of molecular viscosity and magnetic diffusivity coupled with the small-scale fluctuations. However, we stress the following point. The dissipation rates of the turbulent MHD energy and cross helicity, $\\varepsilon$ and $\\varepsilon_W$, can be considered from another aspect. In the intermediate range of turbulence, called the inertial range, the energy and cross helicity supplied from the energy-containing range compensate the energy and cross-helicity lost in the dissipation range. For this cascade picture of turbulence, the energy and cross-helicity transfer from lower to higher wavenumber ranges are most important quantities. In an equilibrium turbulence, $\\varepsilon$ and $\\varepsilon_W$ represent these transfer rates of $K$ and $W$, respectively. This makes the construction of the $\\varepsilon$ and $\\varepsilon_W$ equation possible as we show later. Finally, the transport rates express the flux of a quantity that enters the fluid volume through the boundary. The expression for the transport rates suggests in what situation the quantities considered can be supplied to turbulence. Thanks to these clear-cut pictures associated with the evolution equation, the cross helicity (density) $W$, as well as the turbulent MHD energy (density) $K$, may play an important role in the turbulence modeling of MHD fluids. However, as compared with $K$ and other pseudoscalar turbulence quantities such as the turbulence kinetic and magnetic helicities, only a limited attention so far has been paid to the cross helicity. In the context of homogeneous isotropic MHD turbulence, some important investigations have been made on the decaying rate of the cross helicity or $\\varepsilon_W$. It was shown that if there is a prevailed sign of the cross helicity in the initial state, the system goes towards a dynamically aligned state. The cross helicity scaled by the MHD energy grows towards +1 or -1 depending on the initially prevailed sign of the cross helicity.\\cite{dob1980a,dob1980b,gra1982,gra1983} In the context of inhomogeneous MHD turbulence, the cross helicity has been investigated mostly in the solar-wind research. By using spacecraft observations, detailed spectra of cross helicity have been examined.\\cite{bel1971,rob1987,tum1995} In order to explain the large-scale behavior of the solar-wind turbulence, several models have been proposed.\\cite{zho1990,tum1995} However, investigations related to the cross helicity are mostly concentrated on arguments of its production rate, and effects of large-scale inhomogeneities such as the mean velocity shear have been discussed. Matthaeus and coworkers have employed a kind of algebraic model for the cross-helicity dissipation rate.\\cite{zho1990,mat2004,bre2005,bre2008} Adopting this algebraic model of $\\varepsilon_W$, Usmanov {\\it et al}.\\ have recently performed a series of elaborated numerical simulations on the large-scale evolution of solar-wind turbulence.\\cite{usm2011} However, generally speaking, arguments concerning the dissipation rate of $W$ are still far from sufficient. Also in the context of the turbulence dynamo, the transport equation for the turbulent cross helicity has been considered, where an algebraic model for the cross-helicity dissipation rate has been proposed.\\cite{yos1990,yos1993} Sur and Brandenburg wrote down an evolution equation for the cross-helicity effect, and argued the cross-helicity destruction in the context of quenching mechanism.\\cite{sur2009} In order to examine the evolution of the turbulent cross helicity $W$, it is indispensable to properly estimate the dissipation rate of $W$, $\\varepsilon_W$, as well as the cross-helicity production rate $P_W$. We address this problem; modeling the cross-helicity dissipation rate on the basis of a statistical analytical theory. Spacecraft observations of solar-wind turbulence have revealed detailed information on the large-scale behavior of turbulent statistical quantities, which includes the radial evolution of the cross helicity both in the low- and high-speed wind regions. Comparison of the satellite observations with the numerical simulation with the aid of a turbulence model provides a good test for the cross-helicity dissipation models. In this work, we will delve into the problem of cross-helicity dissipation modeling by using such comparisons. The organization of this paper is as follows. After briefly showing the evolution equation of the cross helicity in Section~\\ref{sec:level2}, we present the exact equation of the cross-helicity dissipation rate $\\varepsilon_W$ in Section~\\ref{sec:level3}. Then, we consider two candidates for the model of $\\varepsilon_W$ in Section~\\ref{sec:level4}; one is the algebraic model and another is a transport-equation model. These expressions are systematically derived with the aid of a statistical analytical theory of inhomogeneous turbulence. Some features of the adopted turbulence model, which is an expansion of the hydrodynamic $k - \\epsilon$-type one-point turbulence model in the engineering field, are noted in Section~\\ref{sec:level5}. An application of the model to the solar wind is presented in Section~\\ref{sec:level6}. A brief summary is given in Section~\\ref{sec:level7}. %------------------------------------------------------------- %\t2 Equation for the cross helicity %------------------------------------------------------------- ", "conclusions": "\\label{sec:level7} The evolution of turbulent cross helicity ($W \\equiv \\langle {{\\bf{u}}' \\cdot {\\bf{b}}'} \\rangle$) was investigated from the viewpoint of generation and destruction mechanism of it. In particular, two possibilities of expressing the dissipation rate of $W$, $\\varepsilon_W$, were presented; (i) the algebraic model and (ii) the evolution equation for $\\varepsilon_W$. It was shown that both model expressions can be systematically derived from the zeroth- and first-order calculations of the statistical analytical theory of inhomogeneous turbulence, respectively. Validity of the model expressions was examined with the aid of a turbulence model constituted by four one-point turbulent statistical quantities (the turbulent MHD energy $K$, its dissipation rate $\\varepsilon$, the turbulent cross helicity $W$, and the turbulent residual energy $K_{\\rm{R}}$). It was shown that, as far as the application to the solar-wind turbulence is concerned, the algebraic model for $\\varepsilon_W$ with one model constant gives results plausible enough. Dependence of the cross-helicity evolution on the large-scale velocity structure was also discussed. In the context of the solar wind, it was shown that both of flow expansion and large-scale velocity shear contribute to decreasing the magnitude of the scaled cross helicity, which confirmed earlier results.\\cite{mat2004,bre2005,bre2008} More detailed expressions for the $\\varepsilon_W$ equation were also indicated from the higher-order calculation of the statistical analytical theory of inhomogeneous turbulence. %------------------------------------------------------------- %\t\tAcknowledgments %-------------------------------------------------------------" }, "1005/1005.2538_arXiv.txt": { "abstract": "We determine an expression for the cosmic variance of any ``normal'' galaxy survey based on examination of $M^* \\pm 1$ mag galaxies in the SDSS DR7 data cube. We find that cosmic variance will depend on a number of factors principally: total survey volume, survey aspect ratio, and whether the area surveyed is contiguous or comprised of independent sight-lines. As a rule of thumb cosmic variance falls below 10\\% once a volume of $10^7h_{0.7}^{-3}$Mpc$^3$ is surveyed for a single contiguous region with a 1:1 aspect ratio. Cosmic variance will be lower for higher aspect ratios and/or non-contiguous surveys. Extrapolating outside our test region we infer that cosmic variance in the entire SDSS DR7 main survey region is $\\sim 7$\\% to $z < 0.1$ The equation obtained from the SDSS DR7 region can be generalised to estimate the cosmic variance for any density measurement determined from normal galaxies (e.g., luminosity densities, stellar mass densities and cosmic star-formation rates) within the volume range $10^3$ to $10^7 h^{-3}_{0.7}$Mpc$^3$. We apply our equation to show that 2 sightlines are required to ensure cosmic variance is $<10$\\% in any ASKAP galaxy survey (divided into $\\Delta z \\sim 0.1$ intervals, i.e., $\\sim 1$ Gyr intervals for $z <0.5$). Likewise 10 MeerKAT sightlines will be required to meet the same conditions. GAMA, VVDS, and zCOSMOS all suffer less than 10\\% cosmic variance ($\\sim$ 3\\%-8\\%) in $\\Delta z$ intervals of 0.1, 0.25, and 0.5 respectively. Finally we show that cosmic variance is potentially at the 50-70\\% level, or greater, in the HST Ultra Deep Field depending on assumptions as to the evolution of clustering. 100 or 10 independent sightlines will be required to reduce cosmic variance to a manageable level ($<10$\\%) for HST ACS or HST WFC3 surveys respectively (in $\\Delta z \\sim 1$ intervals). Cosmic variance is therefore a significant factor in the $z>6$ HST studies currently underway. ", "introduction": "The Universe is not homogeneous except on the largest scales ($>1$Gpc, Davis et al.~1985). As a consequence number and density measurements derived from within modest volumes will show greater than Poisson variation (see Szapudi \\& Colombi 1996 for example). This cosmic variance\\footnote{Technically the term sample variance is more correct but here we adhere to the current convention of using the term cosmic variance to describe perturbations in measurements within our Universe due to sampling size.}, or small-scale scale-dependent inhomogeneity, is often the dominant source of error in many contemporary extragalactic measurements. Examples include the galaxy luminosity function/densities (Norberg et al.~2002a; Hill et al.~2010), the HI mass function (Zwaan et al.~2005), the cosmic star-formation history (Hopkins \\& Beacom~2006), and the stellar mass density (Wilkins, Trentham \\& Hopkins~2008). Generally any number or density measurement derived from the galaxy population as a whole is susceptible. Typically, although often neglected in many studies, the cosmic variance can be estimated through one of four methods: Comparison with numerical simulations which encompass larger volumes such as the Millennium Simulation (e.g., Newman \\& Davis~2002; Somerville et al.~2004; Trenti \\& Stiavelli~2008; Moster et al.~2010); analytically using measurements of the 2 or 3-pt correlation functions (e.g., Driver et al.~2003); empirically by Monte-Carlo sampling of a larger survey (e.g., Driver et al.~2005; Hill et al.~2010); or, also empirically, by Jackknife sampling of the volume in question (e.g., Liske et al.~2003). These methods all have strengths and weaknesses can be laborious to impliment and potentially inconsistent depending on the method adopted and the assumptions made. For example to estimate the cosmic variance from numerical simulations requires the adoption of a numerical simulation (e.g., Springer et al.~2005), and either a semi-analytical prescription (Cole et al.~ 2000; Baugh~2006), or a halo occupation distribution (Berlind \\& Weinberg~2002; Moster et al.~2010) before an appropriate cosmic variance estimate can be made (e.g., Moster et al.~2010). The analytical method (e.g., Driver et al.~2003) implicitly assumes Poisson statistics, radial symmetry, and (because it is parametric) smoothes over potential 'features' in the underlying distribution (e.g., BAOs). The empirical method is not always practical if a large suitable survey does not exist, and Jackknife sampling (where one divides the sample into many parts and recomputes the value in question with each part missing in turn) is only capable of revealing the cosmic variance on scales smaller than the volume in question (nevertheless a useful indicator as cosmic variance should generally decrease with increasing scale). In this paper we aim to use the largest volume survey to date, the Sloan Digital Sky Survey (SDSS), to empirically determine some generically useful formulae for estimating the cosmic variance as a function of survey volume and survey shape. These formulae should also assist in the design of future surveys where trade offs between area and depth need to be made. Throughout we adopt a standard cosmology with the following parameter set: $\\Omega_M=0.3, \\Omega_{\\Lambda}=0.7, H_o=70$kms$^{-1}$Mpc$^{-1}$ although as our analysis is based on very local volumes only the value of the adopted Hubble constant is significant. ", "conclusions": "We have derived a simple empirical expression for calculating cosmic variance for almost any extragalactic survey. The results are entirely empirical and based on resampling the SDSS DR7. The resulting equations agree extremely well with the recent numerical results by Moster et al.~(2010). The two resulting equations provide corrections for $z<0.1$ robustly and for $z>0.1$ under the following caveats: ~ \\noindent (1) Te derived cosmic variance is for $M^* \\pm 1$ mag population only and assumed not to evolve with lookback time -- this is clearly incompatible with our understanding of the evolution of structure and hence beyond $z \\sim 1$ the derived values should be taken as indicative only. ~ \\noindent (2) That above $250h^{-1}_{0.7}$Mpc cosmic variance scales with radial co-moving length according to Poisson statistics. ~ The two equations are then used to determine cosmic variance values for a number of recent, ongoing and planned surveys." }, "1005/1005.0684_arXiv.txt": { "abstract": "We analysed the IGR~J16465--4507 Burst Alert Teelescope survey data collected during the first 54 months of the \\sw\\ mission. The source is in a crowded field and it is revealed through an ad hoc imaging analysis at a significance level of $\\sim$14 standard deviations. The 15--50 keV average flux is $\\sim3 \\times 10^{-11}$ \\ferg. The timing analysis reveals an orbital period of 30.243$\\pm$0.035 days. The folded light curve shows the presence of a wide phase interval of minimum intensity, lasting $\\sim 20\\%$ of the orbital period. This could be explained with a full eclipse of the compact object in an extremely eccentric orbit or with the passage of the compact source through a lower density wind at the orbit apastron. The modest dynamical range observed during the BAT monitoring suggests that IGR~J16465$-$4507 is a wind-fed system, continuously accreting from a rather homogeneous wind, and not a member of the Supergiant Fast X-ray Transient class. ", "introduction": "} \\begin{figure}% \\begin{center} \\centerline{\\includegraphics[width=7.5cm,angle=0]{figure1a.ps}} \\centerline{\\includegraphics[width=6cm,angle=270]{figure1b.ps}} \\caption[IGRJ16465-450 sky map]{{\\bf Top}: 15--50 keV significance map in the neighborhood of IGR~J16465$-$4507. The color-bar represents the significance levels. The cyan star marks the XMM position \\citep{ZuritaHeras2004:16465-4507} of IGR J16465$-$4507. The white circles are centered on the position derived by fitting the significance profile extracted along the white line with two Gaussians plus a constant; their radius corresponds to the 90\\% error on the position. {\\bf Bottom}: Significance profile extracted along the white line in the above map. The plot shows the data (stars) and the best fit model (blue line, the sum of two Gaussian profiles plus a constant value). The higher peak (red line) corresponds to IGR J16479$-$4514, the lower peak (green line) corresponds to IGR J16465$-$4507. } \\label{map} \\end{center} \\end{figure} High mass X--ray binaries (HMXBs), stellar systems composed of a compact object and an early-type massive star, are traditionally divided in two subclasses \\citep[e.g.][and references therein]{vanParadijs1995:binaries}, depending on the nature of the high mass primary and, consequently, the different mass-transfer and accretion mechanism. On one side are the systems with main sequence Be primaries (Be-HMXBs). They are generally wide ($P_{\\rm orb}\\ga 10$\\,d) eccentric (eccentricity $e\\sim 0.3$--0.5) systems in which the primaries are not filling their Roche lobe, and accretion onto the compact object occurs from the equatorial region of the rapidly rotating Be star. Most of these systems are highly variable: in some of them recurrent outbursts are observed caused by an enhanced rate when the compact star passes close to the Be star. On the other side are the systems with an evolved OB supergiant primary (sgHMXB). Their periods are shorter ($P_{\\rm orb}\\la 10$\\,d) and their orbits more circular than in Be-HMXBs. They are powered either by a geometrically thin accretion disc or by the strong radiation-driven stellar wind, depending on whether the primary fills its Roche lobe or not. Their X-ray emission is bright and persistent. Recently, this rather clear-cut picture was made more structured with the INTEGRAL observations of the Galactic plane. Two additional classes were added to the classical OB primary HMXBs: the highly absorbed persistent systems (Walter et al. 2004, 2006) and the supergiant fast X--ray transients \\citep[SFXTs, ][]{Sguera2005,negueruela06esa, smith04,intzand05}. The former are characterized by orbital and spin periods consistent with those observed in wind-accreting systems, but a much higher absorbing column density. The latter are transient sources showing a large dynamic range of 3--5 orders of magnitude with sporadic outbursts (which however are significantly shorter than those of typical Be-HMXBs), characterized by bright flares lasting up to days with peak luminosities of 10$^{36}$--10$^{37}$~erg~s$^{-1}$ \\citep{Sguera2005,Romano2009:sfxts_paperV,sidoli09}. The Burst Alert Telescope \\citep[BAT,][]{Barthelmy2005:BAT} on board \\sw\\ \\citep{Gehrels2004mn} is performing a continuous coverage of the hard X-ray sky (50 to 80\\% of the sky every day). This allowed the detection of many of the new INTEGRAL HMXBs (e.g. \\citealp{cusumano10}) and the collection of their long term light curves. In this Letter we analyse the hard X-ray data collected during the first 54 months of \\sw-BAT sky monitoring using data in the region of the IGR~J16465--4507. This source was discovered by INTEGRAL in 2004 \\citep{Lutovinov2004:16465-4507} and X-ray activity was observed with IBIS/ISGRI starting on September 6, at a flux level of 8.8 $\\pm$ 0.9 mCrab (18-60 keV), followed by a flare (up to 28 mCrab) on September 7. Follow-up observations with XMM-Newton revealed pulsations at 228$\\pm$6 s \\citep{Lutovinov2005} and allowed the identification of the optical counterpart with 2MASS~J16463526--4507045 \\citep{ZuritaHeras2004:16465-4507}. This was classified as a B0.5 Ib supergiant at a distance of $\\sim 8$\\,kpc \\citep{Negueruela2007} or as a O9.5 Ia supergiant at a distance of $9.5_{-5.7}^{+14.1}$\\,kpc \\citep{Nespoli2008}. The supergiant nature of the companion, combined with the observed hard X-ray variability \\citep{Lutovinov2004:16465-4507} and the X-ray spectral distribution, modeled by a hard power law with photon index 1.0$\\pm$0.52 \\citep{Lutovinov2005}, suggested a classification of this source as a SFXT \\citep{Negueruela2006}. However, \\citet{Walter2007}, based on INTEGRAL measurements, suggested that IGR~J16465--4507 is likely a classical supergiant HMXB, with an average flux just below the IBIS/ISGRI sensitivity undergoing sporadic long periods of enhanced activity. This Letter is organized as follows. Section 2 describes the BAT data reduction and the imaging analysis. Section 3 reports on the timing analysis. Sect. 4 describes the analysis of the pointed soft X-ray observation with \\sw-XRT. In Sect.\\ 5 we briefly discuss our results. Errors are at 90\\,\\% confidence level, if not stated otherwise. ", "conclusions": "} We have analysed the data collected by \\sw-BAT during the first 54 months of the \\sw\\ mission in the region of the supergiant HMXB \\src16465 {}. The source is detected at a significance level of $\\sim$ 14 standard deviations with an average flux of $\\sim 2.36 \\times 10^{-11}$ \\ferg\\ in the 15--50 keV energy band. The light curve reveals a periodicity of $30.243 \\pm 0.035$ days that we interpret as the orbital period of the binary system. By applying Kepler's third law, the semi-major axis of the binary system is given by $a^3=P_{\\rm orb}^2\\times \\rm G(M_{\\star}+M_{\\rm X})/4\\pi^2$, where M$_{\\star}$ and M$_{\\rm X}$ are the masses of the supergiant and compact object, respectively. We adopt M$_{\\rm X}=1.4$~M$_{\\sun}$ and M$_{\\star}=27.8$~M$_{\\sun}$ \\citep[][for an O9.5 I star of radius R$_{\\star}=22.1$~R$_{\\sun}$]{Martins2005}. This yields $a\\sim125$~R$_{\\sun} \\sim 6$~R$_{\\star}$. We note that the assumption of a stellar type B0.5 Ib would lead to an estimate of $a\\sim150$~R$_{\\sun} \\sim 5$~R$_{\\star}$ \\citep[][M$_{\\star}=47$~M$_{\\sun}$, R$_{\\star}=32.2$~R$_{\\sun}$]{Searle2008}. The folded profile shows the presence of a dip with a count rate consistent with no emission, that could be intepreted as a full eclipse. However, the width of this dip ($\\sim 20\\%$ of P$_{\\rm orb}$) is not consistent with the duration of the eclipse expected for a circular orbit with 6 R$_{\\star}$ semimajor axis, assuming an edge-on inclination. To obtain such a long eclipse, the system should have an eccentricity of at least 0.8 coupled with a high inclination. This scenario is unlikely. Alternatively, this wide phase interval of minimum intensity could be explained with an eccentric orbit and a lower wind density at apastron. Given our knowledge of both a spin period and an orbital period we can locate the source on the Corbet diagram \\citep{Corbet1986:diagram} together with the known $P_{\\rm spin}$ and $P_{\\rm orb}$ of other binary systems \\citep{Bildsten1997,Liu2006:hmxb}. We also plot in the diagram the position of the two SFXTs IGR J18483-0311 \\citep{Levine2006:igr18483,Zurita2009:sax1818.6_period} and IGR~J11215-5952 \\citep{Swank2007:atel999,Sidoli2007,Romano2009:11215_2008}. \\src16465 {} sits at the boundary of the wind-fed OB-HMXBs (systems with OB supergiants that underfill their Roche-lobes) and the locus of the Be transients (Fig.~\\ref{corbet}. The characteristics of the system \\citep{Lutovinov2004:16465-4507,Lutovinov2005} suggested a classification of this source as a SFXT \\citep{Negueruela2006}. However, the outburst history of this source is rather scarce. After the initial discovery by \\inte\\ no further outbursts have been reported. \\citet{Walter2007} mention three episodes of enhanced hard X-ray activity observed by \\inte\\ at the limit of the instrumental sensitivity, that cannot be considered as real flares. The BAT 15--50\\,keV light curve during the 54 months of monitoring indicates that this is a faint persistent source, with an average luminosity of $2.6\\times10^{35}$ erg s$^{-1}$ (at a distance of 9.5 kpc), no evidence for flaring activity and a weak variability with a dynamical range lower than 10. Based on their X-ray variability, sgHMXB can be classified as classical and absorbed systems ( variability factor $<$20) or as SFXT (variability factor $>100$). The timing behavior of IGR~J16465$-$4507 suggests a wind-fed system, with a neutron star continuously accreting from a rather homogeneous wind. The luminosity, lower than what observed in classical sgHMXB ($10^{36}-10^{37}$ erg s$^{-1}$), can be explained with the larger orbital separation ($\\sim$5 R$_{\\star}$ compared to $\\sim$2 R$_{\\star}$ in classical systems). \\begin{figure}% \\begin{center} \\centerline{\\includegraphics[width=6cm,angle=90]{figure5.ps}} \\caption[]{The Corbet diagram showing the neutron star $P_{\\rm spin}$ vs.\\ binary period $P_{\\rm orb}$. Black circles are HMXBs with an OB primary, empty circles those with a Be one. Larger (red) symbols represent SFXTs: triangles are IGR~J18483$-$0311 and IGR~J11215$-$5952, the star the newly determined position of \\src16465 . } \\label{corbet} \\end{center} \\end{figure}" }, "1005/1005.2681_arXiv.txt": { "abstract": "{Neutrinos emitted from a supernova encode useful information about neutrino physics and astrophysics. Interpreting the neutrino signal depends crucially on understanding neutrino production, flavor mixing during propagation, and detection. In this talk, we review the physics potential of a SN neutrino observation.} \\FullConference{35th International Conference of High Energy Physics - ICHEP2010,\\\\ July 22-28, 2010\\\\ Paris France} \\begin{document} ", "introduction": "Neutrinos from a core-collapse supernova (SN)~\\footnote{Neutrinos refer to both neutrinos and antineutrinos. Only neutrinos will be written as $\\nu$. Similarly, antineutrinos will be written as $\\bar{\\nu}$. Also, SN refers to core-collapse supernovae only.} provide a rare and valuable physics opportunity~\\cite{Dighe:2008dq}. The gargantuan fluxes and low interaction rates of SN neutrinos, makes them ideal candidates to probe neutrino mixing properties and study the extreme conditions in the depths of a star. Observation of the SN $1987$A in neutrinos opened the field of neutrino astronomy~\\cite{Koshiba:1992yb}, and confirmed our overall understanding of SN neutrino physics~\\cite{Raffelt:1996wa}. With present and planned detectors~\\cite{Scholberg:2010zz}, a galactic SN is expected to result in a high-statistics detection. This would lead to significant advances in neutrino physics and SN astrophysics, if we can disentangle the relevant information. Decoding the neutrino signal requires a detailed understanding of neutrino production, mixing during propagation, and detection. For neutrinos, the mixing scenario is reasonably well-determined, except the value of $\\theta_{13}$, the sign of $\\Delta m^2_{\\rm atm}$, and the CP-violating phase~\\cite{GonzalezGarcia:2010er}, and the main detection channels are well-calibrated. However for SNe, initial neutrino fluxes and spectra predicted from SN theory have a significant variance~\\cite{Simulations}. Similarly, the stellar conditions are poorly constrained. The signal interpretation is plagued by these uncertainties, and the strategy for disentangling information from a SN neutrino signal must rely on generic features that are insensitive to model assumptions. Some information is obtained directly, e.g. the direction and time structure of the event, and the above mentioned uncertainties do not affect our inferences. But a lot more information, e.g. flavor dependent energy spectra of the neutrinos, clues to the unknown neutrino parameters, and some signatures of stellar dynamics, is encoded in a flavor dependent way. A detailed treatment of neutrino mixing is required to extract that information. Traditional analyses of neutrino mixing for SN took into account interactions with matter, through the Mikheev-Smirnov-Wolfenstein (MSW) effect. This MSW-based paradigm was believed to be complete, and numerous results followed. See the review~\\cite{Dighe:2004xy} for the traditional expectations. However, that picture was incomplete. Deep in the SN, neutrino densities are large enough to make their collective interactions extremely important. The mixing angles are highly matter suppressed, and one may expect no flavor conversion in that region. However, this naive expectation is incorrect. Neutrino-neutrino interactions entangle the flavor evolution of all neutrinos and create an instability in the flavor composition. Thus flavor conversions take place even for extremely small mixing angles, with a rich phenomenology. See the recent review~\\cite{Duan:2010bg} on these so-called ``Collective effects''. In this talk, we summarize the main aspects of SN neutrinos, focussing on the collective effects and their impact on the physics potential of SN neutrino observations. ", "conclusions": "The main hurdle in interpreting SN neutrino data, besides the absence of it at present, is the lack of knowledge of initial conditions, i.e. the initial fluxes and densities. This is likely to lead to degeneracies. However, one may expect various aspects of the time and flavor dependent SN signal to be used in synergy. Future developments in SN theory/simulations could be expected to reduce or eliminate some of these degeneracies. Another area of improvement would be a better understanding of the flavor conversion. In particular, the effect of anisotropy and inhomogeneity on collective effects and MSW conversions. A future galactic SN can be expected to provide a wealth of scientific information for neutrino oscillation physics and SN astrophysics. This is a rare opportunity, occurring once-in-a-lifetime, and we must be ready with suitable detectors and the required theoretical understanding to interpret the data. A significant step towards this, would be a detailed understanding of the rich phenomenology of neutrinos from supernovae." }, "1005/1005.5068_arXiv.txt": { "abstract": "Radio pulsar surveys are producing many more pulsar candidates than can be inspected by human experts in a practical length of time. Here we present a technique to automatically identify credible pulsar candidates from pulsar surveys using an artificial neural network. The technique has been applied to candidates from a recent re-analysis of the Parkes multi-beam pulsar survey resulting in the discovery of a previously unidentified pulsar. ", "introduction": "Since the discovery of pulsars by Jocelyn Bell and Antony Hewish at Cambridge in 1967 using a pen chart recorder (Hewish et al. 1968\\nocite{hbp+68}), pulsar searching has come a long way. Modern pulsar surveys use high performance computing facilities to perform an extensive range of signal processing and search algorithms. These methods are designed to maximize sensitivity to weak, rapid, and dispersed pulsar signals often buried in large amounts of terrestrial radio frequency interference (RFI), or even in binary systems. There is little doubt that these complex algorithms have aided searches for pulsars, however there remain certain search tasks for which standard computer programs are of little use. In particular, the final stage of a pulsar search, the selection of credible pulsar candidates for follow-up observations, which still remains a task for a human since the decision is visual and based on a number of combined properties of the pulsar signal. The process can be time consuming and inefficient in analysis of large-scale surveys that produce many millions of pulsar candidates. Large-scale pulsar surveys such as the Parkes multi-beam pulsar survey (PMPS) (Manchester et al. 2001)\\nocite{mlc+01} have dramatically increased the number of known pulsars. Finding more pulsars elucidates the properties of their Galactic population, and also offers the possibility of uncovering new and extreme phenomena in neutron-star astrophysics. Future pulsar surveys will be done with the next generation of radio telescopes, such as LOw Frequency ARray (LOFAR), the Five hundred metre Aperture Spherical Telescope (FAST), and the Square Kilometre Array (SKA) (e.g. van Leeuwen \\& Stappers 2010\\nocite{2010A&A...509A...7V}, Smits et al., 2009a,b\\nocite{2009A&A...505..919S}\\nocite{2009A&A...493.1161S}, Cordes et al., 2004\\nocite{2004NewAR..48.1413C}). These radio telescopes will be excellent survey tools because of their large collecting areas, capability to form many simultaneous beams on the sky, and in the case of the interferometers, wide fields of view. It is expected that these instruments will detect a large fraction of the observable pulsars in the Galaxy. The inevitable flood of pulsar candidates that will require inspection to achieve this will certainly require some form of multi-person or machine based candidate selection. Some large-scale astronomical surveys and data mining projects have resorted to employing many online volunteers to search for or classify their objects of interest in so-called `citizen science' projects (e.g. Lintott et al., 2008\\nocite{2008MNRAS.389.1179L}, Westphal et al., 2005\\nocite{2005LPI....36.1908W}). In searches for pulsars, the pulsar Arecibo L-band Feed Array (ALFA) survey collaboration has enlisted the help of High School and undergraduate students in a successful outreach and science program to identify potential pulsar candidates (Jenet et al., 2007\\nocite{2007AAS...211.0517J}). Recent machine solutions include, candidate ranking based on likelihoods calculated from parameter distributions of pulsar and non-pulsar signals (Lee, private communication), and the sorting of candidates based on a number of `scores' that indicate the similarity of the signal to that of a typical pulsar (Keith et al., 2009)\\nocite{kel+09}. In this paper we present an alternative method whereby an artificial neural network (ANN) has been trained using a particular set of scores to automatically identify credible pulsar candidates from a recent re-analysis of the PMPS. ANNs have long been used in other areas of astronomy, for example in the morphological classification of galaxies (e.g. Storrie-Lombardi et al., 1992\\nocite{1992MNRAS.259P...8S}, Zhang, Li \\& Zhao 2009\\nocite{2009MNRAS.392..233Z}), the estimation of photometric redshifts of sources in the Sloan Digital Sky Survey (Firth, Lahav \\& Somerville 2003\\nocite{2003MNRAS.339.1195F}) and in the selection of microlensing events from large variability surveys \\nocite{2003MNRAS.341.1373B} (Belokurov, Evans \\& Du 2003). The outline of this paper is as follows: In Section 2 we describe how pulsar candidate selection in our recent re-analysis of the PMPS is typically done and some of the problems associated with this method. Section 3 gives a brief introduction to ANNs. In Section 4 we describe our implementation of an ANN to classify pulsar candidates from our re-analysis of the PMPS and the results from a test on a sample of the search output data. Preliminary parameters of the pulsar discovered using an ANN during our re-analysis of the PMPS are given in Section 5. Finally Section 6 gives a summary, followed by a discussion of the future application of ANNs to pulsar surveys. ", "conclusions": "ANNs have been developed to aid searches for radio pulsars in a database of $\\sim$ 16 million pulsar candidates generated from a recent re-analysis of the PMPS. The ANNs were trained using small sets of characteristic scores derived from the pulsar candidate plots generated for human inspection. Around $92$ per cent of the pulsars present in a test sample of $\\sim$ 2.5 million pulsar candidates can be recovered by our ANNs. It is likely that pulsars from the test sample were missed due to one of three reasons: poor training of the ANNs on MSPs, abnormal candidate plots generated by our search software, or unbalanced training sets. Future implementations should avoid unbalanced training sets or bias the training sets toward the objects of interest. A larger number of false positives from the ANN is beneficial for searches of the inherently rare objects in the data set viz. pulsars \\nocite{2003MNRAS.341.1373B} (e.g. Belokurov, Evans \\& Du 2003). To test if the `contamination' from non-pulsars in the results set could be further reduced the threshold for pulsar identification, $z_{1}$, was increased to $0.7$, $0.8$, and $0.9$. In each case the number of genuine pulsar to non-pulsar candidates recovered by the ANN were 1:24, 1:23, and 1:20 respectively, however the corresponding recovered fraction of pulsars were 89, 86, and 81 per cent. In a search for pulsars these few per cent could be potential new discoveries. As such higher thresholds should only be used when the number of candidates produced by the ANN are unmanageable. Improving the detection efficiency of our ANNs will be the subject of future work. Improvements in this area are likely to depend upon better representation of the two-dimensional diagnostic plots in our input vectors. These planes are particularly useful for discriminating against narrow-band and impulsive RFI. ANNs are only capable of classifying the inputs given to them based on the training received. By training the ANN using a set of pulsars with particular characteristics we reduce the possibility of uncovering atypical, unusual, or unexpected phenomena. Therefore, it is important to use training sets of pulsars with a wide variety of properties and displaying phenomena such as scintillation, scattering, intermittency and binary motion. Without training on pulsars that exhibit such properties an ANN will only select against these objects. As described in Section~\\ref{s:tests} MSPs display somewhat different candidate plots to their standard pulsar counterparts. Separate ANN training using only data from MSPs may be required to find these pulsars. For more exotic pulsars, such as relativistic binary pulsars, any training sample will be limited in size. To find these systems it might be possible to form training sets from simulated pulsar signals covering a wide range of pulsar and binary parameters. However, it is expected that ANNs trained with simulated training sets will not perform as well as those based on real data due to the subtle effects of instrumentation and RFI environments. In new pulsar surveys where the search data may vary from that of previous surveys, for example in the number of frequency channels and the sampling time, dedicated observations of known pulsars will be required to build a training database of MSPs and the standard population of pulsars. A complementary class of ANNs are unsupervised ANNs which require no desired or `target' vector during training. The best example of such a ANN is the Kohonen Self Organizing Map (SOM) (e.g. Kohonen, 2001)\\nocite{Kohonen01}. SOMs provide a way of representing multi-dimensional data in a lower number of dimensions, generally the two-dimensions of a map. Similar input vectors share similar regions on these maps. Such a ANN could be used to divide the pulsar candidates into sub-classes such as MSPs, standard pulsars, noise signals, and even common RFI signals. Although not optimal for candidate selection such ANNs will be investigated in future work. Future pulsar surveys will probe larger volumes of both real space and the parameter space associated with radio pulsars. The SKA will have a sensitivity that will allow the discovery of all $20\\;000$ to $30\\;000$ pulsars beaming towards Earth and visible from the short listed sites in South Africa and Western Australia. Inspecting the candidates from such surveys will be a tremendous data analysis task. For illustration, scaling the PMPS up to an all sky survey would result in approximately a factor of 40 increase in the number of pulsar candidates. Using the same search algorithm as in our recent re-analysis of the PMPS and assuming an inspection time of one second per pulsar candidate, viewing all the candidates from such as survey would take over 20 years. This figure does not take into account beams that contribute many erroneous candidates due to RFI and which are relatively easy to flag and avoid using graphical selection tools. However, as we have already discussed these cuts can increase the risk of missing genuine pulsars that lie near to RFI in phase-space. Using an ANN and assuming the same recovered fraction of candidates of $0.5$ per cent the total inspection time would come down to just over one month, vastly reducing the work load on human observers. We would like to stress that simple ANNs, like those presented in this work, should not yet be treated as a complete replacement for human inspection of pulsar candidates. Humans are still best placed to spot anything unusual or interesting about the individual pulsar candidates being viewed. Because of the significant time savings, of around two orders of magnitude, ANNs might be best used for first-passes over search output data before a more detailed manual inspection with graphical selection tools." }, "1005/1005.1966_arXiv.txt": { "abstract": "We investigate the radio and $\\gamma$-ray beaming properties of normal and millisecond pulsars by selecting two samples from the known populations. The first, Sample G, contains pulsars which are detectable in blind searches of $\\gamma$-ray data from the \\emph{Fermi} Large Area Telescope. The second, Sample R, contains pulsars detectable in blind radio searches which have spin-down luminosities $\\dot{E} > 10^{34}$~erg s$^{-1}$. We analyse the fraction of the $\\gamma$-ray-selected Sample G which have detectable radio pulses and the fraction of the radio-selected Sample R which have detectable $\\gamma$-ray pulses. Twenty of our 35 Sample G pulsars have already observed radio pulses. This rules out low-altitude polar-cap beaming models if, as is currently believed, $\\gamma$-ray beams are generated in the outer magnetosphere and are very wide. We further find that, for the highest-$\\dot{E}$ pulsars, the radio and $\\gamma$-ray beams have comparable beaming factors, i.e., the beams cover similar regions of the sky as the star rotates. For lower-$\\dot{E}$ $\\gamma$-ray emitting pulsars, the radio beams have about half of the $\\gamma$-ray sky coverage. These results suggest that, for high-$\\dot{E}$ young and millisecond pulsars, the radio emission originates in wide beams from regions high in the pulsar magnetosphere, probably close to the null-charge surface and to the $\\gamma$-ray emitting regions. Furthermore, it suggests that for these high-$\\dot E$ pulsars, as in the $\\gamma$-ray case, features in the radio profile represent caustics in the emission beam pattern. ", "introduction": "The observed pulses from pulsars are interpreted as an emission beam, fixed relative to the rotating neutron star, sweeping across the Earth. The magnetospheric emission-region structure, the viewing angle with respect to the rotation axis and the angle between the magnetic and rotation axes define the shape of the observed pulse profile. Pulsar radio emission is generally thought to originate within a few percent of the light-cylinder radius ($R_{\\rm LC}$) above the polar caps \\citep[e.g.][]{lm88}. Many pulsars exhibit a ``radius-to-frequency'' mapping \\citep{cor78} which is indicative of polar cap emission. The increasing pulse width at lower radio frequencies is interpreted as emission at increasing heights above the poles along more widely directed field lines. The radio profiles of many young pulsars and millisecond pulsars (MSPs) have an ``interpulse'', a second pulse component separated from the main pulse by close to $180\\degr$ of pulse phase, or widely separated pulse components. In most cases, such pulsars do not exhibit radius-to-frequency mapping, that is, they have frequency-independent pulse component separations \\citep{hf86}. The known population of $\\gamma$-ray pulsars has recently been greatly expanded by the six-month data release of the \\emph{Fermi} Large Area Telescope (LAT) \\citep[][hereafter A10]{aaa+10c}. This sample of 46 $\\gamma$-ray pulsars includes both young pulsars and MSPs, all with high spin-down luminosities: \\begin{equation} \\dot{E}=\\frac{4\\pi^{2}I\\dot{P}}{P^{3}}>10^{33}\\,{\\rm erg\\,s^{-1}}, \\end{equation} where $I=10^{45}$\\,gm\\,cm$^{-2}$ is the assumed neutron star moment of inertia, $P$ is the rotation period in seconds and $\\dot{P}$ is the period time-derivative. Recent studies indicate that $\\gamma$-ray emission from pulsars originates at high altitudes in the pulsar magnetosphere, near $R_{\\rm LC}$ and from the last open field lines \\citep[e.g.,][A10]{aaa+09,aaa+09d}. Two classes of high-altitude particle acceleration and emission model are generally considered: slot-gap accelerators \\citep{mh04a,dr03} from the polar cap outwards along the last open field lines, and outer-gap accelerators \\citep{ry95,cz98} where the last closed field lines cross the null-charge surface defined by ${\\bf \\Omega\\cdot B} = 0$ ($\\bf \\Omega$ is the rotation velocity vector, and $\\bf B$ is the magnetic field vector). These models result in what we shall term \\emph{wide-beam} emission geometries. Low-altitude models for $\\gamma$-ray emission have been largely ruled out by the generally wide and multi-component $\\gamma$-ray pulse profiles and the existence of pulsed $\\gamma$-ray emission at energies greater than a few GeV; at low altitudes such emission is precluded by electron-positron pair absorption. Furthermore, \\citet{rw09} find that outer-gap models are statistically preferred over slot-gap models based on fits to LAT pulse profiles for a sample of bright $\\gamma$-ray pulsars. Four of the five brightest $\\gamma$-ray pulsars detected by the Energetic Gamma-Ray Experiment Telescope (EGRET) are also radio pulsars \\citep{tho08}. This, together with the similar radio and $\\gamma$-ray pulse morphology for many young pulsars and MSPs, led \\citet{man05} to suggest that, for these high-$\\dot{E}$ pulsars, the radio emission is also emitted in wide beams from the outer magnetosphere near $R_{\\rm LC}$. Within the framework of polar cap emission, \\citet{jw06} and \\citet[][hereafter KJ07]{kj07} proposed that the radio emission from young pulsars is emitted at intermediate altitudes of $\\sim1000$\\,km or $\\sim 0.2 R_{\\rm LC}$ in order to account for wide profiles of young pulsars. In this letter, we identify samples of $\\gamma$-ray-selected and radio-selected pulsars and consider the pulsars in each sample that also emit radio and $\\gamma$-ray pulses respectively. We then discuss the implications of our analysis for the radio beams of young pulsars and MSPs. ", "conclusions": "By comparing the relative occurrence of radio and $\\gamma$-ray pulsed emission in $\\gamma$-ray-selected and radio-selected samples, we find that the radio beaming fraction $f_r$ appears to be close to unity for the highest-$\\dot{E}$ pulsars, decreasing to $\\sim0.5$ for lower-$\\dot{E}$ $\\gamma$-ray pulsars. Our estimated lower bound $f_r$, averaged over all $\\gamma$-ray-selected pulsars, of $\\sim 0.57$ is inconsistent with low-altitude polar-cap models for the radio emission but is consistent with KJ07's intermediate-altitude polar-cap model. However, the KJ07 model is marginally inconsistent with our result that $f_r \\sim 1$ for the highest-$\\dot{E}$ pulsars. With other evidence from radio and $\\gamma$-ray pulse morphologies, these results suggest that, for high-$\\dot{E}$ pulsars, the radio emission originates in wide beams from the vicinity of the null-charge surface, possibly with a slot-gap configuration, with profile components representing caustics in the emission pattern. Improved statistics from further $\\gamma$-ray and radio discoveries will help to test our conclusions. Modelling of radio pulse profiles with slot-gap or other wide-beam emission geometries for both young high-$\\dot{E}$ pulsars and MSPs would be valuable. If outer-magnetoshpere wide-beam radio emission from these pulsars is confirmed, it would have profound implications for pulsar population and evolution studies as well as our understanding of radio and $\\gamma$-ray pulse emission mechanisms." }, "1005/1005.3478_arXiv.txt": { "abstract": "We utilize detailed time-varying models of the coupled evolution of stars and the HI, $\\rm H_2$, and CO-bright H$_2$ gas phases in galaxy-sized numerical simulations to explore the evolution of gas-rich and/or metal-poor systems, expected to be numerous in the Early Universe. The inclusion of the CO-bright H$_2$ gas phase, and the realistic rendering of star formation as an H$_2$-regulated process (and the new feedback processes that this entails) allows the most realistic tracking of strongly evolving galaxies, and much better comparison with observations. We find that while galaxies eventually settle into states conforming to Schmidt-Kennicutt (S-K) relations, significant and systematic deviations of their star formation rates (SFRs) from the latter occur, especially pronounced and prolonged for metal-poor systems. The largest such deviations occur for gas-rich galaxies during early evolutionary stages but also during brief periods at later stages. Given that gas-rich and/or metal-poor states of present-epoch galaxies are expected in the Early Universe while a much larger number of mergers frequently resets non-isolated systems to gas-rich states, even brief periods of sustained deviations of their SFRs from those expected from S-K relations may come to characterize significant periods of their stellar mass built-up. This indicates potentially serious limitations of (S-K)-type relations as reliable sub-grid elements of star formation physics in simulations of structure formation in the Early Universe. We anticipate that galaxies with marked deviations from the S-K relations will be found at high redshifts as unbiased inventories of total gas mass become possible with ALMA and the EVLA. ", "introduction": "} Since it was first proposed as a phenomenological relation linking HI gas mass and star formation surface density in galaxies (Schmidt 1959), and subsequently better constrained and re-formulated as to include also CO-bright H$_2$ gas (Kennicutt 1998, 2008), the Schmidt-Kennicutt (hereafter S-K) relation: $\\rm \\Sigma_{SFR}\\propto [\\Sigma (HI)]^k$ (k$\\sim $1-2), has provided the standard observational framework relating the star formation rate (as a surface density rate $\\rm \\Sigma_{SFR}$) to the gas supply in galaxies. It is also an important element of the sub-grid star formation physics incorporated in galaxy evolution and structure formation models (e.g. Baugh et al. 2005 and references therein; Springel, Di Matteo, \\& Hernquist 2005, Schaye \\& Della Vecchia 2008) where numerical resolution limitations preclude a more detailed treatment of star formation over the scales involved. Many theoretical (Dopita \\& Ryder 1994; Robertson \\& Kravtsov 2008), and observational (e.g. Wong \\& Blitz 2002; Bigiel et al. 2008) studies have been made to demonstrate its validity, with the most important recent advances being the identification of the CO-bright H$_2$ gas as better correlated to star formation in galaxies than atomic hydrogen (Wong \\& Blitz 2002), and the direct star-formation role of its dense phase ($\\rm n(H_2)>10^4\\,cm^{-3}$) with a k$\\sim $1 (Gao \\& Solomon~2004). Unfortunately past analytical and numerical investigations of the S-K relation did not include a multi-phase ISM (though see Gerritsen 1997 for an early investigation that includes it) or assumed sub-grid models reacting instantaneously to changes in the global state of the ISM. Thus they are ill-suited to explore very gas-rich systems and/or early galaxy evolutionary stages, when the various ISM phases and their interplay with the stellar content have not established equilibrium. Finally, such models cannot be compared directly to observations since they do not include the H$_2$ gas phase (the direct fuel of star formation), or do so but assume that all such gas is CO-bright (e.g. Gnedin et al. 2009). The latter is not the case, especially in a metal-poor and/or far-UV intense ISM environment (e.g. Israel 1997; Maloney \\& Black 1988; Pak et al. 1998), which is common during the gas-rich and vigorously star-forming epochs in early galaxy~evolution. \\subsection{ISM+stars galaxy models: features and limitations } Here, we use our time-varying models of the coupled evolution of HI, H$_2$ gas phases and stars in galaxy-sized numerical simulations to investigate the emergence and possible deviations from the S-K relations. Full details and tests of our method can be found in Pelupessy, Papadopoulos, \\& van der Werf (2006) and Pelupessy \\& Papadopoulos (2009). We use an N-body/SPH code and solve for the full thermodynamic evolution of the WNM and CNM HI phases (see Wolfire et al. 2003) assuming neither an equillibrium nor an Effective Equation of State (EOS), unlike most current cosmological or galaxy-sized structure formation models (e.g. Springel \\& Hernquist 2003; Cox et al. 2006a,b; Narayanan et al. 2009). It is worth pointing out that in such models the coldest ISM phase tracked is usually at $\\rm T_{kin}$$\\sim $10$^4$\\,K (i.e. thermodynamically far removed from the one truly forming the stars), and a Schmidt-Kennicutt (S-K)-type relation between that phase and star formation rate is postulated (e.g. Kravtsov et al. 2004; Governato et al. 2007). In recent models colder gas ($\\sim $300\\,K) is tracked, but instantaneous equillibrium gas thermodynamics remains an assumption (e.g Tasker \\& Bryan 2008) while the molecular gas phase and the feedback effects of star formation on it are not included (e.g. Tasker \\& Tan 2009). In our approach star formation is controlled by gravitational instability via a Jeans mass criterion, the thermal state of the gas is tracked explicitely, while an H$_2$-richness criterion for star formation can be applied in addition to that of gravitational instability (see Pelupessy \\& Papadopoulos 2009 for details). The code tracks the H$_2$ phase with a physical model for substructure using a minimal set of assumptions. It follows the H$_2$ formation on dust grains and its thermal \\& far-UV induced destruction, accounting for self-shielding and dust shielding. The CO-bright H$_2$ phase is indentified as a post-processing step, using the most important chemical reactions (R{\\\"o}llig at al. 2006) solving for the C$^{+}$ envelope per gas cloud, as regulated by the local far-UV field. The aspects making our code particularly suitable for examining the validity of the S-K relation over a range of conditions are: a) a time-varying treatment of the H$_2$$\\leftrightarrow$HI gas mass exchange while tracking the ISM from Warm Neutral Medium (WNM) HI ($\\rm T_k$$\\sim $10$^4$\\,K, $\\rm n\\sim $(0.1--1)\\,cm$^{-3}$) to Cold Neutral Medium (CNM) HI and H$_2$ gas ($\\rm T_{k}$$\\sim $(30--200)\\,K, $\\rm n$$\\sim$(10--200)\\,$\\rm cm^{-3}$), b) the versatility of using an H$_2$-regulating star formation criterion, {\\it in addition} to the regular gravitational instability criterion, and c) CO formation and destruction (for this specific version of our code). The latter allows direct comparisons with observations since CO line emission rather than H$_2$ is the real observable in~galaxies. The most realistic models employ our H$_2$-regulated star formation which uses the local $\\rm H_2$ gas mass fraction as a star formation regulator in the dynamical setting of an evolving galaxy. We implement this molecular regulated (MR) star formation by converting only the molecular ($\\rm f_m$) mass fraction of an unstable gas particle to stars. Unlike our other simulations (where H$_2$ is tracked but plays no role in star formation) that need an adhoc parameter $\\epsilon_{SF}$ designating the local gas mass fraction converted into stars, the MR models contain a physical basis for this parameter while retaining the original requirement of SF gas as Jeans unstable (see Pelupessy et al 2006 for more details and tests). Finally here we limit ourselves to systems with $\\rm M_{baryons} \\le 10^{10}\\,M_{\\odot}$ where we can maintain the ability to track the full range of gas densities and temperatures necessary for the HI$\\rightarrow $H$_2$ phase transition to take place (typically in the densest and coldest regions of CNM HI~gas). ", "conclusions": "Our results are of special importance for the modeling of very gas-rich and/or metal-poor progenitors of present-epoch galaxies found in the distant Universe, or systems where major gas mass accretion events frequently ``reset'' their evolutionary states back to gas-rich ones. In such cases the non-equilibrium, non-linear, mass/energy exchange between the various ISM phases and the stellar component may come to dominate significant periods of intense star formation and stellar mass built-up during which the S-K relation is not applicable. In short we find that it may work well for present-epoch metal-rich spirals with modest remaining gas mass fractions (i.e. systems for which the (S-K) relations were originally deduced), but not for their very gas-rich/metal-poor progenitors in the Early Universe. The importance of such deviations of the actual star formation rates from those expected from S-K relations does not lie so much in their magnitude (though this must be explored further for gas-rich systems larger than the ones modelled here) rather than their systematic nature (e.g. star-forming systems spending certain periods with always higher or always lower SFRs than those estimated from the S-K relations). It is the latter that may make such phenomenological relations poor choices for the sub-grid physics of star-formation in rigorous structure formation models in a cosmological~setting. Finally we note that when it comes to the large gas-rich galaxies that are currently accessible at high redshifts our results, drawn for less massive systems, remain provisional. Nevertheless for more massive and very gas-rich star-forming systems the larger amplitudes of ISM equilibrium-perturbing agents (e.g. SNs, far-UV radiation fields), and the shorter timescales characterizing their variations, will more likely than not exaggerate the deviations of true star formation versus the one expected from (S-K)-type phenomenological relations. An unbiased observational effort to find and study (S-K)-deviant galaxies at high redshifts (soon to be possible with ALMA and the EVLA over a wide range of galaxy masses), as well as extending detailed numerical modeling of gas and stars towards larger systems (as computational capabilities improve), are important in establishing the validity range of S-K relations and thus their utility as an important sub-grid element in models of structure formation in the Universe." }, "1005/1005.5095.txt": { "abstract": "The study of multiple extrasolar planetary systems has the opportunity to obtain constraints for the planetary masses and orbital inclinations via the detection of mutual perturbations. The analysis of precise radial velocity measurements might reveal these planet-planet interactions and yields a more accurate view of such planetary systems. Like in the generic data modelling problems, a fit to radial velocity data series has a set of unknown parameters of which parametric derivatives have to be known by both the regression methods and the estimations for the uncertainties. In this paper an algorithm is described that aids the computation of such derivatives in case of when planetary perturbations are not neglected. The application of the algorithm is demonstrated on the planetary systems of HD~73526, HD~128311 and HD~155358. In addition to the functions related to radial velocity analysis, the actual implementation of the algorithm contains functions that computes spatial coordinates, velocities and barycentric coordinates for each planet. These functions aid the joint analysis of multiple transiting planetary systems, transit timing and/or duration variations or systems where the proper motion of the host star is also measured involving high precision astrometry. The practical implementation related to the above mentioned problems features functions that make these kind of investigations rather simple and effective. ", "introduction": "\\label{sec:introduction} As of this writing, $36$ multiple planetary systems are known around main sequence stars. Most of these systems bear two detected planets, $8$ of them have $3$ and $2$ of them have $4$ planets while the star 55~Cnc has 5 companions\\footnote{See e.g. \\texttt{http://exoplanet.eu} for an up-to-date list.}. With the exception of HR~8799 \\citep{marois2008}, all of the detections are based on or confirmed by the measurements of the radial velocity (RV) variations of the host stars\\footnote{The planetary system around HR~8799 with 3 confirmed planets has been detected by direct imaging.}. The planet HAT-P-13b \\citep{bakos2009} also transits its host star. The detection of this planet was based on transit photometry while the second companion in this system has been revealed by radial velocity measurements. In general, analysis of RV variations constrains the mass ($m$) of planets by a lower limit. Namely, only the quantity $m\\sin i$ is determined by RV data where $i$ is the orbital inclination (relative to the tangential plane of sky). With the exception of transiting planets and systems were the spatial motion of the host star is detected via astrometry, there is no direct evidence for the actual value of the orbital inclination (and therefore the mass of the planet). In case of transiting systems, inclinations are constrained by measuring the impact parameter from the light curves \\citep[see e.g.][]{pal2010}, while astrometry yields not only the inclination but the orientation of the orbital plane as well \\citep{bean2009,benedict2010,mcarthur2010}. The planetary system around GJ~876 is the only known one where mutual inclination is also detected with a 2-$\\sigma$ confidence \\citep{bean2009}. A great advantage of multiple planetary systems is the possibility of detecting mutual perturbations via the deflection of RV values from the purely Keplerian solution \\citep[see e.g.][]{laughlin2001}. Therefore, precise analysis of accurate RV series may yield to an acceptable constraint for both the inclination and the planetary masses. Additionally, planet-planet interactions depend on the mutual inclinations, thus more complex models (such as non-coplanar orbits) for the whole planetary systems can be investigated. Like in the majority of data modelling problems, RV variations (of the host star) in single or multiple planetary systems are modelled with a function of a few external (unknown) parameters. These parameters include the orbital elements and the masses of the planets as well as the barycentric velocity of the host star. For most of the regression methods involved in data modelling \\citep[see e.g. the Levenberg-Marquard algorithm,][]{press1992}, and for the analytic estimation of the covariances, uncertainties and correlations of the model parameters \\citep[see e.g.][]{finn1992,pal2009}, the partial derivatives of the model functions (with respect to the model parameters) have to be known in advance. The simplest way of RV curve modelling does not take into account the mutual interactions between the planets and characterizes the observed RV variations as a sum of independent Keplerian models. \\cite{wright2009b} describes an algorithm detailing the efficient computation of the parametric derivatives of RV model functions where the mutual planetary perturbations are neglected. The main objective of this paper is to present an algorithm that calculates parametric derivatives when the planet-planet interactions are also taken into account. The structure of the paper is as follows. In the next section, we describe the mathematical tools used to construct the algorithm itself (including the discussion of optimal orbital parameterization, and the numerical integration). In Section~\\ref{sec:implementation} the practical implementation is detailed while in Section~\\ref{sec:applications} we demonstrate the usage of the algorithm for three specific multiple planetary systems. The results are summarized in the last section. ", "conclusions": "In this paper we described an algorithm based on the Lie-integration method that efficiently computes the parametric derivatives of radial velocity model functions for multiple planetary systems when the planet-planet interactions are also taken into account. The analysis of these systems yields more accurate constrains for planetary masses since the orbital and mutual inclinations can also be derived if precise radial velocity data are available. Additionally, the presented analytic formulae and integration method aid to plan observation schedules in order to optimize the telescope time utilization in order to detect planetary perturbations. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "1005/1005.0637_arXiv.txt": { "abstract": "Gas-phase complex organic molecules have been detected toward a range of high- and low-mass star-forming regions at abundances which cannot be explained by any known gas-phase chemistry. Recent laboratory experiments show that UV irradiation of CH$_3$OH-rich ices may be an important mechanism for producing complex molecules and releasing them into the gas-phase. To test this ice formation scenario we mapped the B1-b dust core and nearby protostar in CH$_3$OH gas using the IRAM 30m telescope to identify locations of efficient non-thermal ice desorption. We find three CH$_3$OH abundance peaks tracing two outflows and a quiescent region on the side of the core facing the protostar. The CH$_3$OH gas has a rotational temperature of $\\sim$10~K at all locations. The quiescent CH$_3$OH abundance peak and one outflow position were searched for complex molecules. Narrow, 0.6-0.8 km s$^{-1}$ wide, HCOOCH$_3$ and CH$_3$CHO lines originating in cold gas are clearly detected, CH$_3$OCH$_3$ is tentatively detected and C$_2$H$_5$OH and HOCH$_2$CHO are undetected toward the quiescent core, while no complex molecular lines were found toward the outflow. The core abundances with respect to CH$_3$OH are $\\sim$2.3\\% and 1.1\\% for HCOOCH$_3$ and CH$_3$CHO, respectively, and the upper limits are 0.7--1.1\\%, which is similar to most other low-mass sources. The observed complex molecule characteristics toward B1-b and the pre-dominance of HCO-bearing species suggest a cold ice (below 25~K, the sublimation temperature of CO) formation pathway followed by non-thermal desorption through e.g. UV photons traveling through outflow cavities. The observed complex gas composition together with the lack of any evidence of warm gas-phase chemistry provide clear evidence of efficient complex molecule formation in cold interstellar ices. ", "introduction": "Complex organic molecules have been detected toward a range of astrophysical environments, including low-mass protostars \\citep{vanDishoeck95, Cazaux03, Bottinelli07}; however, the origins of these complex molecules as well as their fates are uncertain. Commonly-suggested formation routes for the detected molecules include various gas-phase reactions starting with thermally evaporated CH$_3$OH ice, atom-addition reactions on dust grains, and UV- and cosmic ray-induced chemistry in the icy grain mantles that form during the pre-stellar stages \\citep[][for a review]{Charnley92, Nomura04, Herbst09}. The focus is currently on an ice formation pathway \\citep[e.g.][]{Garrod06, Garrod08} because of the failures of gas phase chemistry to explain the observed abundances of some of the most common complex molecules toward low-mass protostars, especially HCOOCH$_3$. Recent experiments on the photochemistry of CH$_3$OH-rich ices have shown that 1) UV irradiation of CH$_3$OH ices at 20--70~K result in the production of large amounts of the complex molecules observed around protostars and 2) the chemistry has a product branching ratio which is temperature and ice composition dependent \\citep{Oberg09d}. A key result is that HCOOCH$_3$ and other HCO-X ices are only abundantly produced in CO:CH$_3$OH mixtures. A similar result has been reported for proton-bombarded CH$_3$OH and CH$_3$OH:CO ices \\citep{Bennett07b}. CO:CH$_3$OH ice mixtures are probably common since CH$_3$OH forms from hydrogenation of CO \\citep[e.g.][]{Watanabe03,Cuppen09}. CO evaporates at 17--25~K on astrophysical time scales \\citep{Bisschop06}, and thus HCO-X ices will mainly form in cold, UV-exposed ices. At higher temperatures, closer to the protostar, UV irradiation of the remaining pure CH$_3$OH ice will instead favor the production of C$_2$H$_5$OH, CH$_3$OCH$_3$ and (CH$_2$OH)$_2$. In light of this proposed formation scenario, protostars rich in CH$_3$OH ice are natural targets when searching for complex molecule sources. The low-mass protostar B1-b is such a source. From previous SCUBA and 3 mm continuum maps, B1-b consists of two cores B1-bN and B1-bS, separated by 20'' \\citep{Hirano99}. The cores are similar with $T_{\\rm dust}\\sim18$~K, $M=1.6-1.8\\:M_{\\odot}$ and $L_{\\rm bol}=2.6-3.1\\:L_{\\odot}$. The {\\it Spitzer Space Telescope} only observed one protostar, to the south-west of the B1-bS dust core at 03:33:20.34, +31:07:21.4 (J2000), but it was still named B1-b -- to avoid confusion it will be referred to as the `protostar' position. From the {\\it Spitzer} $c2d$ (From Molecular Cores to Planet Forming Disks) ice survey the B1-b protostar has a CH$_3$OH ice abundance of 11\\% with respect to H$_2$O ice \\citep{Boogert08}, corresponding to $\\sim5\\times10^{-6}$ with respect to H$_2$. The fractional abundance of H$^{13}$CO$^+$ decreases by a factor of 5 toward the SCUBA core, indicative of CO freeze-out \\citep{Hirano99}. A HCO-dominated complex chemistry is thus expected where the ice is exposed to UV radiation. The B1-b region is complicated by a number of outflows, which may enhance the UV field through shocks and open cavities through which stellar UV photons can escape \\citep{Jorgensen06, Walawender09, Hiramatsu10}. The outflow of most interest to this study runs in the south-west direction from B1-b protostar, though it is not known whether B1-b is its source or just happens to lie in its path. The outflow is not observed north-east of the B1-b protostar, where the B1-b SCUBA core is situated. It is therefore unclear whether the outflow terminates at the protostar or continues into the core, hidden from view. If it does penetrate into the SCUBA core, this may increase the UV flux in the region by orders of magnitude, enhancing both the UV photochemistry and photodesorption of ices. The latter is important, since once formed, the complex molecules must be (partly) released into the gas phase to be observable at millimeter wavelengths. The second reason for targeting B1-b is the detection of a large CH$_3$OH gas phase column density of $\\sim$2.3--2.5$\\times10^{14}$ cm$^{-2}$ toward the protostar \\citep{Oberg09a,Hiramatsu10} and thus ice evaporation. The narrow line width and low excitation temperature of the observed CH$_3$OH suggest non-thermal ice evaporation -- the abundances are consistent with UV photodesorption. Assuming that complex ice chemistry products desorb non-thermally at an efficiency similar to CH$_3$OH ($\\sim$10$^{-3}$ molecules per incident UV photon \\citep{Oberg09d}), the protostellar environment should contain a cold gas-phase fingerprint of the complex ice composition; B1-b may offer the first possibility to observe the earliest stages of complex molecule formation, untainted by either warm ice chemistry or gas-phase processing. Addressing these predictions, we present IRAM 30-m telescope observations of a $80''\\times80''$ CH$_3$OH map of the B1-b protostellar envelope and nearby dust cores, followed by a search for HCOOCH$_3$, CH$_3$CHO, CH$_3$OCH$_3$, C$_2$H$_5$OH and HOCH$_2$CHO toward two of the identified CH$_3$OH peaks. The CH$_3$OH gas abundances and the complex abundances and upper limits with respect to CH$_3$OH are then discussed in terms of different ice desorption mechanisms and complex molecule formation scenarios. The results are finally compared with complex molecule observations toward other low-mass protostars and outflows. ", "conclusions": "\\begin{enumerate} \\item Cold CH$_3$OH gas (excitation temperature of $\\sim$10~K) is abundant and widespread toward B1-b. Quiescent CH$_3$OH is most abundant in between the B1-b SCUBA dust core and the B1-b protostar, indicative of UV photodesorption of ice on the side of the quiescent dust core because of radiation from the protostar escaping through outflow cavities. \\item Cold HCOOCH$_3$ and CH$_3$CHO are both detected and CH$_3$OCH$_3$ is tentatively detected toward the quiescent CH$_3$OH peak. No complex molecules are observed toward an outflow associated with B1-b, but 3-$\\sigma$ upper limits with respect to CH$_3$OH are only slightly lower compared to the quiescent core. In addition an asymmetry in the CH$_3$CHO emission lines is suggestive of some contribution from a small outflow included in the beam. \\item Assuming a 10~K excitation temperature the calculated beam-averaged abundances are 2.3\\% for HCOOCH$_3$, 1.1\\% for CH$_3$CHO, $\\lesssim$0.8\\% for CH$_3$OCH$_3$ and $<$1.0-1.1\\% for C$_2$H$_5$OH and HOCH$_2$CHO with respect to CH$_3$OH. \\item Building on recent experiments, the observations of large abundances of HCO-containing molecules are explained by UV/cosmic ray processing of cold CH$_3$OH:CO ice followed by non-thermal desorption of a fraction of the produced organic ice. \\item The beam-averaged abundances with respect to CH$_3$OH and the ratios between different complex molecules are similar to other single-dish observations toward low-mass protostars. In contrast resolved observations of the cores around protostars are more abundandant in complex molecules without a HCO-group. This is consistent with a complex ice chemistry that evolves from HCO-rich to HCO-poor as the ice warms up and the CO-ice evaporates when the icy dust falls toward the protostar. \\end{enumerate} Including B1-b, there are still only a handful of observations of complex molecules toward low-mass protostars. In addition few complex molecules are detected toward each source. This makes it difficult both to determine the prevalence of a complex organic chemistry around low-mass protostars and to firmly establish the main formation path of such molecules. The detections of complex molecules in the vicinity of the CH$_3$OH-ice-rich protostar suggests that targeting other CH$_3$OH-ice rich protostars may increase our sample of complex molecule detections significantly. Maybe more important from a formation pathway point of view is to increase our understanding on how the complex chemistry varies between warm, luke-warm and cold regions i.e. how it varies with distance from the protostar. This should be pursued both on larger scales with single dish observations and on smaller scales with interferometers -- similarly to what has been done toward IRAS 16293 and NGC1333 2a \\citep[e.g.][]{Jorgensen05,Bisschop08}. In the meantime the complex molecules observed toward B1-b, and their exceptionally low excitation temperatures and line widths, provides clear evidence, for the efficient formation of complex ices during star- and planet-formation." }, "1005/1005.5012_arXiv.txt": { "abstract": "{The physical conditions of the solar photosphere change on very small spatial scales both horizontally and vertically. Such a complexity may pose a serious obstacle to the accurate determination of solar magnetic fields.} {We examine the applicability of Milne-Eddington (ME) inversions to high spatial resolution observations of the quiet Sun. Our aim is to understand the connection between the ME inferences and the actual stratifications of the atmospheric parameters.} {We use magnetoconvection simulations of the solar surface to synthesize asymmetric Stokes profiles such as those observed in the quiet Sun. We then invert the profiles with the ME approximation. We perform an empirical analysis of the heights of formation of ME measurements and analyze the uncertainties brought about by the ME approximation. We also investigate the quality of the fits and their relationship with the model stratifications.} {The atmospheric parameters derived from ME inversions of high-spatial resolution profiles are reasonably accurate and can be used for statistical analyses of solar magnetic fields, even if the fit is not always good. We also show that the ME inferences cannot be assigned to a specific atmospheric layer: different parameters sample different ranges of optical depths, and even the same parameter may trace different layers depending on the physical conditions of the atmosphere. Despite this variability, ME inversions tend to probe deeper layers in granules as compared with intergranular lanes. } {} ", "introduction": "\\label{cap5:sec:intro} The solar spectrum carries information about the properties of our star. In general, a broad range of atmospheric layers contribute to the shape of the spectral lines, making it difficult to extract this information directly. Both the measurement process and the method of analysis introduce uncertainties in the physical quantities retrieved from the observations. Sources of error are photon noise and instrumental effects like limited spectral resolution, wavelength sampling, and angular resolution, but also the simplifications and approximations of the model used to interpret the measurements. In this paper we want to evaluate the merits of Milne-Eddington (ME) inversions for the analysis of the polarization line profiles emerging from the solar atmosphere. The ME approximation does not account for vertical variations of the parameters \\citep{1956PASJ....8..108U, Rakk1,Rakk2}, so it cannot accurately describe the solar plasma when rapid changes in height are present. What is, then, the significance of the ME parameters? To answer this question it is necessary to simulate the processes of line formation and data inversion. Usually one prescribes a set of model atmospheres, performs spectral synthesis calculations, inverts the synthetic profiles, and compares the results with the known input. A common approach is to use ME models both to generate the spectra and to invert them \\citep[e.g.,][]{nortonetal2006, borreroetal2007}. In that case the analysis is internally consistent and the uncertainties of the retrieved ME parameters are mostly due to the noise and, to a smaller extent, to the convergence of the algorithm, provided that the spectral resolution and wavelength sampling are appropriate. Uncertainties caused by photon noise are known as statistical errors and can be evaluated by means of numerical tests or, more efficiently, by using ME response functions \\citep{2007A&A...462.1137O,2010ApJ...711..312D}. However, they represent only a small fraction of the total error. Another source of error is the very assumption of height-independent parameters, which leads to symmetric line profiles. What happens when realistic (i.e., asymmetric) Stokes spectra are analyzed in terms of ME models? Do the uncertainties of the retrieved parameters increase significantly? Answering these questions is the aim of the present work. \\begin{figure*}[!t] \\centering \\resizebox{0.425\\hsize}{!}{\\includegraphics{Fig1a.ps}} \\resizebox{0.425\\hsize}{!}{\\includegraphics{Fig1b.ps}} \\resizebox{0.425\\hsize}{!}{\\includegraphics{Fig1c.ps}} \\resizebox{0.425\\hsize}{!}{\\includegraphics{Fig1d.ps}} \\caption{Magnetic field strength, inclination, azimuth, and LOS velocity at $\\tau=1$ in a simulation snapshot with $\\langle B \\rangle=140$~G. Negative velocities represent upflows.} \\label{Fig1} \\end{figure*} A first study of the capabilities and limitations of ME inversions was carried out by \\cite{1998ApJ...494..453W} using simple (non-ME) model atmospheres. They made a quantitative comparison of results obtained with the ME code of the High Altitude Observatory \\citep{lites2,lites1} and the SIR code \\citep[Stokes Inversion based on Response functions;][]{1992ApJ...398..375R}. The main conclusion of their work was that ME inversions provide accurate values of the physical parameters averaged along the line of sight, at least when the stratifications are smooth. More recently, \\cite{2007MmSAI..78..166K} have investigated whether the magnetic field stratification itself can be determined reliably through inversion of high resolution data. To that end, they synthesized the Stokes profiles of the \\ion{Fe}{i} 630~nm lines with the help of MHD models and inverted them with SIR, allowing for vertical gradients of the atmospheric parameters. The analysis showed that SIR is able to recover the actual magnetic stratification for fields as weak as 50~G if no noise is present. This work extends the results of \\cite{1998ApJ...494..453W} to the case in which the stratifications are not smooth. To determine the uncertainties associated with ME inversions of asymmetric Stokes profiles we use state-of-the-art magnetohydrodynamic simulations (Sect.~\\ref{MHD}). Our goal is to describe the solar photosphere as realistically as possible. We construct model atmopheres from the simulations and synthesize the Stokes profiles of the \\ion{Fe}{i} 630.2~nm lines emerging from them (Sect.~\\ref{synthesis}). The SIR code is used for the spectral synthesis, so the profiles are asymmetric. Finally, we apply a ME inversion to the data (Sect.~\\ref{inversion}). In our numerical experiments, the spatial sampling of the MHD models, $0\\farcs 0287$, is preserved. There are two reasons why we neglect the effects of solar instrumentation: first, they have already been studied in the past (e.g., Orozco Su\\'arez et al. 2007, 2010); second, this sampling is close to critical for the observations to be delivered by large telescopes like the Advanced Technology Solar Telescope\\footnote{http://atst.nso.edu/} \\citep{wagner2006} and the European Solar Telescope\\footnote{http://www.iac.es/project/EST/} \\citep{collados2008}. We invert the profiles with the MILOS code \\citep{2007A&A...462.1137O}.\\footnote{MILOS is programmed in IDL and can be downloaded from our website, http://spg.iaa.es/download.asp} A direct comparison of the retrieved and true parameters allows us to determine the effective ``heights of formation'' of the ME parameters (Sect.~\\ref{results}) and to quantify the errors caused by the ME approximation (Sect.~\\ref{results2}). The conclusions of our work are given in Sect.~\\ref{conclus}. For completeness, the results of ME inversions are compared with those of tachogram/magnetogram-like analyses in the Appendix. ", "conclusions": "\\label{conclus} In this paper we have analyzed radiative MHD simulations of the quiet Sun. We have used them to synthesize Stokes line profiles in three different spectral regions (525.0, 617.3, and 630.2~nm). The comparison of the synthetic profiles with the FTS atlas suggests that the simulations describe quite satisfactorily the physical conditions of the solar photosphere, although the MHD models are slightly hotter than the HSRA around $\\tau=1$. After synthesizing the Stokes profiles, the applicability of ME inversions to high spatial resolution observations has been examined. We have considered the case of the \\ion{Fe}{i} lines at 630.2~nm. The analysis of the profiles by means of ME inversions has allowed us to characterize the uncertainties that can be expected from the ME approximation. For this reason, the synthetic profiles were not degraded by noise, instrumental effects, or spatial resolution. The main limitation of ME inversions is that they provide constant atmospheric parameters, whereas the MHD models feature physical properties that change with height. This limitation means that ME models are unable to reproduce spectral line asymmetries. Consequently, the ME inferences cannot be assigned to a specific optical depth. Depending on the conditions of the atmosphere, the retrieved ME parameters sample different layers. However, from a statistical point of view we conclude that ME inversions provide fair estimates of the physical conditions prevailing at $\\log\\tau \\sim -1$. The rms uncertainty is smaller than 30~G for the magnetic field strength, 13\\degree\\/ and 20\\degree\\/ for the field inclination and azimuth, and 500~\\ms\\/ for the LOS velocity. Thus, ME inversions are appropriate for statistical analyses of the solar photosphere. This being said, it is important to realize that the errors may be large for individual pixels, even if the best-fit profiles reproduce the observations satisfactorily (the field strength in case {\\em a} of Fig.~\\ref{Fig5} is a good example of this). Finally, we want to stress that the uncertainties associated with the ME approximation are larger than those due to photon noise \\citep{2006ASPC..358..197O,2010ApJ...711..312D}. However, the noise has another undesirable effect: it hides the weaker polarization signals. This fact has not been considered in our study. \\begin{figure*}[!t] \\centering % \\resizebox{0.98\\hsize}{!}{\\includegraphics{FigApp.eps}} \\caption{Scatter plots of the longitudinal magnetic field, $B_{z}$ (first two rows), of the LOS velocity (third row), and of the magnetic field inclination (fourth row), as functions of their corresponding values in the MHD simulations at $\\log\\tau = -1$. The first column refers to results obtained with the 630.15 nm line; the second column displays results obtained with the 630.25 nm line; results with the two lines at the same time are shown in the third column. Color and line codes are the same as in Fig.\\ \\ref{Fig8}. The mean value of the rms difference between the inferences and the simulations at $\\log\\tau = -1$ is given in the upper left corner of each panel.} \\label{FigApp} \\end{figure*} \\appendix" }, "1005/1005.5362_arXiv.txt": { "abstract": "The bright and highly variable X-ray and radio source known as Cygnus X-3 was among the first X-ray sources discovered, yet it remains in many ways an enigma. Its known to consist of a massive, Wolf-Rayet primary in an extremely tight orbit with a compact object. Yet one of the most basic of parameters - the mass of the compact object - is not known. Nor is it even clear whether its is a neutron star or a black hole. In this Paper we present our analysis of the broad-band high-energy continua covering a substantial range in luminosity and spectral morphology. We apply these results to a recently identified scaling relationship which has been demonstrated to provide reliable estimates of the compact object mass in a number of accretion powered binaries. This analysis leads us to conclude that the compact object in Cygnus X-3 has a mass greater than $4.2M_\\odot$ thus clearly indicative of a black hole and as such resolving a long-standing issue. The full range of uncertainty in our analysis and from using a range of recently published distance estimates constrains the compact object mass to lie between $4.2M_\\odot$ and $14.4M_\\odot$. Our favored estimate, based on a 9.0 kpc distance estimate is $\\sim 10 M_\\odot$ with the error margin of 3.2 solar masses. This result may thus pose challenges to shared-envelope evolutionary models of compact binaries, as well as establishing Cygnus X-3 as the first confirmed accretion-powered galactic gamma-ray source. ", "introduction": "The bright X-ray binary source known as Cygnus X-3 (herein Cyg X-3) was discovered over 40 years ago by \\cite{gia67}, and has been the subsequent focus of extensive study. It nonetheless remains enigmatic in that the nature of the compact object; a neutron star or black hole has not been unambiguously determined, its high-amplitude intensity and spectral variations in the X-ray and radio bands are not well understood, and its low-amplitude and featureless power-density spectrum does not resemble other known neutron star or black hole X-ray binaries, e.g. \\cite{McC99}. Many of these unique attributes are due to the nature of the donor star, widely believed to be a Wolf Rayet (WR) star, e.g. \\cite{vankerk92} . It is also the strongest radio source among X-ray binaries, with a quiescent flux $<100$ mJy, a flux of 0.1 Jy flaring to several Jy in high-activity states. It has recently been detected in the ~0.1-1-GeV domain with the Fermi Gamma-Ray Space Telescope \\citep{abdo09} and is thus among a small number of gamma-ray bright X-ray binaries. Its orbital period of 4.8 hours is typical for a low-mass binary, but its WR donor star could have a mass upwards of 30 M$_\\odot$, thus making it an extremely tight binary system, all the more so if the compact star is a black hole. Distance estimates to Cyg X-3 range from 7.2 to 9.3 kpc; see details in \\cite{lzt09}. Here we adopt the distance of 9 kpc, inferred from dust-halo scattering measurements \\citep{pred00} however, we consider the full range of these distance determinations in evaluating the uncertainty of our result. Such a large distance means that it is viewed thorough many magnitudes of visual extinction. It is in addition likely to be enshrouded in the dense wind environment of the WR donor star. This combination of high local and Galactic line-of-sight column densities has rendered the usual photometric and spectroscopic techniques for deriving a binary solution intractable, and as a result, no reliable mass estimate for the compact has emerged. The local absorption has also limited conclusions based on X-ray spectroscopy, which may be further complicated by the presence of a synchrotron component associated with the radio and gamma-ray emission and the effects of Compton downscattering. While hard-to-soft spectral transitions suggestive of a black hole binary are evident, e.g. \\cite{hz09} they cannot always be easily reconciled within the context of familiar black-hole low-hard to high-soft state transition patterns, nor do they resemble neutron-star Z or atoll source spectral behavior. No pulsations are detected, which would be a clear signature of a magnetized neutron star, however, it is possible that pulsed (radio or X-ray) emission could be present in the system but unobservable due to the effects of dense ambient plasma. It is noted that this type of scenario has been proposed and is considered credible in explanation of another radio-loud, gamma-ray-bright X-ray binary LSI +61 303. Previous estimates of the compact object mass based on a variety of methods have been attempted, but the results tend to be highly uncertain. For example Schmutz, Geballe and Schild (1996) employing IR spectroscopy obtained a likely mass in the range of $7-40 M_\\odot$ with a best estimate of $17 M_\\odot$. Hanson, Still and Fender (2000) constructed a radial velocity curve from which derive a mass function of $0.027 M_\\odot$, which leads to a blackhole mass of less than $10 M_\\odot$ but could still accommodate a neutron star depending on the true value of the binary inclination. Vilhu et al. (2009) assess orbital modulation of X-ray emission lines and find a most likely compact object masses between 2 and $8 M_\\odot$ while Hjalmarsditter et al. (2008) estimate a $\\sim 30 M_\\odot$ black hole based on interpretation of color-luminosity diagrams. Our approach involves a multi-epoch spectral analysis and assessment of the inferred parameters. \\cite{st09} and \\cite{ts09}, hereafter ST09 and TS09 respectively established well defined correlations between certain temporal and spectral parameters. Furthermore, these correlation curves observed in different sources exhibited differing patterns for different binaries, but that the variety of those patterns was limited, and thus scalable. Recently, the application of this method was extended to study of an another class of X-ray source, the ULX NGC 5408 X-1 (Strohmayer and Mushotzky 2009). That study resulted in tangible evidence for the existence of an intermediate mass object in that system. The ubiquitous nature of these correlations led to the suggestion that the underlying physical processes leading to the observed variability properties are closely tied to the Comptonizing media. Furthermore, they vary in a well defined manner as the source makes a transition between spectral states. The fact that the same correlations are seen in many sources, which vary widely in both luminosity (and thus presumably with mass accretion rate) and spectral state, suggests a common set of underlying physical conditions. TS09 showed that in GRS 1915+105 the photon index monotonically increases with disk mass accretion rate, followed by saturation. The radio luminosity does not correlate with low frequency quasi-periodic oscillations or X-ray luminosity for the entire range of spectral states, from low-hard to high-soft through intermediate. In this Paper we present the results of a broad-band multi-epoch spectral analysis of Cyg X-3, applied to a novel approach for compact object mass determination, using photon index$-$disk mass accretion rate correlation first implemented by ST09. We compiled a data base of some 35 observations of Cyg X-3, primarily from the RXTE satellite, applied similar modeling methods, and studied the inferred parameter interdependencies. From the results of that analysis, we have applied the scaling laws derived in ST09 deriving an estimate of the compact object mass, which we find to be consistent with a black hole. In \\S 2 we briefly discuss our methodology for mass-determination, which is described in detail elsewhere. In \\S 3 we describe our database compilation and analysis techniques, in \\S 4 we present our results, draw comparison to similar results previously calibrated from objects with independent mass estimates, and in \\S 5 we summarize our conclusions. ", "conclusions": "We have studied a large set of data covering a number of spectral-state transitions and intensity variations in the galactic binary Cyg X-3. We find that for about half of the 35 ~2-100-keV spectra analyzed the observations are well represented by our modeling approach - an absorbed BMC model including a iron line and edge components and a high-energy cutoff term. We then examined the correlation between the inferred photon indices of the Comptonized spectral component and the mass accretion rate, a proxy for which is the normalization term of our model. Based on those results we have applied the scaling method employing previously determined results for the well known Galactic binary GRS 1915+105 as a reference dataset. The calibration of our method for GRS 1915+015 is supported by the dynamical mass determination of \\cite{gre01}. This analysis led us to a lower-limit mass determination of about $ 4.2M_\\odot$ and a best-estimate value of $\\sim 10.1\\pm3.2 M_\\odot$ assuming a 9.0 kpc distance. The previous success of this scaling method for the BH mass determination strongly supports our results for Cyg X-3. Another result of our study is that a basic prediction of the theory of the converging inflow is supported by the observations of the index-mass accretion rate correlation seen in various black-hole X-ray binaries. Specifically, we argue that the spectral-index vs $\\dot m$ flattening, or saturation, seen is an observational signature of the presence of a converging inflow. The implication is that {\\it this effect provides robust observational evidence for the presence of a black hole in Cyg X-3.}" }, "1005/1005.2297_arXiv.txt": { "abstract": "{} {Within the framework of the \\texttt{HERM33ES} Key Project, using the high resolution and sensitivity of the \\textit{Herschel} photometric data, we study the compact emission in the Local Group spiral galaxy M\\,33 to investigate the nature of the compact SPIRE emission sources. We extracted a catalogue of sources at 250\\,\\micron\\ in order to investigate the nature of this compact emission. Taking advantage of the unprecedented \\textit{Herschel} resolution at these wavelengths, we also focus on a more precise study of some striking H$\\alpha$ shells in the northern part of the galaxy.} {We present a catalogue of 159 compact emission sources in M\\,33 identified by SExtractor in the 250\\,\\micron\\ SPIRE band that is the one that provides the best spatial resolution. We also measured fluxes at 24\\,\\micron\\ and H$\\alpha$ for those 159 extracted sources. The morphological study of the shells also benefits from a multiwavelength approach including H$\\alpha$, far-ultraviolet from \\textit{GALEX}, and infrared from both \\textit{Spitzer} IRAC 8\\,\\micron\\ and MIPS 24\\,\\micron\\ in order to make comparisons.} {For the 159 compact sources selected at 250\\,\\micron, we find a very strong Pearson correlation coefficient with the MIPS 24\\,\\micron\\ emission ($r_{24} = 0.94$) and a rather strong correlation with the H$\\alpha$ emission, although with more scatter ($r_{\\rm H\\alpha} = 0.83$). The morphological study of the H$\\alpha$ shells shows a displacement between far-ultraviolet, H$\\alpha$, and the SPIRE bands. The cool dust emission from SPIRE clearly delineates the H$\\alpha$ shell structures.} {The very strong link between the 250\\,\\micron\\ compact emission and the 24\\,\\micron\\ and H$\\alpha$ emissions, by recovering the star formation rate from standard recipes for \\HII\\ regions, allows us to provide star formation rate calibrations based on the 250\\,\\micron\\ compact emission alone. The different locations of the H$\\alpha$ and far-ultraviolet emissions with respect to the SPIRE cool dust emission leads to a dynamical age of a few Myr for the H$\\alpha$ shells and the associated cool dust.} ", "introduction": "% Within the framework of the open time key project ``{\\it Herschel} M\\,33 extended survey ({\\tt HERM33ES})'', we are studying the galaxy M\\,33 to understand the origin of various diagnostic lines in heating and cooling and other processes in the interstellar medium (ISM). We refer the reader to \\citet{2010A&A...501L...1K} for more details about the overall {\\tt HERM33ES} project goals, as well as for a first presentation of the PACS and SPIRE maps of the entire galaxy, together with spatially averaged spectral energy distributions. The local group, late-type, spiral galaxy --M\\,33-- is an ideal target for studying the detailed processes of star formation (SF). Indeed, the distance, 840\\,kpc \\citep{1991ApJ...372..455F}, and inclination, 56 degrees \\citep{1994ApJ...434..536R}, of M\\,33 allow us to reach an unprecedented spatial resolution in the far-infrared of an external spiral galaxy and investigate the phenomena in the interstellar medium which lead to the SF. With {\\it Herschel} \\citep{2010A&A...001L...1P}, we are able to resolve 85\\,pc, 109\\,pc, and 170\\,pc structures of the cool dust in the SPIRE 250, 350, and 500\\,\\micron\\ bands, respectively, which is the typical spatial scale of giant \\HII\\ regions and their complexes. In this article, we follow a two-fold approach in order to study the nature of the compact 250\\,\\micron\\ emission and take advantage of the resolution and sensitivity reached by {\\it Herschel} to resolve the cool dust emission associated with \\HII\\ regions. ", "conclusions": "For the Local Group spiral galaxy M\\,33, using the unprecedented resolution and sensitivity of the {\\it Herschel} SPIRE photometric data, we focus on the compact emission and conclude the following. \\begin{itemize} \\item We created a catalogue of 159 sources from the SPIRE 250\\,\\micron\\ emission using a high detection threshold to focus on the compact emission of the cool dust; the compact emission sources associated to the cool dust follow the spiral pattern of the galaxy. \\item Very high Pearson correlation coefficients, $r_{24} = 0.94$ for the 24\\,\\micron\\ emission and $r_{\\rm H\\alpha} = 0.83$ for the H$\\alpha$ emission, confirm that the 159 sources are closely linked to SF. \\item By using standard (H$\\alpha$+24\\,\\micron\\ and 24\\,\\micron\\ alone) SFR recipes for \\HII\\ regions provided by \\citet{2007ApJ...666..870C,2010ApJ...714.1256C}, we are able to calibrate the 250\\,\\micron\\ compact emission as an SFR tracer and provide conversion factors that can be used in further studies (Eqs.~\\ref{eq:sfrMix} and \\ref{eq:sfr24}). \\item The morphological study of a set of three H$\\alpha$ shells shows that there is a displacement between far-ultraviolet and the SPIRE bands, while the H$\\alpha$ structure is in general much more coincident with the cool dust. \\item The different locations of the H$\\alpha$ and far-ultraviolet emissions with respect to the SPIRE cool dust emissions leads to a dynamical age of a few Myr for a set of H$\\alpha$ shells and the associated cool dust. \\end{itemize} \\vspace{-.4cm}" }, "1005/1005.4687_arXiv.txt": { "abstract": "We present an analysis of the host galaxy dependencies of Type Ia Supernovae (SNe Ia) from the full three year sample of the SDSS-II Supernova Survey. We re-discover, to high significance, the strong correlation between host galaxy type and the width of the observed SN light curve, i.e., fainter, quickly declining SNe Ia favor passive host galaxies, while brighter, slowly declining Ia's favor star-forming galaxies. We also find evidence (at between 2 to $3\\sigma$) that SNe Ia are $\\simeq0.1\\pm0.04$ magnitudes brighter in passive host galaxies, than in star--forming hosts, after the SN Ia light curves have been standardized using the light curve shape and color variations: This difference in brightness is present in both the SALT2 and MCLS2k2 light curve fitting methodologies. We see evidence for differences in the SN Ia color relationship between passive and star--forming host galaxies, e.g., for the MLCS2k2 technique, we see that SNe Ia in passive hosts favor a dust law of $R_V = 1.0\\pm0.2$, while SNe Ia in star-forming hosts require $R_V = 1.8^{+0.2}_{-0.4}$. The significance of these trends depends on the range of SN colors considered. We demonstrate that these effects can be parameterized using the stellar mass of the host galaxy (with a confidence of $>4\\sigma$) and including this extra parameter provides a better statistical fit to our data. Our results suggest that future cosmological analyses of SN Ia samples should include host galaxy information. ", "introduction": "Over the last decade, Type Ia Supernovae (SNe Ia) have become important cosmological probes as they can be used to measure distances to high redshift ($z\\lesssim1.5$). In recent years, numerous samples of SN Ia have been compiled, e.g., CSP \\citep{2006PASP..118....2H}, SNLS \\citep{2006A&A...447...31A}, ESSENCE \\citep{2007ApJ...666..694W}, SDSS \\citep{2008AJ....135..338F}, CfA \\citep{2009ApJ...700.1097H}, and combined, we are approaching $\\sim1000$ spectroscopically-confirmed SNe Ia available for cosmological analysis \\citep{2010arXiv1004.1711A}. With such large samples, it is becoming increasingly important to understand the systematic uncertainties (photometric calibration, SN color variations, etc.) associated with using SNe Ia for cosmology, including any additional physical parameters that could reduce the intrinsic scatter of the population. One such parameter could be related to the environment of the supernova. First and foremost, one would expect differences in the colors of SNe Ia based on the different dust content of their hosts, i.e., potential variations in local circumstellar dust around the progenitor star \\citep{wang2005, 2008ApJ...686L.103G} and/or differences in the global dust content of different galaxy types\\footnote{Dust in our own Galaxy is usually corrected for using the dust maps of \\citet{1998ApJ...500..525S}}. Despite these concerns, most analyses account for dust during the fitting of the supernova light curves by assuming a single absorption law ($R_V$) for all SNe, which minimizes the scatter around the Hubble diagram \\citep{1998A&A...331..815T}. This process, however, has led to a dust law that is significantly different from the canonical value for our Galaxy ($R_V \\approx 3.1$), e.g., \\citet{2007ApJ...664L..13C} find $R_V \\approx 1$ for nearby SNe Ia, while \\citet{2009ApJS..185...32K} obtained a best fit of $R_V = 2.18 \\pm 0.14{\\rm (stat)} \\pm {\\rm 0.48(sys)}$ for the first year sample from the Sloan Digital Sky Survey II (SDSS-II) Supernova Survey. In a recent near-infrared study of nearby SNe, \\citet{2010AJ....139..120F} found $R_V \\approx 1-2$ for their whole sample, but obtained $R_V \\approx 3.2$ if they exclude their most reddened objects. Alternatively, in a study of 80 nearby SNe Ia, \\citet{2008A&A...487...19N} found a value of $R_V=1.75\\pm0.27$ for their whole sample and a lower value of $R_V\\sim1$ if they restrict the sample to low reddening values. These differences could suggest that the effects of dust may also be dependent on the particular line of sight \\citep{2007ApJ...666..694W} or on the inclination of the host galaxies \\citep{2010arXiv1001.1744M}. Secondly, the details of the supernova progenitor system could systematically vary between the different galaxy types. Our present theoretical understanding of SNe Ia suggests they are the thermonuclear explosion of a carbon-oxygen white dwarf which has reached the Chandrasekhar limit \\citep{1973ApJ...186.1007W, 2000ARA&A..38..191H}. The mechanism for how the progenitor system accretes mass could be different between galaxy types, either accretion from a nearby companion star (which could have different metallicities depending on the stellar populations in the host galaxy types) or the merger with another white dwarf \\citep{2009ApJ...699.2026R}. Therefore, there are clear reasons to search for correlations between the properties of Ia's and the properties of their host galaxies. For example, there is a well-established difference between the rates of Ia's in passive and star-forming galaxies, potentially indicating two different paths or timescales for Ia's \\citep{1979AJ.....84..985O, 1990PASP..102.1318V, 2005A&A...433..807M, 2006ApJ...648..868S}. There have also been indications that the host galaxy type correlates with the observed residuals on the SN Hubble diagram, even after standardizing each SN Ia \\citep{2003MNRAS.340.1057S, 2008ApJ...685..752G, 2009arXiv0912.0929K}. For example, \\citet{sullivan2010} recently reported that SNe Ia in massive host galaxies are 0.08 magnitudes brighter than those in lower mass hosts after correction for the light curve shape and color (at a statistical significance of 4$\\sigma$). Such correlations would have important consequences for supernova surveys and could improve the use of SNe Ia as ``standard candles\" \\citep{2009ApJ...699L.139W, sullivan2010}. We investigate the environmental dependencies of SNe Ia by studying the residuals on the Hubble diagram (around the best fit cosmology) as a function of host galaxy type. We further ask if there are differences in the assumed dust law depending on the type of the host galaxy. In Section 2, we outline the data used in this analysis, which is taken from the full SDSS-II Supernova Survey \\citep{2008AJ....135..338F}. This sample of SNe Ia has several advantages for such environmental studies including high survey efficiency, multi-color ($ugriz$) photometry for all host galaxies and a significant cosmological volume, thus providing a fair sampling of the galaxy distribution. Also, the overall SN rate, as a function of galaxy type, has been measured using this data \\citep{smithm} In Section 2, and Appendix A, we outline the details of our analysis using two, well-known, public light curve fitting procedures; SALT2 \\citep{2007A&A...466...11G} and MLCS2k2 \\citep{2007ApJ...659..122J, 2009ApJS..185...32K}. We also describe our methodology for defining passive and star-forming host galaxies. In Section 3, we present our main results, while in Section 4, we discuss these results in light of other work in the field. We conclude in Section 5. ", "conclusions": "We present an analysis of the host galaxy dependencies for the SDSS-II Supernova Survey. We have used 361 SNe Ia (see Table \\ref{tbl_sample}) taken from the full three years of this survey, and then applied several data cuts to ensure we have a clean, well-understood, sample of low redshift SNe ($z<0.21$). We have analysed these data using two well-known light--curve fitting routines (SALT2 and MLCS2k2) to demonstrate that our results are not dependent on the details of the light curve analysis. We summarise below the main conclusions of this work: \\begin{itemize} \\item We confirm, to high significance, the strong correlation between host galaxy type and the observed width of the light curve, i.e., quick decline--rate SNe (small $x_1$ values in SALT2), favor passive host galaxies, while bright, slower decline SNe Ia (larger $x_1$ values) favor star-forming galaxies. This has been seen before by several authors. However, we find no correlation between the color of individual SNe Ia and their host galaxy, as illustrated in Figure 2. \\item We find that SNe Ia are $\\simeq0.1$ magnitudes brighter in passive host galaxies after light curve fitting. This effect is true for both SALT2 and MCLS2k2 analyses. The statistical significance of this difference is between $2$ and $3\\sigma$ dependent upon the details of the fitting methodology and the inclusion of outliers in the color and $x_1$ distributions of these data. \\item We find evidence for differences in the SN color relationship between passive and star--forming host galaxies. For SALT2, we detect differences in $\\beta$, with passive hosts showing $\\beta\\simeq2.5$ and star--forming hosts prefering $\\beta>3$. For MLCS2k2, we see a similar trend for passive hosts preferring a dust law with $R_V\\simeq1$ and star-forming hosts giving $R_V\\sim2$. The significance of these trends depends on the color range considered, but is greater than $3\\sigma$ for the full SN sample considered herein. \\item We find that the required intrinsic dispersion for passive galaxy hosts is smaller than that needed for the whole SN sample (and for star--forming hosts), e.g., only $\\sigma_{int}=0.13$ mags is required to obtain a reduced $\\chi^2$ close to unity for passive hosts compared to 0.17 mags for the star-forming sample. This lower intrinsic dispersion for passive hosts is true for both SALT2 and MLCS2k2 light curve fitters. \\item We demonstrate that the dependence on host galaxy type can be parameterized using the stellar mass of the host galaxies. We show that a 4--parameter fit to the distance modulus of SNe Ia ($M$, $\\beta$, $\\alpha$, $\\gamma$) -- where $\\gamma$ scales with stellar mass -- is better than the usual 3--parameter model given in Eqn. 1. For the data in Figure \\ref{fig9}, we find $\\gamma=0.069\\pm0.014$, or a $4\\sigma$ detection of this parameter. \\end{itemize} These conclusions are in good agreement with other work, especially Kelly et al. (2009) and Sullivan et al. (2010). In particular, Sullivan et al. (2010) see the same trends in both $M$ and $\\beta$ discussed herein with a similar level of statistical significance. This indicates that these trends are common to several SN surveys and appear to not change significantly with redshift. One possible cause for these correlations is a difference in the host galaxy metallicity (see Gallagher et al. 2008), which is correlated with the host stellar mass and host star-formation activity, and could affect the metallicity of the progenitor star thus leading to changes in the peak brightness of SNe Ia. However, the origin of these correlations requires further study, especially to ensure deficiencies in the light curve fitting techniques are not directly responsible. The host galaxy dependencies presented in this paper will be important for future supernova cosmology surveys, which may wish to exploit these dependencies to minimize the scatter on the SN Hubble diagram. This could be achieved by including further parameters in the light curve fitting or distance modulus calculation (e.g., Eqn. 2). We will explore these issues in the future with the full SDSS-II SN Survey similar to the recent analysis of Sullivan et al. (2010) for the SNLS." }, "1005/1005.3919_arXiv.txt": { "abstract": "{Two main scenarios for the formation of the Galactic bulge are invoked, the first one through gravitational collapse or hierarchical merging of subclumps, the second through secular evolution of the Galactic disc.} {We aim to constrain the formation of the Galactic bulge through studies of the correlation between kinematics and metallicities in Baade's Window ($l=1\\degr$, $b=-4\\degr$) and two other fields along the bulge minor axis ($l=0\\degr$, $b=-6\\degr$ and $b=-12\\degr$).} {We combine the radial velocity and the [Fe/H] measurements obtained with FLAMES/GIRAFFE at the VLT with a spectral resolution of R=20000, plus for the Baade's Window field the OGLE-II proper motions, and compare these with published N-body simulations of the Galactic bulge.} {We confirm the presence of two distinct populations in Baade's Window found in Hill et al. 2010: the metal-rich population presents bar-like kinematics while the metal-poor population shows kinematics corresponding to an old spheroid or a thick disc one. In this context the metallicity gradient along the bulge minor axis observed by Zoccali et al. (2008), visible also in the kinematics, can be related to a varying mix of these two populations as one moves away from the Galactic plane, alleviating the apparent contradiction between the kinematic evidence of a bar and the existence of a metallicity gradient.} {We show evidences that the two main scenarios for the bulge formation co-exist within the Milky Way bulge.} ", "introduction": "Although the Milky Way bulge is our closest opportunity to study in detail such a complex chemo-dynamical system, its formation and evolution is still poorly understood. Indeed the high extinction, the crowding, and the superposition of multiple structures along the line of sight make studies of the inner Galactic regions challenging. Two main scenarios have been invoked for bulge formation. The first one is gravitational collapse \\citep{Eggen62} or hierarchical merging of subclumps (\\cite{Noguchi99}, \\cite{Aguerri01}). In this case the bulge formed before the disc and the star-formation time-scale was very short. The resulting stars are old ($>$ 10 Gyr) and have enhancements of $\\alpha$ elements relative to iron which are characteristics of classical bulges. The other scenario is secular evolution of the disc through a bar forming a pseudo-bulge (see, e.g. \\cite{Combes90}, \\cite{Norman96}, \\cite{Kormendy04}, \\cite{Athanassoula05}). After the bar formation it heats in the vertical direction (\\cite{Combes81}, \\cite{Merrifield96}) giving rise to the typical boxy/peanut aspect. Observational data of individual stars in our Galaxy provides evidence of both scenarios. The presence of a bar in the inner Galaxy has been first suggested by \\cite{Vaucouleurs64} from gas kinematics and confirmed since then by numerous studies including infrared luminosity distribution (e.g. \\cite{BlitzSpergel91}), photometric data (e.g. \\cite{Stanek94}, \\cite{Babusiaux05}, \\cite{Benjamin05}), OH/IR and SiO-maser kinematics (e.g. \\cite{Habing06}), stellar kinematics (e.g. \\cite{Rangwala09a}) and microlensing surveys (e.g. \\cite{Alcock00}). Several triaxial structures may even coexist within the Galactic inner regions (\\cite{Alard01}, \\cite{Nishiyama05}, \\cite{LopezC07}). The boxy aspect of the bulge, detected in near-infrared light profile \\citep{Dwek95}, also argues for a secular formation of the bulge. On the other hand, from medium and high resolution spectroscopic data enhancements of $\\alpha$ elements have been observed (\\cite{McWilliam94}, \\cite{Zoccali06}, \\cite{Fulbright07}, \\cite{Lecureur07}) suggesting a short formation time-scale. The discovery of a possible dwarf remnant in the bulge by \\cite{Ferraro09} also argues for a hierarchical merging scenario. A metallicity gradient has been observed in the minor-axis direction of the bulge \\citep{PaperI} favouring a bulge formation through dissipational collapse. Recently some observed similarities between the [$\\alpha$/Fe] bulge trend and the trend of the local thick stars (\\cite{Melendez08}, \\cite{Ryde09}, \\cite{AlvesBrito10}) suggest that the bulge and the local thick disc experienced similar chemical evolution histories. In what concerns the age, colour-magnitude diagrams show that most of the Galactic bulge stars are older than 10 Gyr (\\cite{Ortolani95}, \\cite{Feltzing00}, \\cite{Zoccali03}, \\cite{Clarkson08}) although an intermediate-age population exists (\\cite{vanLoon03}, \\cite{GroenewegenBlommaert05}).\\par Chemo-dynamical modelling of disc galaxy formation in a Cold Dark Matter (CDM) universe stresses the fact that both types of bulges can coexist in the same galaxy. In the \\cite{Samland03} simulation the bulge contains two stellar populations, an old population formed during the collapse phase and a younger bar population, differing in the [$\\alpha$/Fe] ratio. \\cite{Nakasato03} suggest that the Galactic bulge may consist of two chemically different components: one rapidly formed through subgalactic clump merger in the proto-Galaxy, and the other one formed later in the inner region of the disc. The [Fe/H] abundance of the merger component tends to be smaller than that of the second component. Recently \\cite{Rahimi09} simulated bulges formed through multiple mergers and analysed the chemical and dynamical properties of accreted stars with respect to locally formed stars: accreted stars tend to form in early epochs and have lower [Fe/H] and higher [Mg/Fe] ratios, as expected.\\par Exploring correlations of abundances and kinematics in Baade's Window, \\cite{Soto07} suggest a transition in the kinematics of the bulge, from an isotropic oblate spheroid to a bar, at [Fe/H] $= -0.5$ dex. Their study was based on 315 K and M giants with proper motions of \\cite{Spaenhauer92}, radial velocities of \\cite{Sadler96} and low-resolution abundances of \\cite{Terndrup95} re-calibrated with the iron abundances scale of \\cite{Fulbright06}. Similar conclusions were obtained earlier by \\cite{Zhao94} using a sample of 62 K giants who pointed out the triaxiality of the bulge from kinematics in Baade's Window and the fact that the metal poor and metal rich populations were not drawn from the same distribution. \\par In the present paper we analyse the correlations between kinematics and metallicity along three different minor-axis fields: Baade's Window ($b=-4\\degr$), $b=-6\\degr$ and $b=-12\\degr$, for which [Fe/H] abundances (\\cite{PaperI} hereafter Paper\\,I and \\cite{PaperII} hereafter Paper\\,II) and radial velocities data for about 700 stars have been determined. Paper\\,II obtained [Fe/H] and [Mg/H] metallicity distributions for red clump stars in Baade's Window and showed that the sample seems to be separated into two different populations in the metallicity distributions. We here use the kinematic properties of the samples to confirm and constrain the nature of these two populations in Baade's Window and study their relative proportion change along the bulge minor axis fields of Paper\\,I. The paper is organized as follows: section 2 summarizes the data used in this study. In section 3 we analyse Baade's Window by combining our spectroscopic data with OGLE proper motion data. In section 4 we analyse the radial velocity versus metallicity trend along the bulge minor axis. In section 5 we compare our data with some published N-body models. Section 6 discusses our results. ", "conclusions": "\\begin{center} \\begin{tabular}{lrrrrcccc} \\hline\\hline Field & $l$ & $b$ & $z_{GC}$ & $N$ & $\\langle V_r \\rangle$ & $\\sigma_{r}$ & skew & kurtosis\\\\ & (\\degr) & (\\degr) & (pc) & & (km/s) & (km/s) \\\\ \\hline $b = -4\\degr$ & 1.1 & $-4.0$ & 606 & 194 & { }~{ }~5 $\\pm$ 7 & 104 $\\pm$ 5 & $-0.07 \\pm 0.1$ & $-0.54 \\pm 0.2$ \\\\ $b = -6\\degr$ & 0.2 & $-6.0$ & 844 & 201 & $-10 \\pm 6$ & { }~83 $\\pm$ 4 & { }~0.01 $\\pm$ 0.1 & $-0.19 \\pm 0.2$ \\\\ $b = -12\\degr$ & 0.0 & $-12.0$ & 1700 & 99 & $-14 \\pm 8$ & { }~80 $\\pm$ 8 & { }~0.38 $\\pm$ 0.4 & { }~1.71 $\\pm$ 0.7 \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} We first compare the kinematic behaviour along the bulge minor axis (Table \\ref{tab:fieldsSummary}) with previous results from the literature. We saw in section 3 that our Baade's Window radial velocity dispersion is in excellent agreement with previous measurements. At $b=-6\\degr$ our radial velocity dispersion is lower than the BRAVA measure ($\\sigma_{r}=108 \\pm 7$ km/s, \\cite{Howard08}). The decrease of the radial velocity dispersion with galactic latitude we observe is consistent with the SiO maser measurements of \\cite{Izumiura95paperIII}. All the fields show a negligible skew, while the kurtosis are different. This variation of the kurtosis was not detected in \\cite{Howard08}. At $b=-12\\degr$ the distribution is significantly pointy (with 99.3\\% confidence according to the Anscombe-Glynn kurtosis test), indicating that the kinematics are significantly affected by the disc. At $b=-6\\degr$ the kurtosis becomes consistent with zero. In Baade's Window the distribution is significantly flattened (with 93\\% confidence in the RGB sample and 99.8\\% in the red-clump sample). \\cite{Sharples90} and \\cite{Rangwala09a} also measured a skew consistent with zero and kurtosis significantly negative in Baade's Window, indicating that the distributions are flat-topped rather than peaked. \\cite{Rangwala09a} concludes that this seems to be consistent with a model of the bar with stars in elongated orbits forming two streams at different mean radial velocities, broadening and flattening the total distribution. \\begin{figure*} \\begin{minipage}[b]{0.5\\linewidth} \\centering \\includegraphics[width=9cm]{methist.eps} \\caption{Distribution of the metallicity for the different galactic latitudes of Paper\\,I} \\label{fig:metdens} \\end{minipage} \\hspace{0.5cm} \\begin{minipage}[b]{0.5\\linewidth} \\centering \\includegraphics[width=9cm]{metdispVr.eps} \\caption{Dispersion of the radial velocity for the different galactic latitudes as a function of metallicity by bins of 0.4 dex. } \\label{fig:metdispVr} \\end{minipage} \\end{figure*} We analyse now kinematic data versus metallicity. The mean radial velocities do not show significant variation with metallicity in any of our fields. Figure \\ref{fig:metdispVr} shows that the velocity dispersion at the rich end decreases significantly with latitude. However the velocity dispersion on the metal-poor end does not vary significantly with latitude. Summing all stars with [Fe/H] $< -0.5$ dex in all fields leads to $\\sigma_{r}=97 \\pm 7$ km/s. The field $b=-6\\degr$ is close enough to Plaut's Window ($l=0\\degr$, $b=-8\\degr$) so that we can compare our results to the proper motion study of \\cite{Vieira07}. The Besan\\c{c}on model indicates a mean distance of our selected bulge stars to be 6.5 kpc, compatible with the \\cite{Vieira07} selection. Their metallicity distribution is compatible with ours. They do not find neither any change in the proper-motion dispersion as a function of metallicity. They found $\\sigma_l=102\\pm3$ km/s and $\\sigma_b=88\\pm3$ km/s at $b=-8\\degr$, we obtain $\\sigma_r=83$ km/s at $b=-6\\degr$, leading to a full picture coherent with the predictions of the model of \\cite{Zhao96} of $\\sigma_b \\sim \\sigma_r$ and $\\sigma_l > \\sigma_b$ at these latitudes (see the bottom of Fig. 6 of \\cite{Zhao96}). We have seen in the previous section that Baade's Window kinematic behaviour as a function of metallicity can be interpreted as a mix of two populations, a metal poor component with kinematics that can be associated to an old spheroid population and a metal rich component with bar-driven kinematics. Under this light we can interpret the variation with galactic latitude of both the metallicity distribution function (Fig. \\ref{fig:metdens}) and the kinematic as a function of metallicity (Fig. \\ref{fig:metdispVr}) as the bar population disappearing while moving away from the plane. At high latitudes the foreground disc component dominates the metal rich part of the kinematic behaviour. The metal poor component associated with the old spheroid stays present along the bulge minor axis. We have obtained an estimation of the variation of the populations with latitude by using the SEMMUL Gaussian components decomposition in the metallicity distribution of the Paper\\,I samples. Table \\ref{tab:b4decomp} gives the decomposition of the Baade's Window RGB sample. The proportion of the metal-rich component is higher than in the red clump sample of Paper\\,II and its spread in metallicity is higher also, most probably due to a difference in the sample selection which is biased towards stars closer to the Sun in the RGB sample. Although both samples are not exactly on the same scale at the high metallicity end (cf section 5.1.4 of Paper II), the mean metallicities for the metal-poor population is extremely similar in both samples (see also the decomposition of the full Baade's Window sample on the same metallicity scale at the end of section 3.1). At $b=-6\\degr$ the Wilks' test allows to keep a solution with 3 components presented in Table \\ref{tab:b6decomp} rather than the 2 components one. Population A and population B could correspond to the population A and B observed in Baade's Window. The mean metallicities are coherent although their spread is smaller. The radial velocity dispersions of population A are identical. The radial velocity dispersion of population B decreases at $b=-6\\degr$ as expected for a bar-like kinematic behaviour (see Fig. \\ref{fig:compFuxSigVr} and associated text). Population C represents only $6\\pm2$\\% of the sample with a low mean metallicity of $-1.1\\pm0.1$ dex and a high velocity dispersion of 127$\\pm$26 km/s, which could therefore be associated to the halo. The Besan\\c{c}on model prediction of only $0.4\\pm0.1$\\% of halo star in the sample could therefore have been underestimated. This population could also have been hidden in the population A at $b=-4\\degr$. Selecting all stars with [Fe/H]$<-0.9$ in our 3 fields we obtain 16 stars with a radial velocity dispersion of $\\sigma_r= 116\\pm21$km/s, which is coherent with the solar neighbourhood velocity dispersions measured in this metallicity range (e.g. \\cite{ChibaBeers00}: $\\sigma_U \\sim 110 \\pm 10$ km/s) containing both thick disc and halo stars. We are not in a position to clearly associate this population either to the halo or to a metal-poor thick disc in our inner galactic samples. SEMMUL did not converge on the $b=-12\\degr$ field due to the smaller number of stars and the higher contamination with thin, thick discs and halo stars expected in this field. At $b=-12\\degr$ the metal rich component present at $b=-4\\degr$ and $b=-6\\degr$ seems to have fully disappeared. The metal rich velocity part of the velocity dispersion corresponds to a disc like component (see next section) while the metal poor part shows a velocity dispersion still coherent with the metal poor population of $b=-4\\degr$ and $b=-6\\degr$, although we cannot distinguish a spheroid and a thick disc contribution. \\begin{table*} \\caption{SEMMUL Gaussian components decomposition of field $b=-4\\degr$, RGB sample only. The radial velocity dispersion $\\sigma_{r}$ is given in km/s.} \\label{tab:b4decomp} \\begin{center} \\begin{tabular}{lllll} \\hline\\hline Pop & [Fe/H] & $\\sigma_{[Fe/H]}$ & \\% & $\\sigma_{r}$ \\\\ \\hline A & $-0.31 \\pm 0.05$ & $0.39 \\pm 0.02$ & $33 \\pm 3$ & { }~$88 \\pm 8$ \\\\ B & { }~$0.13 \\pm 0.02$ & $0.23 \\pm 0.01$ & $67 \\pm 3$ & $109 \\pm 7$ \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} \\begin{table*} \\caption{SEMMUL Gaussian components decomposition of field $b=-6\\degr$. The radial velocity dispersion $\\sigma_{r}$ is given in km/s. } \\label{tab:b6decomp} \\begin{center} \\begin{tabular}{lllll} \\hline\\hline Pop & [Fe/H] & $\\sigma_{[Fe/H]}$ & \\% & $\\sigma_{r}$ \\\\ \\hline C & $-1.09 \\pm 0.07$ & 0.24 $\\pm$ 0.01 & { }~6 $\\pm$ 2 & 127 $\\pm$ 26 \\\\ A & $-0.27 \\pm 0.02$ & 0.24 $\\pm$ 0.01 & 64 $\\pm$ 3 & { }~83 $\\pm$ 5 \\\\ B & { }~0.14 $\\pm$ 0.02 & 0.13 $\\pm$ 0.01 & 30 $\\pm$ 3 & { }~70 $\\pm$ 6\\\\ \\hline \\end{tabular} \\end{center} \\end{table*} \\begin{figure*} \\vspace{0.3cm} \\begin{minipage}[b]{0.49\\linewidth} \\centering \\includegraphics[width=8cm]{compSigVr.eps} \\caption{Radial velocity dispersion along the bulge minor axis compared to the BRAVA data (\\cite{Howard08} and \\cite{Howard09}) and the model of \\cite{Zhao96}} \\label{fig:compSigVr} \\end{minipage} \\hspace{0.2cm} \\begin{minipage}[b]{0.49\\linewidth} \\centering \\includegraphics[width=8cm]{compFuxSigVr.eps} \\caption{Radial velocity dispersion along the bulge minor axis compared with the model of \\cite{Fux99}. The metal poor and metal rich population are defined by the 33\\% and 66\\% quantile of the [Fe/H] distribution.} \\label{fig:compFuxSigVr} \\end{minipage} \\end{figure*} Our analysis of the kinematics as a function of metallicity in Baade's Window shows that our sample can be decomposed in two distinct populations for which we suggest different formation scenarii. The metal poor component does not show any correlation between its velocity components and is therefore consistent with an isotropic rotating population. Paper\\,II showed that this component is enriched in [Mg/Fe]. We interpret this population as an old spheroid with a rapid time-scale formation. The metal rich component shows a vertex deviation consistent with that expected from tags a population with orbits supporting a bar. Paper\\,II showed that this component has a [Mg/Fe] near solar. We interpret this population as a pseudo-bulge formed over a long time scale through disc secular evolution under the action of a bar. This pseudo-bulge is gradually disappearing when moving away from the Galactic plane. In this context we can give a new consistent interpretation of the metallicity gradient in the bulge. A metallicity gradient is indeed visible in the bulge when observing further away from the plane than Baade's Window (\\cite{Frogel99} and Paper\\,I), while in the inner regions ($\\vert b \\vert\\leq4\\degr$) no gradient in metallicity has been found (\\cite{Ramirez00} and \\cite{Rich07}). This can be understood if the bar as well as the old spheroid population are both present in the inner regions, leading to a constant metallicity at $\\vert b \\vert\\leq4\\degr$, while the bar influence gradually fades further away from the plane than Baade's Window. At high latitudes only the old spheroid remains: the mean metallicity of the outer bulge measured by \\cite{IbataGilmore95} of [Fe/H]$\\sim-0.3$ dex corresponds very well to our metal poor population. This scenario is also consistent with the distribution of bulge globular clusters of \\cite{Valenti09} who found no evidence for a metallicity gradient but all their clusters with [Fe/H]$>-0.5$ dex are located within ($\\vert b \\vert\\leq5\\degr$). In what concerns the age of the two populations, we expect the spheroid component to be old while the pseudo-bulge component may contain both the old stars of the inner disc redistributed by the bar and younger stars whose formation has been triggered by the bar gas flow. \\cite{vanLoon03} found that although the bulk of the bulge population is old, a fraction of the stars are of intermediate age (1 to 7 Gyr). \\cite{GroenewegenBlommaert05} observed Mira stars of ages 1-3 Gyr at all latitudes from $-1.2$ to $-5.8$ in the OGLE-II data. \\cite{Uttenthaler07} found 4 bulge stars with ages lower than 3 Gyr at a latitude of $b=-10\\degr$. \\cite{Bensby09} found 3 microlensed bulge dwarfs with ages lower than 5 Gyr. 87\\% of the variable stars detected by \\cite{KouzumaYamaoka09} are distributed within $\\vert b \\vert\\leq5\\degr$ and most of them should be large-amplitude and long-period variables such as Mira variables or OH/IR stars. This intermediate age population has been shown to trace the Galactic bar (\\cite{vanLoon03}, \\cite{Izumiura95paperIII}, \\cite{GroenewegenBlommaert05}, \\cite{KouzumaYamaoka09}), although providing a larger bar angle ($\\sim40\\degr$) than studies based on older tracers such as red clump stars ($\\sim20\\degr$, \\cite{Stanek94}, \\cite{Babusiaux05}). This discrepancy in the bar angle could well be explained if the old tracers probed a mix of spheroid and bar structures while the young tracers only probe the bar one (although biases on the longitude area surveyed needs also to be taken into account, see \\cite{Nishiyama05}). If this intermediate age population was associated to a part of the bar component, their presence in the CMDs would decrease while going away from the plane as the main bar component and would therefore be a small fraction of the CMD of \\cite{Zoccali03} at $b=-6\\degr$. \\cite{Clarkson08} obtained a proper motion decontaminated CMD with a well defined old turn-off in an inner field ($l=1\\degr$, $b=-3\\degr$). However we would expect an intermediate age population associated with the bar to be metal rich, which, due to the age-metallicity degeneracy, would imply that this population could be hidden in the CMD of \\cite{Clarkson08} if its contribution is small enough compared to the bulk of the bulge population. The new filter combination proposed by the ACS Bulge Treasury Programme to break the age-metallicity-temperature degeneracy \\citep{Brown09} should provide new lights on this issue. We note that in Baade's Window, neither the kinematics nor the chemistry allows to distinguish what we call the old spheroid to the thick disc. The mean metallicity of the solar neighbourhood thick disc is however lower (e.g. \\cite{Fuhrmann08} derived [Fe/H] $=-0.6$ and [Mg/H] $=-0.2$) than the mean metallicity of our metal poor population ([Fe/H]$= -0.27$ dex and [Mg/H]$=-0.04$, Paper\\,II). \\cite{Melendez08}, \\cite{Ryde09}, \\cite{Bensby09} and \\cite{AlvesBrito10} observed similarities between the metallicity of the bulge and the metallicity of thick disc stars for metal poor stars. The sample of \\cite{Ryde09} contains 11 of our stars all with less than solar metallicity. Simulations of the formation of thick stellar discs by rapid internal evolution in unstable, gas-rich, clumpy discs \\citep{Bournaud09} show that thick discs and classical bulges form together in a time-scale shorter than 1 Gyr which allow to explain the observed abundance similarities. The coexistence of classical and pseudo bulge has been observed in external galaxies (\\cite{Prugniel01}, \\cite{Peletier07}, \\cite{Erwin08}) and obtained by N-body simulations (\\cite{Samland03}, \\cite{Athanassoula05}). The chemical and dynamical model of \\cite{Nakasato03} suggests the presence in the bulge of two chemically different components as we found, one formed quickly through the subgalactic clump merger in the proto-Galaxy, and the other has formed gradually in the inner disc. But they do not have a bar in their model. They fitted well the kinematics of \\cite{Minniti96} who observed at $l=8\\degr$, $b=7\\degr$ a decrease of $\\sigma_r$ with metallicity, which is coherent with what we observed at $b=-12\\degr$ and the fact that \\cite{Nakasato03} defined the bulge radius as R$<$2kpc. The chemo-dynamical model of \\cite{Samland03} predicts the different characteristics of our sample: their total bulge population contains two stellar populations: a metal rich population ([Fe/H]$>$0.17) with [$\\alpha$/Fe]$<$0 associated with the bar, and an old population that formed during the proto-galactic collapse, with a high [$\\alpha$/Fe] and a [Fe/H] corresponding to the ``thick disc'' component. Their model also predicts the resulting apparent metallicity gradient along the bulge minor axis (their Fig. 13). Our study highlights the importance to combine metallicity to 3D-kinematic information to disentangle the different bulge populations. This approach needs to be extended to various galactic longitudes. Gaia will not only provide those but also allow to determine the distances (probing the different structures along the line of sight and removing distance induced biases) and work on an impressively large sample of un-contaminated bulge stars." }, "1005/1005.3416_arXiv.txt": { "abstract": "The main objective of this study is to better understand how magnetic helicity injection in an active region (AR) is related to the occurrence and intensity of solar flares. We therefore investigate magnetic helicity injection rate and unsigned magnetic flux, as a reference. In total, 378 ARs are analyzed using $SOHO$/MDI magnetograms. The 24 hr averaged helicity injection rate and unsigned magnetic flux are compared with the flare index and the flare-productive probability in next 24 hr following a measurement. In addition, we study the variation of helicity over a span of several days around the times of the 19 flares above M5.0 which occurred in selected strong flare-productive ARs. The major findings of this study are as follows: (1) for a sub-sample of 91 large ARs with unsigned magnetic fluxes in the range from 3 to 5$\\times$10$^{22}$ Mx, there is a difference in magnetic helicity injection rate between flaring ARs and non-flaring ARs by a factor of 2; (2) the $GOES$ C-flare-productive probability as a function of helicity injection displays a sharp boundary between flare-productive ARs and flare-quiet ones; (3) the history of helicity injection before all the 19 major flares displayed a common characteristic: a significant helicity accumulation of (3--45)$\\times$10$^{42}$ Mx$^2$ during a phase of monotonically increasing helicity over 0.5--2 days. Our results support the notion that helicity injection is important in flares, but it is not effective to use it alone for the purpose of flare forecast. It is necessary to find a way to better characterize the time history of helicity injection as well as its spatial distribution inside ARs. ", "introduction": "A solar flare is a sudden, rapid, and intense brightening usually in a magnetically complicated solar active region (AR). Solar flares produce high energy particles, radiation, and erupting magnetic structures that are related to geomagnetic storms. Their strong electromagnetic radiations from radio waves to gamma-rays have direct effect on cell phones and the global positioning system and heat up the terrestrial atmosphere within minutes so that satellites drop into lower orbits (Schwenn 2006). Enormous economic and commercial losses can be caused by these effects (Baker 2004). Therefore there have been significant efforts to develop flare forecasting systems as part of the space weather service. It is generally thought that flare-productive ARs exhibit complex and non-potential magnetic structures related to the stored magnetic energy to power flares. For this reason, many studies of relationship between the solar flare and photospheric magnetic field properties have been carried out since the flare was first observed and recorded by Carrington (1859) and Hodgson (1859). Some examples include the unbalanced changes in the photospheric line-of-sight magnetic field (Cameron \\& Sammis 1999; Spirock et al. 2002; Wang et al. 2002); rapid changes of the sunspot structure associated with a substantial fraction of flares (Liu et al. 2005; Deng et al. 2005; Wang et al. 2004a, 2005; Chen et al. 2007); the magnetic shear angle evolution (Hagyard et al. 1984; Hagyard \\& Rabin 1986; Sivaraman et al. 1992; Schmieder et al. 1994; Wang et al. 1994, 2004a); the horizontal gradient of longitudinal magnetic fields (Zirin \\& Wang 1993; Zhang et al. 1994; Tian et al. 2002); electric current (Canfield et al. 1993; Lin et al. 1993); magnetic helicity injection (Moon et al. 2002a, 2000b; Sakurai \\& Hagino 2003; Yokoyama et al. 2003; Park et al. 2008). Based on the above mentioned studies, current flare forecasting models are moving toward multiple-magnetic parameter-based approaches (Leka \\& Barnes 2003a, 2003b; Li et al 2008) from sunspot-morphological evolution-based approaches (McIntosh 1990; Gallagher et al. 2002). Magnetic helicity is a measure of twists, kinks, and inter-linkages of magnetic field lines (Berger \\& Field 1984; Pevtsov 2008) and is a useful and important parameter to indicate topology and non-potentiality of a magnetic field system. Naturally magnetic helicity studies have been conducted to the energy buildup and instability leading to solar eruptions and coronal mass ejections (e.g., Rust 2001; Kusano et al. 2004; Phillips et al. 2005; Fan 2009). There were a number of studies related to a rapid magnetic helicity change as an impending condition or a trigger for solar flares (e.g. Moon et al. 2002a, 2002b). LaBonte et al. (2007) surveyed magnetic helicity injection in 48 X-flaring ARs and 345 non-X-class flaring regions, and found that a necessary condition for the occurrence of an X-class flare is that the peak helicity flux has a magnitude $\\textgreater$ 6$\\times$10$^{36}$ Mx$^2$ s$^{-1}$. Park et al. (2008) found that a substantial amount of helicity is accumulated before the flare in all the 11 X-class flare events, and suggested a warning sign of flares can be given by the presence of a phase of monotonically increasing helicity. Motivated by these results we explore the feasibility of using magnetic helicity to build a flare forecasting system utilizing a data sample covering almost one solar cycle. ", "conclusions": "We have investigated the time variations of $\\dot{H}$ and $\\Phi$ in 378 solar ARs and compared the two average parameters, $|$$<$$\\dot{H}$$>$$|$ and $<$$\\Phi$$>$, with $F_{idx}$. Although there is a large amount of scatter in the data samples, we found a moderate correlation between the parameters and $F_{idx}$. The larger $F_{idx}$ an AR has, the larger values of $\\dot{H}$ and $\\Phi$ it presents. To improve the correlation, we have defined a new parameter as an equally weighted linear combination of the two rescaled parameters (0.5 of each). The logarithmic-scale CC of $F_{idx}$ versus the new parameter increased slightly to 0.47. It is not surprising because $|$$<$$\\dot{H}$$>$$|$ is well correlated with $<$$\\Phi$$>$ as shown in Figure 3. Moreover, by considering 48 and 72 hr profiles of $\\dot{H}$ and $\\Phi$ for calculation of the two average parameters, we have executed the same correlation study between the two parameters and the next-day flare index. We found that the longer the period we use for average, the worse the correlation will be, especially in the case of $|$$<$$\\dot{H}$$>$$|$ (CC=0.38 and 0.21 for 48 and 76 hr periods, respectively). This might be because we do not consider the flaring history before or during the measurement period of the parameters, but we compare the parameters with $F_{idx}$ calculated only for the following day of the measurement period. We understand that no matter which method is used, the correlation between the parameters and flare index is not high. This is an intrinsic problem for flare forecasting as the occurrence of a flare depends not only on the amount of magnetic energy built up in an AR, but also on how it is triggered. For example, if new flux-rope emergence is the driver of flares (Schrijver 2009), or if a flare is exactly a result of a small and localized (quite possibly unobservable) perturbation affecting the whole system like self-organized criticality dynamics (Bak et al. 1987; B\\'{e}langer et al. 2007), then it is not feasible to carry out prediction of flare onset time and magnitude by using present-day parameters derived from photospheric magnetic field observations. More specifically for our case, helicity accumulation might be a necessary, but not sufficient condition for flares. Perhaps a triggering mechanism is necessary even a magnetic system has enough non-potentiality to power a flare (so-called metastable state). This idea agrees with the study that a number of X-class flares occurred during the phase of almost constant helicity after the phase of 2--3 days of monotonically increasing helicity (Park et al. 2008). Interestingly, contrary to the expectation that magnetic helicity injection is more closely related to flare productivity than to magnetic flux, our result shows that the correlation between $|$$<$$\\dot{H}$$>$$|$ and $F_{idx}$ is not stronger than that between $<$$\\Phi$$>$ and $F_{idx}$. The logarithmic-scale CCs of $F_{idx}$ versus $|$$<$$\\dot{H}$$>$$|$ and $<$$\\Phi$$>$ are 0.42 and 0.43, respectively. If only the flaring groups with non-zero flare index are considered, then $|$$<$$\\dot{H}$$>$$|$ is not better than $<$$\\Phi$$>$ in predicting how strong the flares will be. This might be due to the fact that we simply use the 1 day average of $\\dot{H}$ in the entire AR for comparison with $F_{idx}$ without more specifically characterizing the temporal and spatial evolution of helicity in the AR related to a flaring condition. Magnetic helicity, however, is useful in predicting whether an AR will produce flares or not. Note that predicting the occurrence of flares is different from predicting the strength of flares. By examining more careful studies such as the helicity injection difference between flare-productive and flare-quiet ARs, the flare-productive probability as a function of $|$$<$$\\dot{H}$$>$$|$, and the temporal evolution of helicity in major flare-producing ARs, we have found that magnetic helicity injection has some interesting features related to flares as follows. \\begin{enumerate} \\item For 91 AR samples in the range (3--5)$\\times$10$^{22}$ Mx of large $<$$\\Phi$$>$ the $flaring$ group has $|$$<$$\\dot{H}$$>$$|$ about twice greater than that of the $non$-$flaring$ group. On the other hand, 118 AR samples of large $|$$<$$\\dot{H}$$>$$|$ do not show the significant difference in $<$$\\Phi$$>$ between the $flaring$ and $non$-$flaring$ groups. \\item The helicity parameter $|$$<$$\\dot{H}$$>$$|$ demonstrates a rapid increase of $P_{i}$ compared to that of $<$$\\Phi$$>$ in the rescaled range of 0--0.15 of the parameter. $P_{C}$($|$$<$$\\dot{H}$$>$$|$), especially, quickly reaches up to $\\sim$90$\\%$ from 46$\\%$ in the very low rescaled range, 0--0.15, of the parameter, and it retains a high value above 90$\\%$ in the rest, 0.15--0.6, of the statistically meaningful range. \\item Helicity of (3--45)$\\times$10$^{42}$ Mx$^2$ accumulates significantly and consistently over 0.5--2 days for all the 19 major flares under investigation. More specifically, following the significant amount of long-term helicity accumulation with fast injection rate, 4 and 8 flares occurred when helicity injection rate starts to become slow (sometimes almost zero) and reverse its sign, respectively. \\end{enumerate} Based on these results, the magnetic helicity can be used for the improvement of flare forecasting. First of all, when an AR has large $<$$\\Phi$$>$, we may better determine whether or not it will produce a flare by considering $|$$<$$\\dot{H}$$>$$|$ of the AR. Second, the helicity parameter $|$$<$$\\dot{H}$$>$$|$ would allow us to establish a better defined cutoff between C-flare-productive and C-flare-quiet ARs than $<$$\\Phi$$>$ if we take into account a sharp increase of $P_{i}$ in the very low rescaled range of the parameter. Third, an early warning sign of flare occurrence could be based on tracking of a phase of monotonically increasing helicity because there is always a significant amount of helicity accumulation a few days before major flares. We might also make an urgent warning sign when helicity injection rate becomes very slow or the opposite sign of helicity starts to be injected after the significant helicity accumulation phase. The sign reversal of the magnetic helicity may support the numerical simulation model for solar flare onset proposed by Kusano et al. (2003b) in which they showed that magnetic reconnection quickly grows in the site of the helicity annihilation with different signs. Some observations of helicity inversion, similar to our result, were also reported around the time of flare onset (Kusano et al. 2003a; Yokoyama et al. 2003; Wang et al. 2004b). For more practical and advanced flare forecasting, we may think how to consider the past history of flare occurrence in an AR under investigation and combine the helicity parameter with others with different weighting coefficients. Besides that it would be required to better characterize not only the time history of helicity injection but also its spatial distribution inside ARs. Finally, our study may lead to some physical understanding of flare on-set. For example, why do only some of the samples with the large helicity injection produce major flares, but not for all? Is a significant amount of helicity accumulation necessary or sufficient conditions for flares? We believe that the study of magnetic helicity in a coronal volume of an AR will help us to better explain physically for these questions and understand pre-flare conditions and energy storage process of flares in more detail. Our future work to calculate the coronal helicity by using three-dimensional non-linear force-free magnetic field extrapolations is therefore in progress." }, "1005/1005.4693_arXiv.txt": { "abstract": "We recently proposed that molecular cloud dense cores undergo a prolonged period of quasi-static contraction prior to true collapse. This theory could explain the observation that many starless cores exhibit, through their spectral line profiles, signs of inward motion. We now use our model, together with a publicly available radiative transfer code, to determine the emission from three commonly used species - N$_2$H$^+$, CS, and HCN. A representative dense core of $3\\,\\,\\Msun$ that has been contracting for 1~Myr has line profiles that qualitatively match the observed ones. In particular, optically thick lines have about the right degree of blue-red asymmetry, the empirical hallmark of contraction. The \\hbox{$J\\,=\\,2\\rightarrow\\,1$} rotational transition of CS only attains the correct type of profile if the species is centrally depleted, as has been suggested by previous studies. These results support the idea that a slow, but accelerating, contraction leads to protostellar collapse. In the future, the kind of analysis presented here can be used to assign ages to individual starless cores. ", "introduction": "Over the last decade, astronomers have carefully scrutinized the large fraction of molecular cloud dense cores that contain no point sources of radiation, in order to understand more fully the onset of star formation. It was already established that the gross properties of these objects differ little from their counterparts with internal stars \\citep{lm99}, except perhaps for a mass that is slightly higher on average \\citep[][Fig.~3]{jm99}. More detailed spectroscopic studies of molecular line emission revealed that many starless cores exhibit signs of inward contraction \\citep{wi99,lmt99,ge00, lmt01,sc07}. The hallmark of this so-called infall signature is an asymmetric emission line, often with a self-absorption dip, that is skewed to the blue. There is a variety of line profiles seen in dense cores, sometimes even in the same object, and the profile just described is by no means universal. It is preferentially found in starless cores with a higher central column density and optical depth \\citep[e.g.][]{lmt99,ge00}. Such cores also exhibit chemical abundances that differentiate them from lower-density objects \\citep{c05,kc08}. The kinematic signature is especially common in more evolved dense cores that contain Class~0 and Class~I sources, i.e., very young, embedded stars \\citep{me00}. These facts, together with the simple and compelling infall interpretation of the asymmetric profiles \\citep{m96}, have convinced most researchers that we are witnessing a type of contraction that precedes the free-fall collapse onto a protostar. The contraction hypothesis is especially convincing for starless cores, where there is no possibility of confusion from stellar outflows. What is the physical cause of this motion? The observed magnitude of the infall velocity is typically a few tenths of the sound speed. This finding rules out ambipolar diffusion as the underlying process, since the characteristic ion-neutral drift speed is much smaller \\citep[e.g.][]{cb01}. Conversely, the subsonic level of the velocity, together with the statistical prevalence of starless cores \\citep[e.g.][]{f09}, indicate that such objects are relatively long-lived \\citep{ki05}, and thus are unlikely to be in a true state of collapse, as proposed by \\citet{kf05}. In a recent contribution \\citep[][hereafter Paper~I]{sy09} we introduced a new model for the contraction. Starless cores, like all self-gravitating objects, undergo bulk oscillations. Indeed, there are now several cases where spectroscopic observations indicate that such oscillations are occurring \\citep{r06,b07,v08}. For an object that has evolved to the point of collapse, by whatever means, the frequency of the lowest normal mode vanishes. We studied, using perturbation theory, how a cloud subject to such a frozen (more commonly called `neutral') mode slowly evolves toward full-blown collapse. We found that this previously unrecognized phase of quasi-static contraction lasts a considerable period (of order $10^6$~yr), during which time the cloud attains internal speeds similar to those observed (see Fig.~4 of Paper~I). Encouraged by this result, we now seek to forge a stronger link to the observations. We calculate, using a publicly available radiative transfer code, the emitted line spectrum from our model cloud. We choose three molecules that are among the most commonly employed in the spectroscopic studies: N$_2$H$^+$, CS, and HCN. Following the early surveys of \\citet{tt77} and \\citet{w92}, N$_2$H$^+$ is often used as a tracer of dense, relatively quiescent cloud gas. Its rotational lines are optically thin in low-mass dense cores. Thus, they do not exhibit either self-absorption or blue-red asymmetry, but serve principally as a gauge of excitation temperature and as a comparison benchmark with lines that are sensitive to internal cloud motion. The rotational lines of CS are often optically thick, and can have asymmetric profiles that have long been interpreted kinematically \\citep{w86}. On the other hand, the infall signature is relatively weak in starless cores, in part because the polar molecule CS is depleted at high density from the gas phase onto grain mantles \\citep{bl97}. Finally, the \\hbox{$J\\,=\\,1\\rightarrow\\,0$} rotational transition of HCN has the strongest infall signature of all, and the molecule appears to suffer little depletion \\citep{a05,s07}. Although our model cloud is a highly idealized isothermal sphere, the results of this study further lend credence to the underlying dynamical picture. For a representative 3\\,\\,$\\Msun$ cloud that has been contracting for 1~Myr, we find that all the calculated line profiles resemble those observed, at least in an average sense. Interestingly, the CS profile is only acceptable once we include central depletion of the molecule. In Section~2 below, we first describe our implementation of the radiative transfer program. Section~3 then summarizes observations to date of the three molecules in starless dense cores. We describe, in Section~4, how the emergent line profiles alter when we change our physical input parameters. Judging from these trends, we then select a canonical model cloud and compare its emission spectrum to the available data. Finally, Section~5 discusses limitations of the dynamical theory and the prospects for improvement. We also indicate how the model, even in its current, simplified form, can be used to assign contraction ages to individual starless cores. ", "conclusions": "This suite of radiative transfer calculations performed on our theoretical cloud model has yielded two important findings. First, we have been able to match the generic, observed properties of the selected line profiles using a reasonable cloud mass, turbulent velocity dispersion, and evolutionary time. This success may indicate that the model has captured the essential physics of the contraction process itself. Note also that the accelerating character of the contraction in our theory explains naturally the key observation that starless cores of lower density exhibit less asymmetry in their line profiles \\citep{ge00,wi06}. Second, the sensitivity of the emergent line profiles to the input parameters means that the model offers a practical route to gauge starless core properties that have previously been difficult to ascertain. For example, given the empirical mass and kinetic temperature of an object, the matching of line profiles would yield a first estimate of its contraction age. Obtaining ages for numerous starless cores in different environments would aid greatly in understanding the onset of star formation. Of course, the observations offer more detail than the model can explain, at least in its current, highly idealized form. We cannot yet account for the rather surprising spatial extent of asymmetric CS emission \\citep{me00,lmt01}. Nor can we explain the relative intensities of the HCN hyperfine lines \\citep{s04}. Explaining both features of the data may require alterations in the spatial abundance gradients, in the manner employed in Sections 4.2 and 4.4. Additionally, an improved model would consider more carefully how the interior, relatively quiescent gas joins onto surrounding, more turbulent, material. In fact, it has long been realized that the pattern of HCN hyperfine anomalies requires the presence of an envelope of relatively high excitation temperature \\citep{g93}. \\citet{go09} have recently simulated the birth of dense cores out of spherical, converging flows. Such calculations, when extended to three dimensions, may offer additional insight into the nature of the core-envelope transition. Finally, we must face the obvious fact that real starless cores are not spherical, as the model currently assumes. The shapes of these objects may be partially determined by the anisotropic pressure of their turbulent environments, as found in recent simulations \\citep[e.g.][]{o08}. Another important factor setting the gross core properties is the force associated with the internal magnetic field \\citep{l04}. It would be instructive to study, using the method of Paper~I, the onset of collapse in marginally stable, magnetostatic equilibria, such as those found by \\citet{t88}. Such an investigation would add a significant degree of realism to our picture of protostellar collapse." }, "1005/1005.3599_arXiv.txt": { "abstract": "{The exact period determination of a multi-periodic variable star based on its luminosity time series data is believed a task requiring skill and experience. Thus the majority of available time series analysis techniques require human intervention to some extent.} {The present work is dedicated to establish an automated method of period (or frequency) determination from the time series database of variable stars.} {Relying on the \\sc SigSpec \\rm method (Reegen 2007), the technique established here employs a statistically unbiased treatment of frequency-domain noise and avoids spurious (i.\\,e. noise induced) and alias peaks to the highest possible extent. Several add-ons were incorporated to tailor \\sc SigSpec \\rm to our requirements. We present tests on 386 stars taken from ASAS2 project database.} {From the output file produced by \\sc SigSpec, \\rm the frequency with maximum spectral significance is chosen as the genuine frequency. Out of 386 variable stars available in the ASAS2 database, our results contain 243 periods recovered exactly and also 88 half periods, 42 different periods etc.} {\\sc SigSpec \\rm has the potential to be effectively used for fully automated period detection from variable stars' time series database. The exact detection of periods helps us to identify the type of variability and classify the variable stars, which provides a crucial information on the physical processes effective in stellar atmospheres.} \\keywords { Astronomical instrumentation, methods and techniques -- Stars: variables: general } ", "introduction": "Iterative measurements of a quantity over time yield a time series. This work deals with time series obtained from photometric observations of variable stars, which is the photometric flux (or magnitude) versus time, with or without data gaps. The majority of astronomical measurements cannot be taken continuously over long periods of time due to several reasons. For ground based observations, daylight and the weather conditions are unavoidable sources of data gaps, and observations from space may suffer from cosmic particle impacts or stray light corruption occasionally producing data points beyond repair (Reegen et al. 2006). The method to extract physical information from time-resolved data is commonly known as time series analysis. The present paper is primarily dedicated to regular variables, which exhibit a strictly periodic photometric signal, so the primary goal of our analysis will be to identify the appropriate frequency in case of a mono-periodic star. If the detected frequency is exact, the intrinsic scatter of a phased light curve will attain a minimum. For multi-periodic variables, a step-by-step prewhitening procedure has established and is widely used in combination with multi-sine least-square fitting techniques (Sperl 1998; Lenz \\& Breger 2005). The identified periods permit to identify the type of variability and deduce important astrophysical parameters. The Fourth Variable Star Working Group meeting held at Geneva Observatory, Switzerland in 2005 elaborates about the Period Search Benchmarks (Laurent Eyer)\\footnote{\\tt http://obswww.unige.ch/\\textasciitilde eyer/VSWG/Meeting4/MINUTES/g\\\\aia-vswg4.html}. ", "conclusions": "We investigated several existing programs and software packages which are used for the time series analysis of variable stars and found that human intervention is required in some way to confirm the true period. Full automation could be achieved only with {\\sc SigSpec} combined with few additional subroutines. The majority of results are found to be satisfying with the published frequencies. Thus {\\sc SigSpec} turned out to be a useful utility for future massive variability surveys and also for the re-analysis of already existing time series data of periodic variables." }, "1005/1005.1887_arXiv.txt": { "abstract": "{Stellar density and bar strength should affect the temperatures of the cool (T~$\\sim 20-30$~K) dust component in the inner regions of galaxies, which implies that the ratio of temperatures in the circumnuclear regions to the disk should depend on Hubble type. We investigate the differences between cool dust temperatures in the central 3~kpc and disk of 13 nearby galaxies by fitting models to measurements between 70 and 500~$\\mu$m. We attempt to quantify temperature trends in nearby disk galaxies, with archival data from \\emph{Spitzer}/MIPS and new observations with \\emph{Herschel}/SPIRE, which were acquired during the first phases of the \\emph{Herschel} observations for the KINGFISH (key insights in nearby galaxies: a far-infrared survey with \\emph{Herschel}) sample. We fit single-temperature modified blackbodies to far-infrared and submillimeter measurements of the central and disk regions of galaxies to determine the temperature of the component(s) emitting at those wavelengths. We present the ratio of central-region-to-disk-temperatures of the cool dust component of 13 nearby galaxies as a function of morphological type. We find a significant temperature gradient in the cool dust component in all galaxies, with a mean center-to-disk temperature ratio of $1.15 \\pm 0.03$. The cool dust temperatures in the central $\\sim3$~kpc of nearby galaxies are 23 ($\\pm3$)\\% hotter for morphological types earlier than Sc, and only 9 ($\\pm3$)\\% hotter for later types. The temperature ratio is also correlated with bar strength, with only strongly barred galaxies having a ratio over 1.2. The strong radiation field in the high stellar density of a galactic bulge tends to heat the cool dust component to higher temperatures, at least in early-type spirals with relatively large bulges, especially when paired with a strong bar.} ", "introduction": "The infrared emission from galaxies contains roughly half of the entire energy budget in the Universe (e.g., Hauser \\& Dwek \\cite{hauser01}). In addition to providing information on the amount of attenuation suffered by the stellar light, the infrared emission provides clues to important physical quantities, such as the metal, dust, and cold gas content of galaxies (e.g., Draine et al.\\ \\cite{draine07a}, Bernard et al.\\ \\cite{bernard08}). Infrared spectral energy distributions (SEDs), especially those extending into the submillimeter regime, can be used to measure the dust mass and temperature in galaxies (e.g., Dunne et al. \\cite{dunne00,dunne01}; Seaquist et al.\\ \\cite{seaquist04}; Vlahakis et al.\\ \\cite{vlahakis05}; Draine et al.\\ \\cite{draine07a}; Willmer et al.\\ \\cite{willmer09}; Liu et al.\\ \\cite{liu10}). These temperature measurements have shown that the dust is warmer in the centers of galaxies than in the outskirts (e.g., Alton et al.\\ \\cite{alton98}; Radovich et al.\\ \\cite{radovich01}; Melo et al.\\ \\cite{melo02}; Dupac et al.\\ \\cite{dupac03}). Cool dust at roughly the same temperature in spiral disks is detected {\\it globally} at longer wavelengths (850~$\\mu$m and 1.3~mm; Siebenmorgen et al.\\ \\cite{siebenmorgen99}; Dunne et al.\\ \\cite{dunne00,dunne01}; Vlahakis et al.\\ \\cite{vlahakis05}), but there is also some evidence of a warmer temperature component associated with the central regions. Warmer dust temperatures tend to be associated with significant star-formation activity, the resulting intense interstellar radiation field (Stevens et al.\\ \\cite{stevens05}) and the earlier Hubble type (Bendo et al.\\ \\cite{bendo03}). Here we use the infrared SEDs, from 70 to 500~$\\mu$m (a range which should be dominated by emission from grains in thermal equilibrium with the radiation field, e.g., Popescu et al.\\ \\cite{popescu00}, Engelbracht et al.\\ \\cite{engelbracht08}), of a local sample of 13 galaxies spanning the range of spiral galaxies in the Hubble sequence, to derive the temperature of the cool (T$ \\sim 20-30$~K) dust component of the central region and the disks separately, and investigate differences in the dust heating in the two regions. To achieve this goal, we use the 250, 350, and 500~$\\mu$m \\emph{Herschel}/SPIRE (Spectral and Photometric Imaging REceiver; Griffin et al.\\ \\cite{griffin10}) images of the galaxies combined with the \\emph{Spitzer}/MIPS (Multiband Imaging Photometer for \\emph{Spitzer}; Rieke et~al.\\ \\cite{rieke04}) 70 and 160~$\\mu$m images. Eventually we will use PACS (Photodetector Array Camera and Spectrometer; Poglitsch et al.\\ \\cite{poglitsch10}) imaging for this study, but at the time of this writing, the data were not yet available. These galaxies have little nuclear activity that might heat the dust, with only NGC~1097 having a strongly active nucleus, so this is the first study to cleanly separate the far-infrared properties of central and disk regions in a sample of normal galaxies. Until recently, little work has been done to dissect the dust emission in galaxies into sub-galactic components, owing to the general paucity of infrared images with the required angular resolution and to poor long-wavelength sensitivity. Some recent work includes a study of the galaxy pair NGC1512/1510 (a target also discussed in this paper), which finds that the dust temperature in the central region of NGC1512 is slightly higher than in the disk and that there is a significantly higher fraction of warm dust, in agreement with the center of NGC1512 being a starburst (Liu et al.\\ \\cite{liu10}). Work by Mu\\~{n}oz-Mateos et al.\\ (\\cite{munoz-mateos09}) examines radial trends in dust properties in a number of nearby galaxies. The \\emph{Herschel} Space Telescope promises to yield a breakthrough in the study of subgalactic components in galaxies. This paper is the first investigation to leverage the longest wavelength \\emph{Herschel} data available for the KINGFISH (key insights on nearby galaxies: a far-infrared survey with \\emph{Herschel}; this program is largely derived from SINGS, the \\emph{Spitzer} infrared nearby galaxy survey by Kennicutt et al.\\ \\cite{kennicutt03}) sample of nearby galaxies, which will eventually total 61. Companion papers from the science demonstration phase for this program showcase the shorter wavelength imaging (Sandstrom et al.\\ \\cite{sandstrom10}) and the spectroscopic data (Beir\\~{a}o et al.\\ \\cite{beirao}). Here we present new SPIRE images acquired during the first few months of \\emph{Herschel} operations, in the context of the KINGFISH Open-Time Key Project. We divide each galaxy into two spatially-resolved zones: the circumnuclear region and the surrounding disk. Then we compare central temperatures with those for the disk. ", "conclusions": "We used far-infrared data from \\emph{Spitzer} and submillimeter data from \\emph{Herschel} to compute separate SEDs for the center and disk regions of 13 nearby galaxies observed as part of the KINGFISH program. We fit those SEDs (at wavelengths longer than 70~$\\mu$m) with blackbody functions (modified by a frequency-dependent emissivity) to compute temperatures. On average, the cool dust temperature of the central component is $15 \\pm 3$\\% hotter than the disk. We find that the central temperature is higher than the disk by 20\\% to 50\\% in galaxies of type S0 to Sb, but only 9\\% higher in later types. This ratio is also higher (at $1.29 \\pm 0.04$) in strongly barred galaxies than in weakly barred galaxies (at $1.09 \\pm 0.03$). The data therefore indicate that the large (or ``classical'') grains that dominate the far-infrared and submillimeter emission are warmer in the centers of those galaxies with a substantial bulge and/or a strong bar. This may simply be caused by the higher density of the radiation field in the centers of early-type spirals, enhanced star formation due to the bar, or some combination of the two. A cleaner separation of morphological components (perhaps with larger samples and/or less distant galaxies) and a more thorough assessment of the density of starlight and star formation activity, plus an evaluation of the impact of central nonthermal sources, may help separate these effects. The analysis presented here illustrates the power of {\\it Herschel} observations in characterizating the spatially resolved distribution of dust in nearby galaxies. This power will grow with the use of better-resolved far-infrared SEDs as measured by PACS (Poglitsch et al.\\ \\cite{poglitsch10}), which will let us measure smaller and/or more distant galaxies and determine radial trends of dust temperature." }, "1005/1005.0848_arXiv.txt": { "abstract": "{High precision astrometry requires an accurate geometric distortion solution. In this work, we present an average correction for the Blue Camera of the Large Binocular Telescope which enables a relative astrometric precision of $\\sim$15 mas for the $B_{\\rm Bessel}$ and $V_{\\rm Bessel}$ broad-band filters. The result of this effort is used in two companion papers:\\ the first to measure the absolute proper motion of the open cluster M~67 with respect to the background galaxies;\\ the second to decontaminate the color-magnitude diagram of M~67 from field objects, enabling the study of the end of its white dwarf cooling sequence. Many other applications might find this distortion correction useful.} ", "introduction": "\\label{sec1} Modern wide field imagers (WFI) equipped with CCD detectors began their operations at the end of the last century, however -- after more than 10 years -- their astrometric potential still remains somehow unexploited (see Anderson et al.\\ \\cite{anderson06}, hereafter Paper~I). It is particularly timely to begin exploring their full potential now that WFI start to appear also at the focus of the largest available 8m-class telescopes. The present work goes in this direction, presenting a correction for the geometric distortion (GD) of the Blue prime-focus Large Binocular Camera (LBC), at the Large Binocular Telescope (LBT). Unlike in Paper~I, in which we corrected the GD of the WFI at the focus of the 2.2m MPI/ESO telescope (WFI@2.2m) with a look-up table of corrections, for the LBC@LBT we will adopt the same technique described in Anderson \\& King (\\cite{AK03}, hereafter AK03), and successfully applied to the new Wide field Camera 3/UV-Optical channel on board the {\\it Hubble Space Telescope} (Bellini \\& Bedin \\cite{bb09}, hereafter BB09). This article is organized as follows: Section 2 briefly describes the telescope/camera set up; Section 3 presents the data set used. In section 4, we describe the steps which allowed us to obtain a solution of the GD, for each detector separately, while in Section 5 we presents a (less accurate) inter-chip solution. Distortion stability is analyzed in Section 6, and a final Section summarizes our results. \\begin{figure*}[ht!] \\centering \\includegraphics[height=10.2cm]{13783f1a.ps} \\includegraphics[height=10.0cm]{13783f1b.ps}\\\\ \\caption{(\\textit{Left}): LBC-Blue mosaic layout; ``$\\ast$'' marks the center for each chip (see Sect.\\ \\ref{sec:3} for the operative definition of centers and chips), while ``+'' marks the here defined center of the mosaic. (\\textit{Right}): Each MEF file consists of 4 images. Each image is composed by one scientific region and two overscan regions, covering the first 50 and the last 206 pixel columns (shaded).} \\label{fig1} \\end{figure*} ", "conclusions": "\\label{sec:10} By using a large number of well dithered exposures we have found a set of third-order-correction coefficients for the geometric distortion solution of each chip of the LBC-Blue, at the prime focus of the LBT. The use of these corrections removes the distortion over the entire area of each chip to an accuracy of $\\sim$0.09 pixel (i.e.\\ $\\sim$20 mas), the largest systematics being located in the 200-400 pixels closest to the boundaries of the detectors. Therefore, we advise the use of the inner parts of the detectors for high-precision astrometry. The limitation that has prevented us from removing the distortion at even higher level of accuracies -- in addition to atmospheric effects and to the relatively sparsity of the studied field -- is the dependency of the distortion on the scale changes that result from thermal and/or gravitational induced variations of the telescope+optical structure. If a dozen (or more) well distributed high S/N stars are available within the same chip, a general 6-parameter linear transformation could register relative positions in different images down to about 15 mas. If the field is even more densely populated, then a local transformation approach [as the one adopted in Bedin et al.\\ (\\cite{bedin03}), from space, or in Paper~I, II, III) from ground] can further reduce these precisions to the mas level. [Indeed, using these techniques and this very same data-set we were able to reach a final precision of $\\sim$1 mas$\\,$yr$^{-1}$ (Bellini et al., submitted to \\aap\\ Letters)]. These are the precisions and accuracies with which we can hope to bring one image into another image by adopting:\\ {\\it conformal}, {\\it general}, or {\\it local} transformations. In the case of absolute astrometry, however, the accuracies are much lower. During the available limited number of nights of observations (and atmospheric conditions), we observed scale-variations up to 5 parts in $10\\,000$, even during the same night. This implies that astrometric accuracy -- which completely relies on our GD solution -- can not be better than $\\sim$250 mas ($\\sim$1 pixel) within a given chip (from center to corners), and can be as large as $0\\farcs5$ ($\\sim$2 pixels) in the meta-chip system. This value is in-line with the meta-chip stability observed in other ground-based WFI (Paper~I), and absolutely excellent for a ground-based prime-focus instrument with such a small focal ratio and large FoV. Thankfully, several stars from astrometric catalogs such as the UCAC-2, GSC-2, 2MASS, will be always available within any given LBC-Blue large FoV. These stars, in addition to provide a link to absolute astrometry (as done for example in Rovilos et al.\\ 2009), will enable constrains of linear terms in our GD solution, and to potentially reach an absolute astrometric precision of 20 mas. The fact that we are able to reach good astrometric precision also for saturated stars will make the comparison between these catalogs and the sources measured in the -- generally deeper -- LBC images, even easier. For the future, more data and a longer time-baseline are needed to better characterize the GD stability of LBC@LBT detectors on the medium and long time term. This could make it possible to:\\ (1) determine a multi-layer model of the distortion which would properly disentangle the contributions given by optical field-angle distortion, light-path deviations caused by filters and windows, non-flat CCDs, CCDs artifacts, alignment errors of the CCD on the focal plane, etc.; and (2) allow for time-dependent and/or mis-alignments of mirrors, filters/windows, and CCDs." }, "1005/1005.5695_arXiv.txt": { "abstract": "The photon polarization operator in superstrong magnetic fields induces the dynamical photon ``mass'' which leads to screening of Coulomb potential at small distances $z\\ll 1/m$, $m$ is the mass of an electron. We demonstrate that this behaviour is qualitatively different from the case of $D=2$ QED, where the same formula for a polarization operator leads to screening at large distances as well. Because of screening the ground state energy of the hydrogen atom at the magnetic fields $B \\gg m^2/e^3$ has the finite value $E_0 = -me^4/2 \\ln^2(1/e^6)$. ", "introduction": "The Larmour radius of the electron orbit $a_H = 1/\\sqrt{eB}$ is much smaller than Bohr atomic radius $a_B = 1/(me^2)$ for homogenius magnetic fields $B\\gg m^2 e^3$ (we are using Gauss system of units, where $e^2 = \\alpha = 1/137$; also in all formulas $\\hbar = c = 1$). It is natural to look for the atomic energy levels in such strong magnetic fields studing the influence of Coulomb potential on the electrons occupying Landau levels \\cite{Lan}. A strong magnetic field confines an electron in the transverse direction while in the longitudinal direction an electron is bound by the weak Coulomb field of a nucleus. The cigar-shape wave function of an electron is formed with transverse size which equals Larmour radius and longitudinal size which is by $\\ln(a_B^2/a_H^2) \\equiv \\ln(B/m^2 e^3)$ smaller than Bohr radius. The ground state energy is larger than Rydberg constant by the square of the same logarithm: $E_0 = -(me^4/2)\\ln^2(B/m^2 e^3)$. One can easily get this logarithmic factor from the fact that in one-dimensional Coulomb potential energy diverges logarithmically at small distances. The divergency is regularized at the longitudinal distances which equals $a_H$, where one-dimensional motion converts to a three-dimensional one. Atomic levels in such strong magnetic fields were found numerically in \\cite{1} (see also \\cite{2, 22}). Our purpose is to understand the behaviour of the energy levels with the growth of a magnetic field. The point is that at superstrong magnetic fields $B\\ga m^2/e^3$ the polarization operator insertions into the photon propagator induce the dynamic photon ``mass'' $m_\\gamma^2 \\approx e^3 B$ \\cite{3, 33}. One would expect that the photon mass should screen Coulomb potential and shift energies of the atomic levels found in tree approximation. Dirac equation spectrum in a constant homogenious magnetic field looks like \\cite{4}: \\begin{equation} \\varepsilon_n^2 = m^2 + p_z^2 + (2n+1)eB + \\sigma eB \\;\\; , \\label{1} \\end{equation} where $n = 0,1,2, ...$, $\\sigma = \\pm 1$ and the field is directed along axis $z$.\\footnote{This spectrum with the substitution of $2n+1+\\sigma$ by $2j, \\,\\,j=0,1,2,... $ was found by I.I. Raby \\cite{IR}.} In the magnetic fields we are interested in $\\varepsilon_n \\ga m/e$, and electrons are ultrarelativistic. The only exception is the lowest Landau level (LLL) which has $n=0$, $\\sigma = -1$. The energy of LLL electron equals its mass for $p_z = 0$ and the consideration of the nonrelativistic electron motion along $z$ axis is selfconsistent. LLL is interesting both practically and theoretically. An analog of the critical electric field $E_{cr} = m^2/e$ is the magnetic field $B_0 = m^2/e = 4.4 \\cdot 10^{13}$ gauss. Two orders larger superstrong fields $B\\ga m^2/e^3$ can exist at special neutron stars named magnetars. The temperature of an outer magnetar layer is not enough to populate the excited Landau levels and one can observe the transitions among the states to which LLL is splitted at the electric field of the nucleus. Freezing of the ground state energy in the superstrong magnetic fields discussed in the paper leads to the upper bound on the spectra of photons radiated from magnetars.\\footnote{I am grateful to S.I. Blinnikov for the discussions of magnetar physics.} To study the stability of the huge magnetic fields \\cite{5} one should also know the energy of the ground state as a function of a field. So we are studying the energies of the states to which LLL splits in the presence of an atomic nucleus. Since the electron at LLL moves along $z$ axis we will study in section 2 QED at $D=2$: the behaviour of electrons in two-dimensional space-time. The coupling constant $g$ has dimension of mass, so Coulomb potential as a function of $|z|$ depends on two dimensionfull parameters: $g$ and electron mass $m$. We will obtain the approximate analytical formula for Coulomb potential in $d= D-1 = 1$ which takes into account the photon polarization operator. We will see that for large $g$ (or small $m$) $g \\gg m$ Coulomb potential is screened. In section 3 we will consider the physical case, $D =4$ QED. The analog of the coupling constant squared $g^2$ in the real world is the product $e^3 B$. The polarization operator in the magnetic fields $B\\gg m^2/e$ at $k_\\parallel ^2(\\equiv k_z^2) \\ll eB$ practically coincides with the one obtained in section 2 \\cite{77}. Nevertheless the screening at large distances $|z| \\gg 1/m$ does not occur: at $|z| \\gg 1/m$ we get a purely Coulomb potential $\\Phi(z) = e/|z|$. The screening occurs at small distances, and its influence on the ground state energy is determined in section 4. The results similar to those presented in sections 3 and 4 were obtained in \\cite{55} with the help of the numerical calculations. ", "conclusions": "The photon polarization operator leads to modifications of the atomic energy levels. The famous example is its contribution to the Lamb shift, the difference of the energies of $2s_{1/2}$ and $2p_{1/2}$ levels of hydrogen. They are numerically small loop corrections to the values of energies determined by the tree level potential. The role of the photon polarization diagram in the superstrong magnetic fields $B > m^2/e^3 = 6.2 \\cdot 10^{15}$ gauss is qualitatively different. It determines the behaviour of the ground state energy: the formula obtained at tree level becomes invalid and the growth of the coupling energy with $B$ terminates at $B\\approx m^2/e^3$. Screening of Coulomb potential should be more important for the energies of even excited states which are more sensitive to the shape of the potential at small distances \\cite{155}. Degeneracy of even and odd excited states in the limit $B\\Longrightarrow\\infty$ is not lifted by the screening. We study the analogy of the electric potential in $d=1$ QED with massive electrons and in $d=3$ QED in strong magnetic fields $B > B_0 = m^2/e$ which originates from the coincidence of the polarization operators in these cases. A simple analytical expression which equals the polarization operator with 10\\% accuracy enables us to obtain an approximate formula for the electric potential of the point charge in $d=1$ QED with massive fermions and asymptotics of the potential in $d=3$ QED. In $d=1$ QED for a coupling constant $g$ larger than a fermion mass $m$ a tree level formula is modified at $|z|> 1/g$. In $d=3$ QED a tree level formula is modified at the distances $1/m> |z| > 1/\\sqrt{e^3 B}$ while at large distances $|z|> 1/m$ Coulomb law is valid. Analogous results for $D=4$ were obtained in \\cite{55}. The other aspect of the Coulomb potential in the strong magnetic field is investigated in paper \\cite{10}: it is supposed that fermions obtained their mass due to a magnetic field (dynamical fermion mass). I am grateful to S.I. Blinnikov, V.A. Novikov, L.B. Okun, V.S. Popov, and A.V. Smilga for useful discussions and to A.I.Rez, who brought to my attention Phys. Rev. publication \\cite{55}. This work was supported by the grants RFBR 08-02-00494, Nsh-4172.2010.2 and by the contract of the RF Ministry of Science and Education No. 02.740.11.5158." }, "1005/1005.5376_arXiv.txt": { "abstract": "The Sun's supergranulation refers to a physical pattern covering the surface of the quiet Sun with a typical horizontal scale of approximately 30~000~kilometres and a lifetime of around 1.8~days. Its most noticeable observable signature is as a fluctuating velocity field of 360~m/s rms whose components are mostly horizontal. Supergranulation was discovered more than fifty years ago, however explaining why and how it originates still represents one of the main challenges of modern solar physics. \\smallskip A lot of work has been devoted to the subject over the years, but observational constraints, conceptual difficulties and numerical limitations have all concurred to prevent a detailed understanding of the supergranulation phenomenon so far. With the advent of 21st century supercomputing resources and the availability of unprecedented high-resolution observations of the Sun, a stage at which key progress can be made has now been reached. A unifying strategy between observations and modelling is more than ever required for this to be possible. \\smallskip The primary aim of this review is therefore to provide readers with a detailed interdisciplinary description of past and current research on the problem, from the most elaborate observational strategies to recent theoretical and numerical modelling efforts that have all taken up the challenge of uncovering the origins of supergranulation. Throughout the text, we attempt to pick up the most robust findings so far, but we also outline the difficulties, limitations and open questions that the community has been confronted with over the years. \\smallskip In the light of the current understanding of the multiscale dynamics of the quiet photosphere, we finally suggest a tentative picture of supergranulation as a dynamical feature of turbulent magnetohydrodynamic convection in an extended spatial domain, with the aim of stimulating future research and discussions. ", "introduction": "The story of supergranulation really started in Oxford when Avril B. Hart reported in 1953 the existence of a ``noisy'' fluctuating velocity field on top of the mean rotation speed of the solar equator that she was measuring \\citep{hart54}. Actually, it is most probable that this ``noise\" was already detected as early as 1915 by \\cite{plaskett16}. Two years after her first detailed report, \\cite{hart56} confirmed the discovery and was able to give the first estimate\\epubtkFootnote{In solar physics, an appropriate length unit is the megametre (Mm), also 1000~km.} of 26~Mm for the typical horizontal length scale of these ``velocity fluctuations\" (sic). Supergranulation was further recognised as a characteristic feature of the whole surface of the quiet Sun (the regions of weak magnetic fields, which represent the most important part of the solar surface) after the seminal work of \\cite{leighton62}, who published the first Doppler images of the Sun (and also the first detection of the five minutes oscillations). This work was soon after completed by another important paper by \\cite{SL64} who showed, amongst other results, the intimate relation between supergranulation and the magnetic network of the quiet Sun. \\smallskip It is remarkable that all the fundamentals of supergranulation have basically been uncovered over that 1954--1964 decade. Since then, progresses have been much less spectacular, especially on the theoretical side. This is certainly why supergranulation is still a fascinating subject. We are still wondering where it comes from, what its exact relation with magnetic fields is, if it is a universal feature of solar type stars, or of stellar surface convection, if it plays a role in the solar dynamo(s), etc. All these questions are pending fifty years after the discovery. There are many reasons why the solar physics community has not yet managed to answer them, several of which are not actually specific to the supergranulation problem. Most of these reasons are described in detail in this review, but it is worth pointing out a few important issues here as an introduction. \\smallskip For a long time only a limited set of observables and restricted time-records of the evolution of the supergranulation pattern have been available, may it be on short (24 consecutive hours) or long timescales (a solar cycle). This has somewhat hindered the study of the detailed spatial structure of supergranulation and the identification of the physical factors that affect it (buoyancy, stratification, magnetic fields, rotation). From the theoretical and numerical perspectives, on the other hand, the strongly nonlinear physical nature of magnetised thermal convection in the outer layers of the Sun makes it extremely difficult to come up with a simple, unique, verifiable physical model of the process (a similar problem arises in many subfields of solar physics, if not of astrophysics). Overall, these observational uncertainties and theoretical or numerical limitations have somewhat negatively interfered to prevent a rigorous solution to the supergranulation problem. The currently fairly obfuscated state of affairs may nevertheless get clarified in a near future, as the solar physics community is now armed with both high-resolution solar observatories such as Hinode\\epubtkFootnote{Hinode is led by the Japanese Aerospace Exploration Agency (JAXA) in collaboration with NASA, the Science and Technology Facilities Council (STFC), and the European Space Agency (ESA). Hinode is a Japanese mission developed, launched and operated by ISAS/JAXA, in partnership with NAOJ, NASA and STFC (UK). Additional operational support is provided by ESA and NSC (Norway). The project website can be found at \\url{http://solarb.msfc.nasa.gov/}} and large supercomputers that allow for increasingly realistic numerical simulations of the complex solar surface flows. However, it is clear that stronger connections between theory, numerics and observations need to be established for the problem to be resolved. \\smallskip In this review, we would like to introduce readers to this subject by first describing the full range of past and current research activities pertaining to the solar supergranulation, from the breadth of historical and modern observational results to the most elaborate numerical models of supergranulation convection. We particularly wish to provide a useful guide to the abundant literature related to that theme, to point out the important findings in the field, but also to stress the limitations and difficulties that have been encountered over the years in order to help overcome them. To this end, we attempt to discuss the already existing or possible connections between various pieces of research and try to identify some important questions whose answer may be crucial to understand how and why supergranulation originates. \\smallskip The review is divided into eight parts, including this introduction. The next two sections offers some introductory material on the physics of deep and surface convection in the Sun (Section~\\ref{generalconsiderations}) and a brief recap on small-scale flows, namely granulation and mesogranulation (Section~\\ref{granule}). Section~\\ref{obs} is dedicated to a presentation of observational facts that have been collected on supergranulation. We then carry on with the discussion of existing theoretical models to explain the origin of supergranulation in Section~\\ref{theory}. In Section~\\ref{numerics}, we discuss related numerical experiments. Our current knowledge on supergranulation is summarised and commented in Section~\\ref{wrapup} for the convenience of hurried readers. In the light of the present understanding of multiscale dynamics of the quiet photosphere, we finally suggest a tentative picture of supergranulation as a dynamical feature of turbulent magnetohydrodynamic (MHD) convection in an extended spatial domain, with the aim of stimulating future research and discussions. We notably propose several numerical and observational diagnostics that could help make important progress on the problem in the near future (Section~\\ref{discussion}). \\smallskip We tried to make the paper readable by all astrophysicists, assuming only little background in that field and trying to avoid as much as possible the solar physicists jargon, or to explain it when necessary. ", "conclusions": "} This long tour of the main observational, theoretical and numerical results on the problem of supergranulation being completed, we are now in a position to provide hurried (as well as less hurried) readers with a synthetic presentation of what has been learned so far on the supergranulation phenomenon and what are the current issues. The presentation of a more personal outlook and suggestions for future research is deferred to Section~\\ref{discussion}. \\subsection{Observations} The Sun's supergranulation is a large-scale coherent pattern detected in the surface layers of the quiet Sun. The \\emph{impression} given by observations is that it is simply superimposed on a stochastic, highly nonlinear background smaller-scale flow pattern, the granulation. Characterising the supergranulation velocity pattern requires monitoring solar surface flows over long times, over wide fields of views, or over a large set of independent observations. The properties of the supergranulation velocity field can be summarised as follows: \\begin{itemize} \\item the \\emph{length scale} of the supergranulation flow, as given by the kinetic energy power spectrum of the horizontal component of solar surface flows, is in the range of 20--70~Mm with a preferred scale of 36~Mm (Section~\\ref{length_sc}). These results come from both Dopplergrams \\citep{hathaway2000} and from granule tracking in wide-field high resolution image series \\citep{rieutord08}. The size of the field must be sufficiently large to secure the statistical convergence of the results. \\item The typical \\emph{size} of supergranules, defined as coherent diverging flow cells at the solar surface, is in the range 10--30~Mm \\citep{hirzberger08}. The derived average size is sensitive to the method used to identify supergranules. \\item The \\emph{kinetic energy excess} associated with \\emph{supergranulation} in the power spectrum of solar surface flows lies \\emph{on the large-scale side of the injection range of photospheric turbulence} located at the granulation scale (Section~\\ref{length_sc} and Figure~\\ref{figure:spectra}). \\item The most recent estimate of the \\emph{lifetime} of supergranules, based on the largest sample of supergranules collected so far, is 1.6$\\pm0.7$~day \\citep{hirzberger08}. The dispersion in the measurements of supergranules lifetimes is also fairly large (Section~\\ref{time_sc}). \\item Rms \\emph{horizontal velocities} at supergranulation scale are of the order of 350~m/s, while rms \\emph{vertical velocities} are around 30~m/s \\citep{hathaway2002}. As velocities depend on the scales considered, the relation between amplitude and scale, namely the power spectrum, provides the most suitable observable to estimate the amplitude of the supergranulation velocity field (Section~\\ref{velo_sc}). \\item Local helioseismology indicates that supergranules are \\emph{shallow structures} (Section~\\ref{obsdepth}), possibly not deeper than 5~Mm \\citep{sekii07}. The mean vertical profile of the supergranulation flow is not very well constrained at the moment. More precise determinations are definitely called for. \\end{itemize} Note that the foregoing determinations are not independent of each other, because velocity scales can be derived from the combination of length and time scales. Namely, 30~Mm divided by 1.7 day gives 205~m/s, which is in reasonable agreement with direct measurements of supergranulation-scale velocities. In our view, the computation of the power spectra of solar surface flows provides one of the most robust methods to make progress on the determination of these various quantities in the future. Most notably, an accurate determination of the \\emph{vertical velocity spectrum} of vertical velocities in the supergranulation range is still lacking. \\smallskip Besides this set of typical scales associated with the supergranulation velocity pattern, several other observational signatures and properties of supergranulation have been studied: \\begin{itemize} \\item \\emph{horizontal intensity fluctuations} at supergranulation scales are very faint (Section~\\ref{obsintensity}). The latest studies indicate that supergranules are slightly warmer at their centre. The temperature drop is less than 3~K \\citep{meunier07b,goldbaum_etal09}. \\item Supergranulation is affected by the \\emph{global solar rotation} (Section~\\ref{obsrot}). Locally, supergranules are anticyclonic structures \\citep{gizon03a}, their mean vertical vorticity is negative in the northern hemisphere and positive in the southern one. The supergranulation pattern has been observed to propagate anisotropically in the prograde direction \\citep{gizon03b}. \\item Supergranulation has \\emph{dynamical interactions} with the \\emph{magnetic fields} of the quiet Sun. Most notably, supergranules are strongly correlated with the magnetic network (Section~\\ref{obsmag}). Correlations between the size of supergranules and the strength of network and internetwork fields have been evidenced recently \\citep{meunier07a}. The solar-cycle dependence of the pattern remains uncertain though, as various papers have been giving contradicting results. \\end{itemize} The bounds on intensity variations seem to be well established now. The proper rotation of supergranules, as measured by their local mean vertical vorticity, is also well constrained by local helioseismology. On the other hand, we believe that more work is required to quantitatively constrain the interaction of supergranules with magnetic fields. A determination of the \\emph{magnetic energy spectrum} of the quiet Sun over a wide range of scales would be extremely useful to put constraints on the physical processes at the origin of the network and internetwork fields and on their interactions with supergranulation (see Section~\\ref{discussion} below). \\smallskip Figure~\\ref{figure:super_pict} is an attempt to depict the standard view of the supergranulation phenomenon, as constrained by the observations summarised above. \\epubtkImage{fig/super_pict.pdf}{ \\begin{figure}[ht] \\centerline{\\includegraphics[width=1.\\linewidth]{fig/super_pict}} \\caption[Schematic view of the supergranulation phenomenon, as constrained by observations.]{A schematic view of the supergranulation phenomenon, as constrained by observations. $\\lambda$ is the scale where the horizontal kinetic energy spectral density is maximum. $d$ is the diameter of ``coherent structures'' (supergranules). The red and blue patches depict the warm and cold regions of the flow. I.N.B denotes the internetwork magnetic field. Note that the indicated internetwork and network fields geometries roughly correspond to the standard historical picture of quiet Sun magnetic fields and their relation to supergranulation (Section~\\ref{obsmag}). As discussed in Sections~\\ref{obsINB} and \\ref{discussion}, this picture must be significantly nuanced in reality, as the dichotomy between network and internetwork fields is probably not quite as clear as indicated in this drawing.} \\label{figure:super_pict} \\end{figure} } \\subsection{Theory} Two major types of physical scenarios have been suggested to explain the origin of supergranulation: \\begin{itemize} \\item \\emph{thermal convection scenarios}, in which buoyancy is the main driver of the supergranulation flow (Section~\\ref{theoryconv}). Various effects (magnetic fields, shear, rotation, effective boundary conditions) have been explored within the framework of linear and weakly nonlinear theory to explain the size of of the supergranulation pattern, its weak thermal signature, its oscillations and propagation. \\item \\emph{Collective interaction scenarios}, whereby supergranulation emerges as a large-scale coherent pattern triggered by nonlinear interactions of vigorous smaller-scale structures like granulation (see \\cite{rieutord2000} and Section~\\ref{theoryLSinst}). These scenarios have mostly been explored quantitatively through ``toy model'' simulations \\citep{rast03b,crouch07} that do not incorporate the full complexity of dynamical MHD equations. In our view, direct numerical simulations provide the most promising way of making progress on this side in the future. \\end{itemize} A shared property of all models is the looseness of the approximations on which they rely (e.g. linear theory with turbulent viscosity parametrisation, or purely phenomenological arguments on the nature of dynamical interactions between granules and their potential large-scale instabilities). Completely distinct theoretical arguments can easily be tuned to produce results that are all broadly consistent with observations. This degeneracy makes it impossible to discriminate between various scenarios and to come up with a proper theoretical explanation for the origin of supergranulation that could be unambiguously validated by observations. \\smallskip Finally, it is possible but certainly not obvious that supergranulation can be explained quantitatively by a simple mathematical theoretical model. In any case, one of the most urgent tasks to overcome some of the previously mentioned shortcomings is to figure out if the basic assumptions and arguments on which current theoretical models rely (linear theory, effective boundary conditions, convection in uniform magnetic fields etc.) are justified, and to test them quantitatively with the help of large-scale numerical simulations. \\subsection{Numerical simulations} The complexity and nonlinearity of the physical environment of supergranulation is extraordinary: vigorous turbulent small-scale flows in a strongly stratified atmosphere, ionisation physics, rotation, shear and tortuous magnetic fields geometries at all observable scales may all have something to do with the supergranulation phenomenon. As argued several times in this review, numerical simulations have a unique potential for approaching this complexity. They have now become an unavoidable tool to uncover the real nature of supergranulation and to test the various qualitative theoretical pictures described in Section~\\ref{theory}. \\smallskip Numerical simulations dedicated to the supergranulation problem are still in their infancy though, mostly because they remain awfully expensive in terms of computing time. The latest generation of numerical experiments, summarised in Table~\\ref{table:performances}, barely accommodates for the scale of supergranulation. The main results obtained so far are summarised below: \\begin{itemize} \\item \\emph{global spherical simulations} \\citep{derosa02} exhibit a supergranulation-like pattern, but the scale of this pattern is dangerously close to the grid scale of the simulations (Section~\\ref{simglobal}). \\item In \\emph{local large-scale idealised simulations} \\citep{cattaneo01,rincon05}, two patterns can be singled out of the continuum of turbulent scales: a granulation pattern forming in the upper thermal boundary layer, and a larger-scale, extremely energetic mesoscale pattern, which extends through the whole convective layer (Figure~\\ref{figure:ideal}). Whether or not this pattern has anything to do with supergranulation or with the hypothetical solar mesogranulation is not understood (see Section~\\ref{simlargelocal} for an in-depth discussion). \\item \\emph{Local large-scale realistic simulations of hydrodynamic convection} \\citep{stein09} do not exhibit any significant energy excess at supergranulation scales in spite of the presence of Hydrogen and Helium ionizations in the model (Section~\\ref{simlargelocal2}). This result therefore tends to disprove the ``classical'' \\cite{SL64} supergranulation theory. \\item \\emph{Local large-scale realistic simulations of MHD convection} reveal the formation of a magnetic network at scales ranging from mesoscales to supergranulation scales (Section~\\ref{simmhd}). What sets the scale of this network and the emergence of supergranulation as a special scale in these simulations has not been investigated yet, but a recent study \\citep{ustyugov09} suggests that strong magnetic flux concentrations play a significant role in the scale-selection process. \\end{itemize} \\smallskip Numericists will have to address several important issues in the forthcoming years. One of the main problems is that all dedicated simulations to date are still fairly dissipative (much more than the Rayleigh-B\\'enard simulations described in Section~\\ref{numconv}, for instance). Local large-scale simulations, for instance, barely accommodate 10 grid points within a granule. This kind of resolution is not sufficient to capture all the dynamics of solar surface flows, as the viscous and magnetic dissipation scales are both much smaller than 100~km (Section~\\ref{turbulentscales}) at the solar surface. As mentioned in Section~\\ref{realvsideal} and \\ref{numconc}, resolving dissipation scales properly has recently turned out to be essential to make progress on several turbulent MHD problems, such as magnetic field generation (dynamo action) by non-helical turbulent velocity fields. A related point is that uncovering the full dynamical physics of large scales and avoiding spurious finite-box effects requires both very large numerical domains and large integration times of the simulations, which is not ensured in today's experiments. This point is easily illustrated by the supergranulation-scale dichotomy between global and local simulations discussed in Section~\\ref{simlarge}. \\smallskip Overall, the current computing limitations are such that numerical simulations are still far away from the parameter regime typical of the Sun. Hence, one cannot exclude that all simulations to date miss some critical multiscale dynamical phenomena, either purely hydrodynamic or MHD. Large-scale simulations are also currently too expensive for any decent scan of the parameter space of the problem to be possible. However, it is fair to say that the perspective of petaflop computations holds the promise of significant numerical breakthroughs in a ten-years future." }, "1005/1005.2283_arXiv.txt": { "abstract": "{The Sun is a magnetic star whose cyclic activity is thought to be linked to internal dynamo mechanisms. A combination of numerical modelling with various levels of complexity is an efficient and accurate tool to investigate such intricate dynamical processes. } {We investigate the role of the magnetic buoyancy process in 2D Babcock-Leighton dynamo models, by modelling more accurately the surface source term for poloidal field.}{ To do so, we reintroduce in mean-field models the results of full 3D MHD calculations of the non-linear evolution of a rising flux tube in a convective shell. More specifically, the Babcock-Leighton source term is modified to take into account the delay introduced by the rise time of the toroidal structures from the base of the convection zone to the solar surface.} {We find that the time delays introduced in the equations produce large temporal modulation of the cycle amplitude even when strong and thus rapidly rising flux tubes are considered. Aperiodic modulations of the solar cycle appear after a sequence of period doubling bifurcations typical of non-linear systems. The strong effects introduced even by small delays is found to be due to the dependence of the delays on the magnetic field strength at the base of the convection zone, the modulation being much less when time delays remain constant. We do not find any significant influence on the cycle period except when the delays are made artificially strong.}{ A possible new origin of the solar cycle variability is here revealed. This modulated activity and the resulting butterfly diagram are then more compatible with observations than what the standard Babcock-Leighton model produces.} ", "introduction": "Our Sun is a prime example of a very turbulent and magnetically active star. Its robust 22-yr activity cycle originates in the periodic polarity reversal of the Sun's internal large-scale magnetic field. As a result, sunspots emerge at the solar surface with statistically well-defined dynamical and morphological characteristics. These sunspots are thought to be the surface manifestations of strong toroidal magnetic structures created at the base of the convection zone and which rise through the plasma under the effect of magnetic buoyancy (\\cite{Parker55}). This toroidal field undergoes cyclic variations along with the poloidal field, which flips polarity at sunspot cycle maxima. Even if a robust regular activity is easily exhibited in the Sun, a significant modulation of both the amplitude and the frequency of the cycle has been observed. In particular, periods of strongly reduced activity have been revealed, the most famous being the Maunder minimum between 1650 and 1700, during when no sunspots were to be seen at the solar surface (\\cite{Eddy76}). Today, understanding the origins of such a variability has become crucial. Not only is the impact on satellites, astronauts, power grids or radio communications significant but the climatologists community now tends to address more and more the question of the influence of solar variability on the Earth climate (e.g. \\cite{Bard06}, \\cite{Lean05}). The classical explanation for the cyclic activity of the large-scale magnetic field is that a dynamo process acts in the solar interior to regenerate the three components of the magnetic field and sustain them against ohmic dissipation. The inductive action of the complex fluid motions would thus be responsible for the vigorous regeneration of magnetic fields and for its non-linear evolution in the solar interior (see \\cite{Charbonneau051} and \\cite{Miesch05} for recent reviews on the subject). Understanding how these complex physical processes operating in the solar turbulent plasma non-linearly interact is very challenging. One successful and powerful approach is to rely on multi-dimensional magnetohydrodynamic (MHD) simulations. In this context, two types of numerical experiments have been performed since the 70's: kinematic mean-field axisymmetric dynamo models which solve only the mean induction equation (\\cite{Steenbeck69, Roberts72, Stix76, Krause80}) and full 3D global models which explicitly solve the full set of MHD equations (\\cite{Gilman83, Glatzmaier85, Cattaneo99, Brun04}). Those two approaches are complementary and needed since 2D mean-field models are limited by the fact that they rely on simplified descriptions of complex physical processes such as turbulence and since the cost of 3D models make it difficult, as of today, to provide any reliable predictions concerning the large-scale magnetic cycle. As far as the first kind of numerical simulations is concerned, a particular model has been favoured by a part of the community, namely the Babcock-Leighton flux-transport model first proposed by \\cite{Babcock61} and \\cite{Leighton69}. This model has proved to be very efficient at reproducing several properties of the solar cycle as the equatorward branch of activity or the phase relationship between toroidal and poloidal fields. It has thus been extensively used since the 90's (\\cite{Wang91, Durney95, Dikpati99, Nandy01}), even to make tentative predictions of the next solar cycle (\\cite{Dikpati06}). In this formulation, the poloidal field owes its origin to the tilt of active regions emerging at the Sun's surface at various latitudes during the solar cycle. It thus relies on a different mechanism than ``classical'' $\\alpha\\Omega$ dynamo models for which the poloidal field is generated by turbulent helical motions twisting and shearing toroidal field lines inside the convection zone. The emergence of tilted active regions at the photosphere is the result of the complex non-linear evolution of strong toroidal structures from the base of the convection zone (CZ), where they are created through the $\\Omega$-effect. It is thus natural to rely on 3D MHD simulations of rising toroidal structures in the convection zone to gain some insight on the best way to model the Babcock-Leighton source term. Many models carried out since the 80's relied on the assumption that toroidal flux is organised in the form of discrete flux tubes which will rise cohesively from the base of the CZ up to the solar surface. These models enabled to demonstrate that the initial strength of magnetic field was an important parameter in the evolution of the tube and that the active regions tilts could be explained by the action of the Coriolis force on the magnetic structure (\\cite{DSilva93, Fan94, Caligari95}). More sophisticated multidimensional models were then developed and extended to the upper part of the CZ and the transition to the solar atmosphere (\\cite{Magara04, Archontis05, Cheung07, Martinez08}). Computations were then performed to study the influence of convective turbulent flows on the dynamical evolution of flux ropes inside the CZ (\\cite{Dorch01, Cline03, Fan03}, Jouve $\\&$ Brun 2009). In particular in \\cite{Jouve09}, such computations were made in a global spherical geometry and thus allowed to assess the combined role of hoop stresses, Coriolis force, convective plumes, turbulence, mean flows and sphericity on the tube evolution and on the subsequent emerging regions, along with the usual parameters such as field strength, twist of the field lines or magnetic diffusion. How can the results of 3D simulations of this presumed major step of the dynamo loop be reintroduced in 2D mean-field models? This is the question we are addressing in this work. A particular point we retain from these simulations is that the rise velocity and thus the rise time of magnetic structures is strongly dependent on the initial field strength at the base of the convection zone. Indeed, magnetic buoyancy competes with convective flows to control the trajectory and the speed of the flux tubes while they rise. Since the Babcock-Leighton source term relies on this process to regenerate poloidal fields, it may be worth trying to introduce magnetic-field dependent time delays caused by rising toroidal fields in those models and analyze the influence on the solar cycle. Indeed, time delays built up in solar dynamo models have been shown to cause long-term modulation of the dynamo cycle and under some circumstances to lead to a chaotic behaviour (\\cite{Yoshimura78}, \\cite{Charbonneau052} or \\cite{Wilmot05}). In the framework of mean-field dynamo theory, several possibilities have been studied to explain the variability of the solar cycle. They mainly fell into two categories: stochastic forcing or dynamical nonlinearities. Indeed, as we mentioned above, the solar convection zone is highly turbulent and it would be surprising if the dynamo processes acting inside this turbulent plasma were nicely regular. The influence of stochastic fluctuations in the mean-field dynamo coefficients has been studied in various models (e.g. \\cite{Hoyng88,Ossendrijver96,Weiss00,Charbonneau00}). Moreover, the dynamical feedback of the strong dynamo-generated magnetic fields is likely to be significant enough to produce non-linear effects on the activity cycle. A number of models have introduced these non-linear effects (\\cite{Proctor77, Tobias97, Moss00,Bushby06,Rempel06}) and have resulted in the production of grand minima-like periods or other strong modulation of the cyclic activity. However, time delays have hardly been considered and if they were, they were mainly due to the advection time by meridional flow. Indeed, the time-scale of the buoyant rise of flux tubes was considered to be so small compared to the cycle period (and to the meridional flow turnover time) that this particular step was assumed to be instantaneous. However, we would like to address the question of the influence of magnetic field dependent delays on the cycle produced by Babcock-Leighton models and especially on its potential modulation. We propose to do so in the present paper. This article is organized as follows: Section \\ref{sect_3D} summarizes the results of 3D calculations which will be used in our modified 2D mean-field Babcock-Leighton model. The formulation of this new model is then presented, with a particular focus on the modification of the surface source term. Results of 2D models are shown in Sect. \\ref{sect_2D} and analyzed in the following two sections. Sect. \\ref{sect_6th} and \\ref{sect_5th} present and study the behaviour of a reduced set of ordinary differential equations designed to gain some insight on the results of the 2D model. We discuss the results and conclude in Sect. \\ref{sect_conclu}. The idea of this work is to design a new 2D mean-field flux transport Babcock-Leighton model which takes into account the findings of 3D calculations concerning the rise of strong toroidal structures from the base of the convection zone up to the surface. In this section, we thus present the usual mean-field equations and the basic ingredients used in the model, with a particular focus on the new Babcock-Leighton surface source term. \\subsection{The mean field Babcock-Leighton model} To model the solar global dynamo, we use the hydromagnetic induction equation, governing the evolution of the magnetic field ${\\bf B}$ in response to advection by a flow field ${\\bf V}$ and resistive dissipation. \\begin{equation} \\frac{\\partial {\\bf B}}{\\partial t}=\\nabla\\times ({\\bf V} \\times{\\bf B})-\\nabla\\times(\\eta\\nabla\\times{\\bf B}) \\end{equation} where $\\eta$ is the magnetic diffusivity. As we are working in the framework of mean-field theory, we express both magnetic and velocity fields as a sum of large-scale (that will correspond to mean field) and small-scale (associated with fluid turbulence) contributions. Averaging over some suitably chosen intermediate scale makes it possible to write two distinct induction equations for the mean and the fluctuating parts of the magnetic field. A closure relation is then used to express the electromotive force in terms of mean magnetic field, leading to a simplified mean-field equation. In this work we will replace the emf by a surface Babcock-Leighton term (Babcock 1961; Leighton 1969; Wang et al. 1991; Dikpati \\& Charbonneau 1999; Jouve \\& Brun 2007a) as described in details below. We work here in spherical geometry and under the assumption of axisymmetry. Reintroducing a poloidal/toroidal decomposition of the magnetic and velocity fields in the mean induction equation, we get two coupled partial differential equations, one involving the poloidal potential $A_{\\phi}$ and the other concerning the toroidal field $B_{\\phi}$. \\begin{eqnarray} \\label{eqA2} \\frac{\\partial {A_{\\phi}}}{\\partial t}&=&\\frac{\\eta}{\\eta_{t}} (\\nabla^{2}-\\frac{1}{\\varpi^{2}})A_{\\phi}- R_{e}\\frac{\\bf{v}_{p}}{\\varpi}\\cdot\\nabla(\\varpi A_{\\phi}) \\nonumber \\\\ &+&C_{s}S(r,\\theta,B_{\\phi}) \\end{eqnarray} \\begin{eqnarray} \\label{eqB2} \\frac{\\partial {B_{\\phi}}}{\\partial t}&=&\\frac{\\eta}{\\eta_{t}} (\\nabla^{2}-\\frac{1}{\\varpi^{2}})B_{\\phi} +\\frac{1}{\\varpi}\\frac{\\partial(\\varpi B_{\\phi})}{\\partial r}\\frac{\\partial (\\eta/\\eta_{t})}{\\partial r} \\nonumber \\\\ &-&R_{e}\\varpi {\\bf v}_{p}\\cdot\\nabla(\\frac{B_{\\phi}}{\\varpi})-R_{e}B_{\\phi}\\nabla\\cdot{\\bf v}_{p} \\nonumber \\\\ &+&C_{\\Omega}\\varpi(\\nabla\\times(\\varpi A_{\\phi}{\\bf \\hat{e}}_{\\phi}))\\cdot\\nabla\\Omega \\end{eqnarray} \\noindent where $\\varpi=r\\sin\\theta$, $\\eta_{t}$ is the turbulent magnetic diffusivity (diffusivity in the convective zone), ${\\bf v}_{p}$ the flow in the meridional plane (i.e. the meridional circulation), $\\Omega$ the differential rotation, $S(r,\\theta,B_{\\phi})$ the Babcock-Leighton source term for poloidal field. In order to write these equations in a dimensionless form, we choose as length scale the solar radius ($R_{\\odot}$) and as time scale the diffusion time ($R_{\\odot}^2/\\eta_{t}$) based on the envelope diffusivity ($\\eta_{t}$). This leads to the appearance of three control parameters $C_{\\Omega}=\\Omega_{0}R_{\\odot}^2/\\eta_{t}$, $C_{s}=s_{0}R_{\\odot}/\\eta_{t}$ and $R_{e}=v_{0}R_{\\odot}/\\eta_{t}$ where $\\Omega_{0}, s_{0}, v_{0}$ are respectively the rotation rate and the typical amplitude of the surface source term and of the meridional flow. Equations $\\ref{eqA2}$ and $\\ref{eqB2}$ are solved in an annular meridional cut with the colatitude $\\theta$ $\\in [0,\\pi]$ and the radius (in dimensionless units) $r \\in [0.6,1]$ i.e from slightly below the tachocline ($r=0.7$) up to the solar surface, using the STELEM code. This code has been thoroughly tested and validated thanks to an international mean field dynamo benchmark involving 8 different codes (\\cite{Jouve08}). At $\\theta=0$ and $\\theta=\\pi$ boundaries, both $A_{\\phi}$ and $B_{\\phi}$ are set to 0. Both $A_{\\phi}$ and $B_{\\phi}$ are set to $0$ at $r=0.6$. At the upper boundary, we smoothly match our solution to an external potential field, i.e. we have vacuum for $r \\geq 1$. As initial conditions we are setting a confined dipolar field configuration, i.e the poloidal field is set to $\\sin\\theta / r^{2}$ in the convective zone and to $0$ below the tachocline whereas the toroidal field is set to $0$ everywhere. \\\\ \\subsection{The standard physical ingredients and the modified surface source term} We now need to prescribe the amplitude and profile of the various ingredients acting in equations $\\ref{eqA2}$ and $\\ref{eqB2}$, namely the rotation, the diffusivity, the meridional flow and the Babcock-Leighton source term. The rotation profile captures some realistic aspects of the Sun's angular velocity, deduced from helioseismic inversions (Thompson et al. 2003), assuming a solid rotation below $0.66$ and a differential rotation above the interface. \\begin{eqnarray} {\\Omega(r,\\theta)}&=&\\Omega_{c}+\\frac{1}{2}[1+\\rm erf(2\\frac{r-r_{c}}{d_{1}})] \\nonumber \\\\ & &(\\Omega_{Eq}+a_{2}\\cos^2\\theta+a_{4}\\cos^4\\theta-\\Omega_{c}) \\end{eqnarray} \\noindent with $\\Omega_{Eq}=1$, $\\Omega_{c}= 0.93944$, $r_{c}=0.7$, $d_{1}=0.05$, $a_{2}=-0.136076$ and $a_{4}=-0.145713$ and where $\\rm erf$ is the error function. With this profile, the radial shear is maximal at the tachocline. We assume that the diffusivity in the envelope $\\eta$ is dominated by its turbulent contribution whereas in the stable interior $\\eta_{\\rm c} \\ll \\eta_{\\rm t}$. We smoothly match the two different constant values with an error function which enables us to quickly and continuously transit from $\\eta_{\\rm c}$ to $\\eta_{\\rm t}$ i.e. \\begin{equation} \\eta(r)=\\eta_{\\rm c}+\\frac{1}{2}\\left(\\eta_{\\rm t}-\\eta_{\\rm c}\\right)\\left[1+{\\rm erf}\\left(\\frac{r-r_{\\rm c}}{d}\\right)\\right], \\label{eqeta} \\end{equation} \\noindent with ${\\eta_{\\rm c}}=10^9 \\,\\rm cm^2\\rm s^{-1}$ and $d=0.03$. In Babcock-Leighton flux-transport dynamo models, meridional circulation is used to link the two sources of the magnetic field namely the base of the CZ and the solar surface. For all the models presented in this paper we use a large single cell per hemisphere, directed poleward at the surface, vanishing at the bottom boundary $r=0.6$ and thus penetrating a little below the tachocline. We take a stream function \\begin{equation} \\psi(r,\\theta)=-\\frac{2}{\\pi}\\frac{(r-r_{b})^2}{(1-r_{b})}\\sin\\left(\\pi\\frac{r-r_{b}}{1-r_{b}}\\right)\\cos\\theta\\sin\\theta, \\label{eqpsi} \\end{equation} \\noindent which gives, through the relation ${\\bf u_{\\rm p}}=\\nabla \\times(\\psi \\hat {\\bf e}_{\\phi})$, the following components of the meridional flow \\begin{eqnarray} u_{r}&=&-\\frac{2(1-r_{\\rm b})}{\\pi r}\\frac{(r-r_{\\rm b})^2}{(1-r_{\\rm b})^2} \\sin\\left(\\pi\\frac{r-r_{\\rm b}}{1-r_{\\rm b}}\\right)(3\\cos^2\\theta-1), \\end{eqnarray} \\begin{eqnarray} u_{\\theta}&=&\\Bigg[\\frac{3r-r_{\\rm b}}{1-r_{\\rm b}} \\sin\\left(\\pi\\frac{r-r_{\\rm b}}{1-r_{\\rm b}}\\right)+\\frac{r\\pi}{1-r_{\\rm b}}\\frac{(r-r_{\\rm b})}{(1-r_{\\rm b})} \\cos\\left(\\pi\\frac{r-r_{\\rm b}}{1-r_{\\rm b}}\\right)\\Bigg] \\nonumber \\\\ & \\times &\\frac{2(1-r_{\\rm b})}{\\pi r}\\frac{(r-r_{\\rm b})}{(1-r_{\\rm b})}\\cos\\theta\\sin\\theta, \\end{eqnarray} \\noindent with $r_{\\rm b}=0.6$. In Babcock-Leighton dynamo models, the poloidal field owes its origin to the tilt of magnetic loops emerging at the solar surface. Since these emerging loops are thought to rise from the base of the convection zone through magnetic buoyancy, we see that we can directly relate the way we model the Babcock-Leighton (BL) source term and the results of 3D calculations shown in Sect \\ref{sect_3D}. In the standard model, the source term is confined in a thin layer at the surface and is made to be antisymmetric with respect to the equator, due to the sign of the Coriolis force which changes from one hemisphere to the other. These features are retained in our modified version. However, the standard term is proportional to the toroidal field at the base of the convection zone at the same time, implying an instantaneous rise of the flux tubes from the base to the surface where they create tilted active regions. The 3D calculations showed that the rise velocity and thus the rise time depend on the field strength at the base of the CZ, we thus introduce in our modified version of the source term, a magnetic field-dependent time delay in the toroidal field at the base of the convection zone. We thus take into account the time delay between the formation of strong toroidal structures at the base of the convection zone and the surface regeneration of poloidal field. The modified expression of the source term is thus \\begin{eqnarray} S(r,\\theta,B_{\\phi})&=& \\frac{1}{2}\\Bigg[1+\\rm erf(\\frac{r-r_{2}}{d_{2}})\\Bigg]\\Bigg[ 1-erf(\\frac{r-1}{d_{2}})\\Bigg] \\nonumber \\\\ & \\times & \\Bigg[1+({\\frac{B_{\\phi}(r_{c},\\theta,t-\\tau_B)}{B_{0}}})^{2}\\Bigg]^{-1} \\nonumber \\\\ & \\times &\\cos\\theta \\sin\\theta \\,\\, B_{\\phi}(r_{c},\\theta,t-\\tau_B) \\label{eq_s} \\end{eqnarray} where $r_{2}=0.95$, $d_{2}=0.01$, $B_{0}=10^4 \\, \\rm G$, with the time delay $\\tau_B$ proportional to the inverse of the magnetic energy at the base of the convection zone, namely $$ \\tau_B(\\theta,t) = \\tau_0/B_{\\phi}(r_c,\\theta,t)^2 $$ In our 3D simulations, the approximate rise time for a $6 \\times 10^5 \\, \\rm G$ field is indeed about 4 times that of a $3 \\times 10^5 \\, \\rm G$ field. This takes into account the more significant effects of convective downdrafts and Ohmic diffusion in the ``weak B'' case. Note that the time delay is then dependent both on space and time. Finally, we take into account the fact that very strong toroidal structures (more than $10^5 \\, \\rm G$) are not influenced by the Coriolis force enough to gain a significant tilt as they reach the surface. To do so, we introduce a quenching term in the surface source which will make the regeneration of the poloidal field less effective when the toroidal field at the base of the CZ is strong enough. Inversely, when the flux tubes are too weak, they will not be able to reach the surface at all, and will not take part into the regeneration term for the poloidal field. % We thus prevent the weakest flux tubes (or equivalently the most delayed ones) to take part in the regeneration of poloidal fields at the surface. To do so, the source term is set to zero when the delays are above a certain threshold value corresponding to a rise time of approximately half a solar cycle. ", "conclusions": "\\label{sect_conclu} In this work, we have introduced a more physically accurate model for the poloidal field source term in Babcock-Leighton flux transport dynamo models. To do so, some results from 3D MHD calculations of rising flux tubes from the base of the convection zone to the surface were reintroduced in 2D mean-field models. In particular, the rise time of flux ropes, which was assumed to be negligible in previous simulations, was here taken into account by introducing magnetic energy-dependent time delays in the BL source term. Through the use of a combination of analytical and numerical techniques applied to an adapted reduced system of non linear equations, we show that the system exhibits a sequence of bifurcations leading to a chaotic behaviour. The time evolution of the solution at that stage is qualitatively very similar to the full 2D model and more importantly to the actual solar activity, with strong modulation especially on the cycle amplitude. Assuming the rise of flux tubes from the base of the convection zone and the surface to be instantaneous and thus the time delays due to magnetic buoyancy to be negligible can sound reasonable since these delays are weak compared to the other time-scales of the system. In particular, we can expect the time delay due to the advection by the meridional flow (which has a characteristic time-scale of a few solar cycles, see for example \\cite{Charbonneau00}) to be much more effective on the behaviour of the dynamo-generated magnetic field. As a consequence, it is mainly the effect of meridional flow which has been tested in previous models (e.g. \\cite{Wilmot05}). However, we showed here that the time delays due to flux tubes rising have in fact a significant effect to modulate the cycle on time scales very large compared to the delays. The main reason for this behaviour is that the delays are not fixed but depend on the modulated toroidal energy itself. To illustrate this explanation, we show on Figure \\ref{figure_bq_mod} the time evolution of toroidal fields (delayed and undelayed) for a case similar to what we showed in the core of the paper (a $\\vert B \\vert^2$-dependent delay with $\\tau_0=3.4 \\times 10^{-2}$, i.e. just after the Hopf bifurcation) and the same evolution when the delay is fixed to the mean value reached by the delay in the previous run. \\begin{figure}[h!] \\centering \\includegraphics[width=8.5cm]{./fig_bq_modul.ps} \\includegraphics[width=8.5cm]{./fig_bq_cst.ps} \\caption{Time evolution of the toroidal fields amplitude (undelayed in plain line and delayed in dotted line) for a delay dependent on $\\vert B \\vert^2$ just after the Hopf bifurcation (upper panel) and for a delay fixed to the mean value of the previous run (lower panel).} \\label{figure_bq_mod} \\end{figure} On this figure, both the basic toroidal field and the delayed field are shown, stressing the time shift existing between the two. The striking result is that a long-term modulation appears only in the case of a varying delay. In the fixed-delay case, although the time-shift between the basic toroidal field and its delayed counterpart is quite significant compared to the cycle period, no modulation is created. On the contrary, in the varying-delay case, even though the strongest fields are almost not affected by the time-shift, a strong modulation on the amplitude is built up. As a result, the fact that small delays can affect the long-term evolution of the dynamo cycle seems to be linked to the variability of these delays and more specifically to their dependence on the magnetic field strength. Since magnetic buoyancy is much more efficient in the solar interior for strong toroidal field structures, these kind of delays are likely to appear in reality. Consequently, a new origin of the solar cycle modulation may have been revealed here, thanks to the input of 3D MHD models into 2D mean-field dynamo calculations. Only a part of the results of 3D calculations were reintroduced here but other effects can also be taken into account. In particular, hoop stresses and the Coriolis force are known since the first thin flux tubes simulations to deflect the trajectory of buoyantly rising magnetic fields inside the convection zone, this effect being stronger when the fields are less intense. This feature could also be introduced in BL models by modifying the source term for poloidal field. In particular, the source term is likely to get a contribution from the same latitude a the initial position of the flux tubes when those are strong enough and from a lower latitude when the flux tubes have been sufficiently influenced by the Coriolis force and the hoop stresses to rise parallel to the rotation axis. This may result in a different shape for the butterfly diagram with slightly more activity at higher latitudes, provided that weak magnetic structures are assumed to be able to reach the surface and create active regions. This could lead to reconsider the correspondence which is often made between toroidal magnetic field generated in 2D dynamo models at the base of the convection zone and the actual sunspot migration observed at the solar surface. The variations of the solar cycle period are also known to be significant. For instance, cycle 23 has been significantly longer than the previous ones, with a duration of about 13 years (1996-2009). In the models computed in this work, the modulation of the solar activity is particularly visible on the cycle amplitude and is much less obvious on the period. However, the influence of rising flux tubes on the meridional flow was not taken into account here. Since the cycle period of the solutions resulting from these flux-transport models is very sensitive to the amplitude of the meridional circulation, a modification of the flow by rising flux tubes is very likely to modify the frequency of the dynamo solution and thus to also introduce a modulation on the cycle period. This feature is currently being investigated. A considerable step forward would obviously be to develop a self-consistent global model with buoyant toroidal structures built up and making their way from the base of the convection zone where they become unstable to the photosphere where they create active regions. Unfortunately, this has not been achieved yet due to numerous physical and numerical difficulties. In the mean time, we have shown here that inputs from 3D MHD models simulating a particular step of the dynamo cycle can significantly improve 2D mean-field calculations. In particular, we have shown that they can even help reproducing a feature of the solar cycle (its variability) which was not present in the standard model before. We will continue to develop these ideas in future work." }, "1005/1005.2626_arXiv.txt": { "abstract": "s{ The Extragalactic Background Light (EBL) from the infrared (IR) through the ultraviolet (UV) is dominated by emission from stars, either directly or through absorption and reradiation by dust. It can thus give information on the star formation history of the universe. However, it is difficult to measure directly due to foreground radiation fields from the Galaxy and solar system. Gamma-rays from extragalactic sources at cosmological distances (blazars and gamma-ray bursts) interact with EBL photons creating electron-positron pairs, absorbing the gamma-rays. Given the intrinsic gamma-ray spectrum of a source and its redshift, the EBL can in principle be measured. However, the intrinsic gamma-ray spectra of blazars and GRBs can vary considerably from source to source and the from the same source over short timescales. A maximum intrinsic spectrum can be assumed from theoretical grounds, to give upper limits on the EBL absorption from blazars at low redshift with very high energy (VHE) gamma-ray observations with ground-based Atmospheric Cherenkov telescopes. The Fermi-LAT observations of blazars and GRBs can probe EBL absorption at higher redshifts. The lower energy portion of the LAT spectrum of these sources is unattenuated by the EBL, so that extrapolating this to higher energies can give the maximum intrinsic spectrum for a source. Comparing this to the observed higher energy LAT spectrum will then give upper limits on the EBL absorption. For blazars which have been detected by both the Fermi-LAT and at higher energies by Cherenkov telescopes, combined LAT-VHE observations can put more stringent constraints on the low redshift EBL. These procedures above can also be reversed: for sources with an unknown redshift, a given EBL model and gamma-ray spectrum can lead to an upper limit on the source's redshift. } ", "introduction": "\\label{intro} The night sky appears dark to the naked eye, but in fact glows faintly in the IR through the optical and UV. At these wavelengths, the background light is dominated by emission from the atmosphere, solar system, and Milky Way. There is also a much smaller extragalactic component from all of the stars which have ever existed, through direct emission (in the UV-optical) and through absorption and reradiation by dust (in the IR). Due to the weakness of this extragalactic background light (EBL) to other components, direct measurement of the EBL is extremely difficult~\\cite{bernstein02,mattila03,bernstein07}. The other background components can be avoided to some extent by using instruments on spacecraft which have left the atmosphere~\\cite{bernstein02,hauser98} or the solar system~\\cite{toller83,edelstein00}. However, it is unlikely that spacecraft will leave our Galaxy in the near future, so uncertainties in direct measurements will remain. Number counts in the IR and optical can be used to find EBL lower limits~\\cite{madau00,marsden09}, as discussed by Beelen and Penin in these proceedings. Modeling~\\cite{kneiske04,stecker06,franceschini08,gilmore09,finke10} has been an important tool for constraining the EBL intensity and tying it to basic astrophysics such as the star formation rate density, dust absorption, initial mass function, cosmological expansion rate, and others. Fig.\\ \\ref{EBLmodels_fig} shows many EBL measurements, constraints and models, and Hauser \\& Dwek~\\cite{hauser01} present a thorough review. \\begin{figure} \\begin{center} \\includegraphics[scale=0.34]{EBLmodels_06_stretch} \\caption{EBL models, measurements, and constraints. See Finke {\\em et al.} for details and references. \\label{EBLmodels_fig}} \\end{center} \\end{figure} The EBL photons interact with $\\g$-rays from cosmological sources to produce $e^+e^-$ pairs, absorbing the $\\g$-rays so that the observed flux $F_{obs}(E) = F_{int}(E)\\exp[-\\tau_{\\g\\g}(E)]$ where $F_{int}(E)$ is the unabsorbed source flux as a function of observed energy $E$, and $\\tau_{\\g\\g}(E)$ is the EBL absorption optical depth. If $F_{int}(E)$ is known, a measurement of the observed $\\g$-ray spectrum from these sources can be used to probe the EBL. The intrinsic spectrum is not generally known, however it is possible to determine an upper limit either from theory or from extrapolating a lower energy, unattenuated spectrum to higher energies. This is discussed further in the next sections. From the upper limit on $F_{int}(E)$ and the measurement of $F_{obs}(E)$ with a $\\g$-ray telescope, an upper limit on $\\tau_{\\g\\g}(E)$ can be calculated and compared to theoretical predictions. ", "conclusions": "" }, "1005/1005.4620_arXiv.txt": { "abstract": "{This paper discusses some of the challenges of spectro-polarimetric observations with a large aperture solar telescope such as the ATST or the EST. The observer needs to reach a compromise among spatial and spectral resolution, time cadence, and signal-to-noise ratio, as only three of those four parameters can be pushed to the limit. Tunable filters and grating spectrographs provide a natural compromise as the former are more suitable for high-spatial resolution observations while the latter are a better choice when one needs to work with many wavelengths at full spectral resolution. Given the requirements for the new science targeted by these facilities, it is important that 1)tunable filters have some multi-wavelength capability; and 2)grating spectrographs have some 2D field of view. } ", "introduction": "The Advanced Technology Solar Telescope (ATST), the European Solar Telescope (EST) and the so-called plan B for the space mission Solar-C are all being designed with multi-wavelength spectro-polarimetry as one of the top priorities. This kind of observations will permit novel studies of the processes taking place through the solar atmosphere, from the photosphere to the corona, as they traverse very different physical regimes with conditions that typically vary over several orders of magnitude. Observing the chromosphere/corona at disparate wavelengths with spectro-polarimetry is still challenging even with the large-aperture telescopes currently under development. The science goals dictate requirements that typically include the following: \\begin{itemize} \\item High spatial resolution. We are interested in observing processes that take place at very small spatial scales. Thus, the telescope needs to be able to observe near the diffraction limit. \\item High cadence. The demand for high spatial resolution imposes also requirements on the time cadence. If a particular feature is moving with a projected velocity $v$, integrating over time scales comparable or longer than $t={\\Delta x}/v$ (where $\\Delta x$ is the spatial resolution) will result in image degradation. The sound speed (which is often taken as a first estimate for $v$) in the photosphere is typically around 6~km~s$^{-1}$. However, in the chromosphere it is significantly larger and one often finds supersonic flows. \\item High spectral resolution. Many science goals require not only imaging but also the observation of detailed line profiles to be able to track shock waves in the shape of chromospheric lines, as well as other interesting phenomena that produce emission features or other imprints on the intensity or polarization spectral profiles. Ideally one would want to resolve the Doppler width of the lines. Many interesting chromospheric lines are from H and He ions and have large Doppler widths, but others (particularly Ca) impose requirements for a better resolution. \\item High signal-to-noise ratio. Polarimetric signals are often weak, especially those coming from the chromosphere. Typically, a continuum signal-to-noise ratio of 1000 is required for diagnostics based on the Zeeman effect (some times an order of magnitude better is required for Hanle-effect studies). \\end{itemize} Unfortunately, it is not possible to meet all four of these requirements simultaneously, regardless of the telescope size. Observers will have to make compromises and optimize the instrument configuration for the problem at hand. Compromising will then be a key notion in observational solar physics for the next decade. ", "conclusions": "The tremendous leap in spatial resolution that will be provided by the ATST and the EST will require careful rethinking of most conventional observing modes, particularly long slit spectroscopy. Without at least some limited 2D capability, it is not clear how to make these observations work in practice, particularly with the requirement of seamless multi-wavelength operation that is so vital to the science goals of these facilities. It should be noted, though, that a large aperture is still useful for moderate- or even low-resolution observations in which one needs to reach very high polarimetric sensitivities (e.g., for some Hanle effect observations). Efficient exploitation of the new large solar telescopes with minimum downtimes will be of paramount importance, especially since it now appears that they will concentrate most of the resources devoted by the solar community to observations. This condition implies shifting the telescope operations to a new model more similar to what is currently employed in the large night-time facilities or the space-borne instrumentation such as Hinode. However, there is the risk that the community will be left without a facility where new experimental ideas can be tested. It is extremely important that the new operational model for the ATST/EST considers a provision of at least a small fraction of the observing time to test new ideas or generally speaking, high risk/high return potential campaigns." }, "1005/1005.4599_arXiv.txt": { "abstract": "We consider the interpretation of the MiniBooNE low-energy anomaly and the Gallium radioactive source experiments anomaly in terms of short-baseline electron neutrino disappearance in the framework of 3+1 four-neutrino mixing schemes. The separate fits of MiniBooNE and Gallium data are highly compatible, with close best-fit values of the effective oscillation parameters $\\Delta{m}^2$ and $\\sin^2 2\\vartheta$. The combined fit gives $\\Delta{m}^2 \\gtrsim 0.1 \\, \\text{eV}^2$ and $ 0.11 \\leq \\sin^2 2\\vartheta \\leq 0.48 $ at $2\\sigma$. We consider also the data of the Bugey and Chooz reactor antineutrino oscillation experiments and the limits on the effective electron antineutrino mass in $\\beta$-decay obtained in the Mainz and Troitsk Tritium experiments. The fit of the data of these experiments limits the value of $\\sin^2 2\\vartheta$ below 0.10 at $2\\sigma$. Considering the tension between the neutrino MiniBooNE and Gallium data and the antineutrino reactor and Tritium data as a statistical fluctuation, we perform a combined fit which gives $\\Delta{m}^2 \\simeq 2 \\, \\text{eV}$ and $ 0.01 \\leq \\sin^2 2\\vartheta \\leq 0.13 $ at $2\\sigma$. Assuming a hierarchy of masses $ m_{1}, m_{2}, m_{3} \\ll m_{4} $, the predicted contributions of $m_{4}$ to the effective neutrino masses in $\\beta$-decay and neutrinoless double-$\\beta$-decay are, respectively, between about 0.06 and 0.49 and between about 0.003 and 0.07 eV at $2\\sigma$. We also consider the possibility of reconciling the tension between the neutrino MiniBooNE and Gallium data and the antineutrino reactor and Tritium data with different mixings in the neutrino and antineutrino sectors. We find a $2.6\\sigma$ indication of a mixing angle asymmetry. ", "introduction": "Introduction} Neutrino oscillations have been observed in solar, atmospheric and long-baseline reactor and accelerator experiments. The data of these experiments are well fitted in the framework of three-neutrino mixing, in which the three flavor neutrinos $\\nu_{e}$, $\\nu_{\\mu}$, $\\nu_{\\tau}$ are unitary linear combinations of three massive neutrinos $\\nu_{1}$, $\\nu_{2}$, $\\nu_{3}$ with the solar (SOL) and atmospheric (ATM) squared-mass differences \\begin{align} \\null & \\null \\Delta{m}^2_{21} = \\Delta{m}^2_{\\text{SOL}} \\simeq 8 \\times 10^{-5} \\, \\text{eV}^2 \\,, \\label{002} \\\\ \\null & \\null |\\Delta{m}^2_{31}| \\simeq |\\Delta{m}^2_{32}| = \\Delta{m}^2_{\\text{ATM}} \\simeq 2 \\times 10^{-3} \\, \\text{eV}^2 \\,, \\label{003} \\end{align} where $\\Delta{m}^2_{jk} = m_{j}^2 - m_{k}^2$ and $m_{j}$ is the mass of the neutrino $\\nu_{j}$ (see Refs.~\\cite{hep-ph/9812360,hep-ph/0211462,hep-ph/0310238,hep-ph/0405172,hep-ph/0506083,hep-ph/0606054,GonzalezGarcia:2007ib,Giunti-Kim-2007}). Besides these well-established observations of neutrino oscillations, there are at least three anomalies which could be signals of short-baseline neutrino oscillations generated by a larger squared-mass difference: the LSND $\\bar\\nu_{\\mu}\\to\\bar\\nu_{e}$ signal \\cite{hep-ex/0104049}, the Gallium radioactive source experiments anomaly \\cite{0901.2200,1001.2731}, and the MiniBooNE low-energy anomaly \\cite{0812.2243}. In this paper we consider the MiniBooNE and Gallium anomalies, which can be explained by short-baseline electron neutrino disappearance \\cite{0707.4593,0711.4222,0902.1992} in the effective framework of four-neutrino mixing, as explained in Sections~\\ref{009} and \\ref{019}. On the other hand, the LSND anomaly is disfavored by the results of the MiniBooNE $\\nu_{\\mu}\\to\\nu_{e}$ experiment \\cite{0704.1500,0812.2243} and may require another explanation \\cite{hep-ph/0010308,0705.0107,0706.1462,0710.2985,0805.2098,0906.1997,0906.5072}. In Refs.~\\cite{0707.4593,0902.1992} we proposed to explain the MiniBooNE low-energy anomaly \\cite{0704.1500,0812.2243} through the disappearance of electron neutrinos due to very-short-baseline oscillations into sterile neutrinos generated by a squared-mass difference $ \\Delta{m}^2 $ larger than about $20\\,\\text{eV}^2$. In that case, the analysis of the MiniBooNE data is simplified by the fact that the effective survival probability $P_{\\nu_{e}\\to\\nu_{e}}$ is practically constant in the MiniBooNE energy range from 200 to 3000 MeV. In this paper we extend the analysis of MiniBooNE data to lower values of $ \\Delta{m}^2 $, considering the resulting energy dependence of the effective short-baseline (SBL) electron neutrino and antineutrino survival probability \\begin{equation} P_{\\boss{\\nu}{e}\\to\\boss{\\nu}{e}}^{\\text{SBL}}(L,E) = 1 - \\sin^2 2\\vartheta \\sin^2\\!\\left( \\frac{ \\Delta{m}^2 L }{ 4 E } \\right) \\,, \\label{004} \\end{equation} where $L$ is the neutrino path length and $E$ is the neutrino energy (CPT invariance implies that the survival probabilities of neutrinos and antineutrinos are equal; see Ref.~\\cite{Giunti-Kim-2007}). The two-neutrino-like effective short-baseline survival probability in Eq.~(\\ref{004}) is obtained in four-neutrino schemes (see Refs.~\\cite{hep-ph/9812360,hep-ph/0405172,hep-ph/0606054,GonzalezGarcia:2007ib}), which are the simplest extension of three-neutrino mixing schemes which can accommodate the two small solar and atmospheric squared-mass differences in Eqs.~(\\ref{002}) and (\\ref{003}), and one larger squared-mass difference for short-baseline neutrino oscillations, \\begin{equation} |\\Delta{m}^2_{41}| = \\Delta{m}^2 \\gtrsim 0.1 \\, \\text{eV}^2 \\,. \\label{005} \\end{equation} The existence of a fourth massive neutrino corresponds, in the flavor basis, to the existence of a sterile neutrino $\\nu_{s}$. In this paper we consider 3+1 four-neutrino schemes, since 2+2 four-neutrino schemes are disfavored by the combined constraints on active-sterile transitions in solar and atmospheric neutrino experiments \\cite{hep-ph/0405172}. For simplicity, we consider only 3+1 four-neutrino schemes with \\begin{equation} m_{1}, m_{2}, m_{3} \\ll m_{4} \\,, \\label{006} \\end{equation} which give the $\\Delta{m}^2_{41}$ in Eq.~(\\ref{005}) and appear to be more natural than the other possible 3+1 four-neutrino schemes in which either three neutrinos or all four neutrinos are almost degenerate at a mass scale larger than $\\sqrt{\\Delta{m}^2}$ (see Refs.~\\cite{hep-ph/9812360,hep-ph/0405172,hep-ph/0606054,GonzalezGarcia:2007ib}). In 3+1 four-neutrino schemes the effective mixing angle in the effective short-baseline electron neutrino survival probability in Eq.~(\\ref{004}) is given by (see Refs.~\\cite{hep-ph/9812360,hep-ph/0405172,hep-ph/0606054,GonzalezGarcia:2007ib}) \\begin{equation} \\sin^2 2\\vartheta = 4 |U_{e4}|^2 \\left( 1 - |U_{e4}|^2 \\right) \\,. \\label{007} \\end{equation} In this paper we assume that the value of $|U_{\\mu4}|^2$ is so small that the effective short-baseline muon neutrino survival probability is practically equal to unity and short-baseline $\\boss{\\nu}{\\mu}\\leftrightarrows\\boss{\\nu}{e}$ are negligible\\footnote{ In 3+1 four-neutrino schemes the effective short-baseline muon neutrino survival probability has the form in Eq.~(\\ref{004}) with $\\sin^2 2\\vartheta$ replaced by $\\sin^2 2\\vartheta_{\\mu\\mu} = 4 |U_{\\mu4}|^2 \\left( 1 - |U_{\\mu4}|^2 \\right)$. The effective short-baseline $\\boss{\\nu}{\\mu}\\leftrightarrows\\boss{\\nu}{e}$ transition probability is given by $ P_{\\boss{\\nu}{\\mu}\\leftrightarrows\\boss{\\nu}{e}}^{\\text{SBL}}(L,E) = \\sin^2 2\\vartheta_{e\\mu} \\sin^2\\!\\left( \\frac{ \\Delta{m}^2 L }{ 4 E } \\right) $, with $\\sin^2 2\\vartheta_{e\\mu} = 4 |U_{e4}|^2 |U_{\\mu4}|^2$ (see Refs.~\\cite{hep-ph/9812360,hep-ph/0405172,hep-ph/0606054,GonzalezGarcia:2007ib}). }. This assumption is justified by the lack of any indication of $\\nu_{\\mu}\\to\\nu_{e}$ transitions in the MiniBooNE experiment \\cite{0704.1500,0812.2243} and the limits on short-baseline muon neutrino disappearance found in the CDHSW \\cite{Dydak:1984zq}, CCFR \\cite{Stockdale:1985ce} and MiniBooNE \\cite{0903.2465} experiments. We do not consider the MiniBooNE antineutrino data \\cite{0904.1958}, which have at present statistical uncertainties which are too large to constraint new physics \\cite{0902.1992}. The plan of the paper is as follows. In Section~\\ref{009} we discuss the analysis of MiniBooNE data. In Section~\\ref{019} we present an update of the analysis of Gallium data published in Ref.~\\cite{0711.4222} and the combined analysis of MiniBooNE and Gallium data. In Section~\\ref{027} we discuss the implications of the measurements of the effective electron neutrino mass in Tritium $\\beta$-decay experiments and their combination with reactor neutrino oscillation data. In Section~\\ref{041} we present the results of the combined analysis of MiniBooNE, Gallium, reactor and Tritium data and in Section~\\ref{045} we present the corresponding predictions for the effective masses measured in $\\beta$-decay and neutrinoless double-$\\beta$-decay experiments. In Section~\\ref{058} we calculate the mixing angle asymmetry between the neutrino and antineutrino sectors which could explain the tension between the neutrino and antineutrino data under our short-baseline $\\nu_{e}$-disappearance hypothesis. In Section~\\ref{060} we draw the conclusions. \\begin{figure*} \\begin{center} \\begin{tabular}{lr} \\includegraphics*[bb=33 147 578 695, width=0.48\\linewidth]{fig-01a.eps} & \\includegraphics*[bb=33 147 578 695, width=0.48\\linewidth]{fig-01b.eps} \\end{tabular} \\end{center} \\caption{ \\label{008} Expected number of $\\nu_{e}$ events compared with MiniBooNE data, represented by the black points. The energy bins are numbered with the index $j$. The uncertainty is represented by the vertical error bars, which represent the sum of statistical and uncorrelated systematic uncertainties. (a) Expected number of $\\nu_{e}$-like events $N_{\\nu_{e},j}^{\\text{cal}}$ calculated by the MiniBooNE collaboration. $N_{\\nu_{e},j}^{\\text{cal}}$ is given by the sum of the $\\nu_{e}$-induced events ($N_{\\nu_{e},j}^{\\nu_{e},\\text{cal}}$) and the misidentified $\\nu_{\\mu}$-induced events ($N_{\\nu_{e},j}^{\\nu_{\\mu},\\text{cal}}$). (b) Best-fit value of the number of $\\nu_{e}$-like events $N_{\\nu_{e},j}^{\\text{the}}$ obtained with the hypothesis of $\\nu_{e}$ disappearance. $N_{\\nu_{e},j}^{\\text{the}}$ is given by the sum of $ N_{\\nu_{e},j}^{\\nu_{e},\\text{the}} = f_{\\nu} P_{\\nu_{e}\\to\\nu_{e}}^{(j)} N_{\\nu_{e},j}^{\\nu_{e},\\text{cal}} $ and $ N_{\\nu_{e},j}^{\\nu_{\\mu},\\text{the}} = f_{\\nu} N_{\\nu_{\\mu}}^{\\text{cal}} $. The best-fit values of $f_{\\nu}$, $\\sin^2 2\\vartheta$ and $\\Delta{m}^2$ are those in the first column of Tab.~\\ref{017} (MB$\\nu$). } \\end{figure*} ", "conclusions": "" }, "1005/1005.2189_arXiv.txt": { "abstract": "{We have carried out two extremely deep surveys with SPIRE, one of the two cameras on {\\it Herschel}, at 250\\,$\\mu$m, close to the peak of the far-infrared background. We have used the results to investigate the evolution of the rest-frame 250-$\\mu$m luminosity function out to $z =\\rm 2$. We find evidence for strong evolution out to $z \\simeq\\rm 1$ but evidence for at most weak evolution beyond this redshift. Our results suggest that a significant part of the stars and metals in the universe today were formed at $z \\preceq\\rm 1.4$ in spiral galaxies.} ", "introduction": "The discovery that approximately half the energy ever radiated by galaxies is received on Earth in the far-infrared waveband (Puget et al.\\ 1996; Fixsen et al.\\ 1998) implies that galaxies must show strong evolution that is hidden from optical telescopes (Gispert, Lagache \\& Puget 2000). A decade ago, deep surveys with the SCUBA camera on the James Clerk Maxwell Telescope resolved much of the far-infrared background (FIRB) at 850 $\\mu$m into individual sources (Barger et al.\\ 1998; Hughes et al.\\ 1998). These sources are mostly extremely luminous dust-enshrouded galaxies at $\\rm z > 2$ (Chapman et al.\\ 2005) with an average implied star-formation rate (if the ultimate source of the energy is star formation) of $\\simeq$400 $\\rm M_{\\odot}\\ year^{-1}$ (Coppin et al.\\ 2006), much greater than the star-formation rates in galaxies like our own. However, the energy density in the FIRB at 850 $\\mu$m is $\\simeq$30 times less than at 200 $\\mu$m where the FIRB is at a maximum, and both the spectral shape of the FIRB and statistical `stacking' analyses (Dole et al. 2006; Pascale et al.\\ 2009) imply that much of the FIRB is actually produced by sources at lower redshift (Gispert et al.\\ 2000; Dole et al. 2006; Pascale et al.\\ 2009). The launch of the Herschel Space Observatory (Pilbratt et al.\\ 2010) in May 2009 has given us the opportunity to resolve a significant fraction of the FIRB at wavelengths where its energy density is at a maximum. In this letter, using the first data from the Herschel Multi-tiered Extragalactic Survey (HerMES; Oliver et al., in prep), we investigate the evolution implied by the existence of the FIRB by measuring the evolution of the galaxy luminosity function at 250 $\\mu$m. We everywhere assume a standard concordance cosmology: $\\rm \\Omega_M=0.28,\\ \\Omega_{\\Lambda}=0.72,\\ H_0 = 72\\ km\\ s^{-1}\\ Mpc^{-1}$. ", "conclusions": "On the assumption that the star-formation rate is proportional to the rest-frame 250-$\\mu$m luminosity, the lack of evidence for any strong evolution at $\\rm z > 1$ is consistent with investigations that have concluded that the overall star-formation rate per unit comoving volume was approximately constant at $\\rm z > 1$ (Steidel et al.\\ 1999; Gispert et al.\\ 2000; Hopkins 2004). We estimate from the GOODS-North catalogue that we have resolved $\\simeq$20\\% of the FIRB at 250 $\\mu$m. Fig.~3 shows that the sources making up this top 20\\% of the FIRB are at moderate redshift ($z \\sim 1$). Although we cannot say anything directly about the sources responsible for the missing 80\\% of the FIRB, the stacking analysis of Pascale et al.\\ (2009) from BLAST data suggests that the remainder of the FIRB at this wavelength is also from sources at moderate redshift. A revealing way to look at these early HerMES results is to use the relationship between the production of metals and the background radiation associated with this metal production (Peacock 1993). Using a value for the integrated FIRB of $\\rm 14\\ nW\\ sr^{-1}$ (Fixsen et al.\\ 1998), we obtain a relationship between the mass of metals and the fraction, $\\epsilon$, of the FIRB produced in a particular redshift interval: \\smallskip $$ M_{metals} = 9 \\times 10^6 \\epsilon (1+z)\\ M_{\\odot}\\ Mpc^{-3}. $$ \\smallskip \\noindent Fig.~5 shows the metals produced as a function of redshift on the assumption that the redshift distributions of GOODS-North and LH-North are representative of the FIRB as a whole. The figure suggests that most of the metals (and therefore most of the massive stars) formed at a moderate redshift ($z \\preceq 1.4$). This conclusion is consistent with the overall star-formation rate in the universe being constant at $\\rm z > 1$, and the apparent decline of metal production at high redshift is simply the consequence of the relationship between cosmic time and redshift. However, the decline in the figure at high redshifts is relatively small, and since we are making the large assumptions that the redshift distributions of our samples are characteristic of the FIRB as a whole, it is possible that observations that resolve more of the FIRB and at several wavelengths rather than a single wavelength will modify this conclusion. The excellent images that exist from the Hubble Space Telescope for the GOODS-North {\\it Herschel} sources mean that we can examine the nature of the galaxies responsible for this metal production. We have used the z-band images, since at $\\rm z \\simeq 1$ this band corresponds in the rest-frame to the B-band. Of the 26 galaxies in the redshift interval $\\rm 0.8 < z < 1.2$, 12 show clear signs of spiral structure, and there may well be more if some of the galaxies have spiral arms that are below the surface-brightness detection threshold of the images. Thus the galaxies found in the first HerMES images appear to be quite different from the major mergers found in the SCUBA surveys (Ivison et al.\\ 2000; Tacconi et al.\\ 2008). Studies of galaxy evolution based on {\\it Spitzer} 24-$\\mu$m surveys have also found that the {\\it Spitzer} galaxies at $\\rm z \\simeq 1$ are often spirals rather than major mergers (Elbaz et al.\\ 2007) but the crucial advance that {\\it Herschel} has made possible is that we can determine the cosmological importance of these objects, in terms of the total masses of stars and metals formed in them. \\begin{figure} \\centering \\includegraphics[width=6cm]{14675fg5.eps} \\caption{An estimate of the total metals formed per cubic Mpc in different redshift intervals using the method described in the text. We have made the estimates from the redshift distributions of GOODS-North (red) and LH-North (green) on the assumption that these redshift distributions represent the FIRB as a whole.} \\end{figure}" }, "1005/1005.2971_arXiv.txt": { "abstract": "In addition to the Sun, six other stars are known to harbor multiple planets and debris disks: HD 69830, HD 38529, HD 128311, HD 202206, HD 82943 and HR 8799. In this paper we set constraints on the location of the dust-producing planetesimals around the latter four systems. We use a radiative transfer model to analyze the spectral energy distributions of the dust disks (including two new $Spitzer$ $IRS$ spectra presented in this paper), and a dynamical model to assess the long-term stability of the planetesimals' orbits. As members of a small group of stars that show evidence of harboring a multiple planets and planetesimals, their study can help us learn about the diversity of planetary systems. ", "introduction": "Surveys with the {\\it Spitzer} Space Telescope have been spectacularly successful at identifying infrared excess emission associated with planetary debris disks around A- through K-type stars. The excesses at 70 $\\mu$m are associated with cool dust located at distances from the stars analogous to the position of the Kuiper Belt (KB) in the solar system (Moro-Mart\\'in et al. 2008), although much larger numbers of objects must lie in these exo-Kuiper belts to account for the detected level of emission (Trilling et al. 2008; Hillenbrand et al. 2008; Carpenter et al. 2009). The most sensitive studies find 70 $\\mu$m excesses around $\\sim$ 15-20\\% of mature solar-type stars over the entire 10 Myr to 10 Gyr age range (Trilling et al. 2008). The excesses at 24$\\mu$m are generally associated with warmer dust and disappear relatively quickly as the host star ages; about 30\\% of the stars of the Pleiades age ($\\sim$ 120 Myr) show excess emission (Sierchio et al. 2010) whereas by the age of Praesepe ($\\sim$ 600 Myr), the 24 $\\mu$m excesses have almost completely disappeared (Meyer et al. 2008; G\\'asp\\'ar et al. 2009). Because the expected lifetimes of the debris dust grains are much shorter than the ages of the stars, it is inferred that the dust originates from collisional activity in reservoirs of planetesimals left over from the planet formation process (hence the term {\\it debris} dust). To sustain the dust production, it is necessary that large planetesimals (1000 km-sized) or unseen planets stir the planetesimals so they continue to collide with each other. Planets are also responsible for constraining the planetesimals in some zones and clearing them from others, thus determining much of the structure of the debris system. The way the patterns of debris disks activity decay with age is consistent with the expectation that the inner zones of a planetary system have relatively short dynamical time scales, whereas dynamical activity unfolds slowly at the distance of the Kuiper Belt. A highlight of the recent surveys is the first detection of debris disks around stars with planets (Beichman et al. 2005; Moro-Mart\\'in et al. 2007a). These planets orbit within several AU of their parent star, whereas the cold dust emitting the observed far-IR radiation generally resides tens of AU away. Despite the separation between the dust and planets, it is still possible for the planet to shape the structure of the dust (and planetesimals) disk. In the multiple-planet system HD 38529, for example, secular resonances excited by planets at 0.13 and 3.7 AU create regions at tens of AUs that are unstable for orbiting planetesimals (Moro-Mart\\'in et al. 2007b). Several additional systems have been identified as having both multiple planets (capable of exciting secular resonances) and orbiting debris (indicating the presence of planetesimals; Bryden et al. 2009, Su et al. 2009). Presumably there are many more stars with debris-disk excesses that also harbor multiple but undiscovered planets. Detailed studies of the known examples can reveal aspects of their behavior that help us understand the diversity of planetary systems. In this paper we study four of these systems: HD 128311, HD 202206, HD 82943 and HR 8799. In Section \\ref{obs} we describe the planet and debris dust detections for each one of these systems, and present new $Spitzer$ $IRS$ detections for the debris disks around HD 202206 and HD 82943. In Section \\ref{sed} we use a radiative transfer model to identify the range of parameters (dust mass and dust location) that would fit the observed spectral energy distribution (SED). Due to the high fractional luminosity, the grain-grain collisional time-scale is shorter than the Poynting-Robertson (P-R) time-scale for all these systems and therefore we expect the dust to trace the location of the dust-producing planetesimals. In Section \\ref{dyn}, we use a dynamical model to assess the long-term orbital stability of the putative dust-producing planetesimals, taking into account the effect of secular resonances. Putting together the results from the SED and dynamical analysis, in Section \\ref{loc} we discuss the potential location of the dust-producing planetesimal belts. ", "conclusions": "In this paper we have studied the possible planet-planetesimal configurations of four multi-planet systems, of which three are radial-velocity systems -- HD 128311, HD 202206, HD 82943 -- and one is a directly imaged system -- HR 8799. We have quantified where the zone of influence lies of planets on the dust-producing planetesimals: for HR 8799 it extends to nearly 20 AU from the orbit of the outermost planet, while for the three radial-velocity systems it extends to about 4 AU. A previous paper that studied HD38529, another radial-velocity multi-planet system, found that the influence of the planets in this case extends out to $\\sim$ 10 AU (determining the inner edge of the disk), and becomes dominant again at a $\\sim$ 55 AU (due to a secular resonance that probably determines the outer edge of the dust disk (Moro-Mart\\'in et al. 2007b). We conclude that radial-velocity multi-planet systems generally have zones of influence within a few to ten AU; more precise determinations will require individual modeling of a system. The influence can be extended much further through secular resonances and similar behavior. For the three radial-velocity multi-planet systems studied in this paper, we have constructed fits to the spectral energy distributions of the debris disks. If we use astronomical silicates with a size distribution down to the blow-out size, the emitting regions of the disks must be so far from the star that they are well outside the zones of influence of the radial velocity planets. This behavior depends critically on the optical properties of the grains (size distribution and optical constants of the grain material). If we were to adopt optical constants typical of an ice-silicate mixture instead of astronomical silicates, the dust may be located closer to the star. In all three cases, we find that single-size 10 $\\mu$m astronomical silicates reproduce well the observed SED, and could lie at the edge of the zone of influence of the planets and therefore the disk would be sculpted by them. \\begin{center} {\\it Acknowledgments} \\end{center} We thank Hal Levison for providing skeel-SyMBA for the dynamical simulations, Sebastian Wolf for providing DDS for the SED models, Alexander Krivov for his careful reading of the manuscript and useful comments and Daniel Fabrycky and Ruth Murray-Clay for useful comments. This work is based on observations made with the {\\it Spitzer} Space Telescope, which is operated by the Jet Propulsion Laboratory, managed for NASA by the California Institute of Technology. A.M.M. acknowledges funding from the Spanish MICINN (Ram\\'on y Cajal Program and grants AYA2009-07304 and CONSOLIDER INGENIO 2010CSD2009-00038), the Michelson Fellowship and the {\\it Spitzer} archival grant 40412. She thanks the Isaac Newton Institute for Mathematical Sciences at Cambridge University for support. RM acknowledges support from grants by NSF (AST-0806828) and NASA (NNX08AQ65G). \\newpage" }, "1005/1005.0373_arXiv.txt": { "abstract": "{We investigate the possibility of using the ratio between the 2-10 keV flux and the [Ne~V]3426 emission line flux (X/NeV) as a diagnostic diagram to discover heavily obscured, possibly Compton-Thick Active Galactic Nuclei (AGN) in the distant Universe. While being on average about one order of magnitude fainter than the more commonly used [O~III]5007 emission line, the [Ne~V]3426 line can be observed with optical spectroscopy up to $z\\sim 1.5$, whereas the [O III]5007 line is redshifted out of the optical bands already at $z\\sim0.8$. First, we calibrate a relation between X/NeV and the cold absorbing column density $N_H$ using a sample of 74 bright, nearby Seyferts with both X-ray and [Ne V] data available in the literature, and for which the column density is determined unambiguously. Similarly to what is found for the X-ray to [O III]5007 flux ratio (X/OIII), we found that the X/NeV ratio decreases towards large column densities, as expected if [Ne V]3426 emission is a good tracer of the AGN intrinsic power. Essentially all local Seyferts with X/NeV values below 15 are found to be Compton-Thick objects. At X/NeV values below 100, the percentage of Compton-Thick nuclei decreases to $\\sim 50\\%$, but still $\\sim 80\\%$ of the considered sample is absorbed with $N_H>10^{23}$ cm$^{-2}$. Second, we apply this diagnostic diagram to different samples of distant obscured and unobscured QSOs in the Sloan Digital Sky Survey (SDSS). SDSS blue, unobscured, type-1 QSOs in the redshift range $z=[0.1-1.5]$ indeed show X/NeV values typical of unobscured Seyfert 1s in the local Universe. Conversely, SDSS type-2 QSOs at $z\\sim 0.5$ classified either as Compton-Thick or Compton-Thin on the basis of their X/OIII ratio, would have been mostly classified in the same way based on the X/NeV ratio. We apply the X/NeV diagnostic diagram to 9 SDSS obscured QSOs in the redshift range $z=[0.85-1.31]$, selected by means of their prominent [Ne~V]3426 line (rest $EW>4$\\AA) and observed with \\chandra\\ ACIS-S for 10ks each (8 of them as part of our proprietary program). Based on the X/NeV ratio, complemented by X-ray spectral analysis, 2 objects appear good Compton-Thick QSO candidates, 4 objects appear as Compton-Thin QSOs, while 3 have an ambiguous classification. When excluding from the sample broad lined QSOs with a red continuum and thus considering only genuine narrow-line objects, the efficiency in selecting Compton-Thick QSOs through the [Ne~V] line is about 50\\% (with large errors, though), more similar to what is achieved with [O~III] selection. We discuss the possibility of applying the X/NeV diagnostic to deep X-ray surveys to search for Compton-Thick Seyferts at $z\\sim 1$, i.e. those objects which are thought to be responsible for the ``missing\" X-ray background. Finally, we compare the optical spectral properties of [Ne~V]-selected QSOs with those of other SDSS populations of obscured and unobscured QSOs. By restricting the analysis to objects in the same redshift (and luminosity) range z=[0.4-1.5], we found evidence that, at any given [Ne~V] luminosity, increasing obscuration is accompanied by increasing [O~II]3727 emission. This correlation is interpreted as evidence for enhanced star formation in obscured QSOs, which is consistent with current popular scenarios of BH-galaxy coevolution. ", "introduction": "\\label{introduction} While the cosmological evolution of unobscured QSOs has been traced up to $z\\sim 6$, the evolution of obscured AGN is much more uncertain and is the subject of intense debate. The {\\it observed} number statistics in current AGN samples is dominated by unobscured objects which are easier to discover (e.g. $>10^5$ QSOs have been identified in the Sloan Digital Sky Survey), but several arguments suggest that obscured AGN must be intrinsically more numerous. Deep X-ray surveys (see \\citealt{bh05} for a review) have indeed shown that, towards faint X-ray fluxes, the surface density of obscured AGN overtakes that of unobscured AGN. Also, population synthesis models of the cosmic X-ray background (XRB), suggest that obscured AGN outnumber unobscured ones by a factor which ranges from $\\sim 2$ to $\\sim 8$, depending on the considered luminosity regime (\\citealt{gch07}; see \\citealt{tuv09} and \\citealt{balla06} for a steeper luminosity dependence). To understand the cosmological history of accretion onto supermassive black holes (SMBHs) it is therefore necessary to map and understand the population of obscured AGN and, in particular, of the most obscured and hence elusive ones, the so-called Compton-Thick (CT) nuclei, i.e. those obscured by column densities above $\\sim 10^{24}$ cm$^{-2}$. The population of moderately obscured AGN in fact does not completely account for the XRB peak intensity at 30 keV, to which CT AGN are expected to contribute from $\\sim 10\\%$ \\citep{tuv09} to $\\sim 25-30\\%$ \\citep{gch07}, depending on the XRB model assumptions. In addition, the presence of a large population of CT AGN across the cosmic epochs, would help in reconciling the measured mass function of local SMBHs with that expected by integrating the accretion history of seed black holes \\citep{marconi04, shankar04}. Finally, popular semi-analytic models of galaxy formation and evolution \\citep{kh00, marulli08}, coupled to hydrodynamical simulations of galaxy mergers \\citep{hop06}, propose that nuclear activity is triggered during major mergers of gas-rich galaxies and that at its early stage, the AGN is embedded within optically thick gas shrouds. Despite these theoretical progresses, the cosmological evolution and luminosity function of CT objects is currently unknown, and in the synthesis models of the XRB it has been usually assumed to be equal to that of less obscured objects. Whether this is the case or not can only be determined by obtaining statistically significant samples of distant CT objects and by comparing them with the local samples (see e.g. \\citealt{c04} and \\citealt{rdc08} for reviews on nearby, {\\it bona-fide} CT objects). The number of techniques devised to select CT AGN is rapidly growing, following the technological development of efficient detectors across the electromagnetic spectrum. These diverse selection techniques have allowed the first estimates of the space density of CT AGN in different redshift and luminosity regimes. Very hard X-ray selection, i.e. at energies above 10 keV, is unaffected by absorption up to a few $\\times 10^{24}$ cm$^{-2}$, but, because of the still limited instrumental sensitivity, is mainly sampling the nearby Universe. The population of CT AGN detected by INTEGRAL/IBIS \\citep{tueller08} and Swift/BAT \\citep{malizia09} at $z<0.02$ is indeed producing only a tiny fraction ($<$1\\%) of the cosmic XRB. Deep X-ray surveys in the more accessible 2-10 keV band are revealing large populations of heavily obscured objects in the redshift range $z\\sim[0.5-4]$. The generally low photon statistics of the detected sources, however, prevent an accurate spectral analysis and $N_H$ measurement. The CT nature of a faint X-ray source is therefore often inferred from the characteristics of the reprocessed spectrum, like the presence of a prominent fluorescence $K\\alpha$ iron line over a flat continuum. Examples of distant CT AGN selected with this technique have been found by \\citet{tozzi06} and \\citet{geo07,geo09} in the \\chandra\\ Deep Fields. The number of CT AGN candidates detected in the deep X-ray surveys appears to be in rather good agreement with the XRB models predictions. A very recent and promising approach to select CT candidates in the distant Universe is based on their strong mid-IR flux, where most of the absorbed radiation should be re-emitted \\citep{alejo05}. Recently, \\citet{daddi07}, \\citet{fiore08, fiore09} and \\citet{alex08} located heavily obscured AGN in objects showing 24$\\mu$m emission in excess of that expected from dust heated by stellar processes. By stacking the \\chandra\\ data of these mid-IR-excess objects, a very hard X-ray spectrum was observed, reminiscent of CT obscuration. These studies span a broad AGN luminosity range ($L_X\\sim10^{42-45}$ erg s$^{-1}$), but mostly sample populations of objects at $z\\sim 2$. The measured space density of CT AGN at these high redshifts is in general as large as expected from XRB synthesis models or possibly even larger (see eg. \\citealt {trei09_ct}). Another way to select obscured QSOs is through their high-ionization narrow optical emission lines, which are thought to be produced on physical scales (from $\\sim0.1$ to a few kpc) mostly free from nuclear obscuration. Recently, the [O IV]26$\\mu m$ line has been used to select obscured AGN among galaxies observed with {\\it Spitzer}/IRS. However, since this line quickly moves out of the observable IR bands as redshift increases, this selection mostly concerns the nearby Universe \\citep{diamond09, rigby09}. The most commonly used marker of obscured nuclear activity therefore remains the [O~III]5007 emission line, which is strong, falls in the optical domain, and allows object selection up to $z\\sim 0.8$. The 2-10 keV to [O~III]5007 flux ratio (X/OIII) has been often used as a diagnostic for heavy obscuration in sources with poor X-ray photon statistics \\citep{maio98, cappi06, panessa06}, being low X/OIII ratios ($\\lesssim 3$, see e.g. Fig.~4 of \\citealt{cappi06}) highly suggestive of heavy nuclear absorption. Based on the [O~III]5007 emission line, \\citet{zak03} and \\citet{reyes08} identified in the Sloan Digital Sky Survey (SDSS) a population of obscured QSOs at a median redshift of $z\\sim0.3$, at least as abundant as that of type-1 QSOs at the same redshifts \\citep{reyes08}. These results have been extended to lower luminosities by \\citet{bongiorno09}, who measured the luminosity function of [O III]-selected type-2 AGN in the zCOSMOS spectroscopic survey \\citep{lilly07}, finding that the fraction of obscured AGN is decreasing with luminosity, in agreement with what is observed in X-ray surveys. X-ray observations of small samples drawn from the \\citet{zak03} catalog, suggest that about half of luminous type-2 QSOs (log$L_{O III}>9.3\\;L_{\\odot}$) could be CT \\citep[hereafter V10; see also \\citealt{lamastra09} for X-ray observations of lower luminosity SDSS type-2 AGN]{ptak06, v06, v10}. Since [O III] selection is likely missing objects in which also the Narrow Line Region (NLR) is extincted (like e.g. in the prototype CT AGN NGC~4945 and NGC~6240), the estimated type-2 QSO abundances should be considered as lower limits \\citep[V10]{reyes08}. In this work we explore the possibility of using the high-ionization [Ne~V]3426 emission line, rather than the [O~III]5007 line, as a tracer of obscured nuclear activity. Despite being on average a factor of $\\sim 9$ weaker than [O~III]5007 \\citep{fo86, zak03} and suffering stronger dust extinction, the [Ne~V]3426 line is commonly observed in nearby Seyfert galaxies and, given that high energy photons ($\\gtrsim 0.1$ keV) are required to further ionize NeIV, it is considered an unambiguous sign of nuclear activity \\citep[e.g.][]{schmidt98}. In addition, the [Ne~V]3426 emission line is observable up to $z\\sim 1.5$ before being redshifted out of the optical bands, whereas the [O~III]5007 line is observable only up to $z\\sim 0.7-0.8$. Indeed, only 13 out of 887 [O~III]-selected QSOs in the \\citet{reyes08} sample lie at $z>0.7$, with only 3 at $z>0.8$. [Ne~V]-selection may then be used to reveal nuclear activity in obscured sources at $z\\sim 1$, i.e. at the epoch where most of the XRB light is thought to be produced. The structure of the paper is the following: in Section 2 we present and discuss the sample of nearby Seyfert galaxies used to calibrate the relation between [Ne~V] and X-ray emission (the details of the sample are given in the Appendix). In Section 3 we present the X/NeV diagnostic diagram and apply it to obscured and unobscured QSO population drawn from the SDSS. In Section 4 we present \\chandra\\ observations of a sample of 9 [Ne~V]-selected obscured QSOs at $z\\sim 1$ in the SDSS and use the X/NeV diagnostic to estimate the fraction of CT objects among them. In Section 5 we discuss efficiency and biases of [Ne~V] selection together with its application to sky areas with deep optical spectroscopy and X-ray coverage. In the same Section, the evidence of enhanced star formation in obscured QSOs at z=0.4-1.5 is also highlighted. The conclusions are drawn in Section 6. \\begin{figure*}[t] \\begin{center} \\includegraphics[width=12cm]{4039f1.ps} \\caption{Observed 2-10 keV to [Ne~V]3426 luminosity ratio (X/NeV) vs absorption column density for a sample of 74 Seyfert galaxies in the local Universe. The solid line shows the expected trend obtained by starting from the mean X/NeV ratio $\\langle$X/NeV$\\rangle$ observed in unobscured objects (i.e. those plotted at log$N_H=20$) and progressively obscuring the X-ray emission with increasing $N_H$ (up to log$N_H=25.5$) as plotted on the y-axis. Lower limits at log$N_H=24$ refer to Compton-Thick objects observed only below 10 keV, while datapoints at log$N_H>24$ and lower limits at log$N_H=25$ refer to CT objects observed also above 10 keV, for which a more stringent determination of the column density is possible. The darker (lighter) shaded region is obtained by making the same computation as above but starting at $\\pm1\\sigma$ ($\\pm 90\\%$) around $ \\langle$X/NeV$\\rangle$ (see text for details). The region at low ($<15$) X/NeV values is essentially populated only by CT objects.} \\label{xnev_local} \\end{center} \\end{figure*} ", "conclusions": "We have presented a diagnostic diagram to identify heavily obscured, Compton-Thick AGN candidates at $z\\sim 1$ based on the ratio between the 2-10 keV flux and the [Ne~V]3426 emission line flux (X/NeV). The diagnostic was calibrated on a sample of 74 local Seyfert galaxies and then applied to populations of type-1 and type-2 QSOs at different redshifts (from $z\\sim 0.1$ to $z=1.5$) selected from the SDSS. The main results obtained in this work can be summarized as follows. \\\\ \\noindent $\\bullet$ The observed X/NeV ratio is found to decrease with increasing absorption: the mean X/NeV ratio for unobscured Seyferts is about 400, about 80\\% of local Seyferts with X/NeV$<100$ are obscured by column densities above $10^{23}$\\cm\\, and essentially all objects with observed X/NeV $<15$ are Compton-Thick. \\\\ \\noindent $\\bullet$ We considered a sample of 83 blue type-1 QSOs and 21 [O III]-selected type-2 QSOs in the SDSS which have been observed in the X-rays and show significant [Ne V] detection. It was verified that they follow the same X/NeV vs X-ray absorption trend which is observed for local Seyferts. Furthermore, SDSS type-2 QSOs classified either as Compton-Thick or Compton-Thin on the basis of their X/OIII ratio, would have been mostly classified in the same way based on the X/NeV ratio. \\\\ \\noindent $\\bullet$ The X/NeV diagnostic was used to investigate the obscuration of 9 SDSS obscured QSOs in the redshift range $z=[0.85-1.31]$, which is not accessible through [O III] selection. The 9 objects were selected by means of their prominent [Ne~V]3426 line ($EW>4$\\AA), and \\chandra\\ snapshot observations for 8 of them were obtained (one object is from the archive). Based on the X/NeV ratio, complemented by X-ray spectral analysis, only 2 objects appear good Compton-Thick QSO candidates. However, when considering the 4 genuine narrow-line objects only (FWHM of the MgII line $\\lesssim 2000$ km $s^{-1}$), the efficiency in selecting Compton-Thick QSOs through the [Ne~V] line is about 50\\% (2/4), which is more similar, despite the large uncertainties, to what is achieved with [O~III] selection (60-70\\%; \\citealt{v10}). \\\\ \\noindent $\\bullet$ We verified that neither extinction nor anisotropy corrections on the [Ne~V] emission would affect our conclusions and that the X/NeV diagnostic is therefore a good method to identify clean, despite not complete, samples of heavily obscured AGN. We discussed the possibility of applying the X/NeV diagnostic to objects in sky areas with deep optical spectroscopy and X-ray coverage. This will allow to identify Compton-Thick Seyferts at $z\\sim 1$, i.e. those objects which are thought to be responsible for a large fraction of the ``missing\" X-ray background. \\\\ \\noindent $\\bullet$ Finally, the optical emission line properties of [Ne~V]-selected QSOs were compared with those of other SDSS populations of obscured and unobscured QSOs. By restricting the analysis to objects in the same redshift (and luminosity) range $z$=[0.4-1.5], we found evidence that the ratio between the [O~II]3727 and [Ne~V]3426 luminosity increases with obscuration. This correlation is interpreted as evidence of enhanced star formation in obscured QSOs, which is consistent with current popular scenarios of BH-galaxy coevolution. {" }, "1005/1005.1929_arXiv.txt": { "abstract": "Precise measurement of the angular power spectrum of the Cosmic Microwave Background (CMB) temperature and polarization anisotropy can tightly constrain many cosmological models and parameters. However, accurate measurements can only be realized in practice provided all major systematic effects have been taken into account. Beam asymmetry, coupled with the scan strategy, is a major source of systematic error in scanning CMB experiments such as Planck, the focus of our current interest. We envision Monte Carlo methods to rigorously study and account for the systematic effect of beams in CMB analysis. Toward that goal, we have developed a fast pixel space convolution method that can simulate sky maps observed by a scanning instrument, taking into account real beam shapes and scan strategy. The essence is to pre-compute the ``effective beams'' using a computer code, ``Fast Effective Beam Convolution in Pixel space'' ({\\tt FEBeCoP}), that we have developed for the Planck mission. The code computes effective beams given the focal plane beam characteristics of the Planck instrument and the full history of actual satellite pointing, and performs very fast convolution of sky signals using the effective beams. In this paper, we describe the algorithm and the computational scheme that has been implemented. We also outline a few applications of the effective beams in the precision analysis of Planck data, for characterizing the CMB anisotropy and for detecting and measuring properties of point sources. ", "introduction": " ", "conclusions": "" }, "1005/1005.3092_arXiv.txt": { "abstract": "{Type Ia supernovae (SNe Ia) play a key role in measuring cosmological parameters, in which the Phillips relation is adopted. However, the origin of the relation is still unclear. Several parameters are suggested, e.g. the relative content of carbon to oxygen (C/O) and the central density of the white dwarf (WD) at ignition. These parameters are mainly determined by the WD's initial mass and its cooling time, respectively. Using the progenitor model developed by Meng \\& Yang, we present the distributions of the initial WD mass and the cooling time. We do not find any correlation between these parameters. However, we notice that the range of the WD's mass decreases, while its average value increases with the cooling time. These results could provide a constraint when simulating the SN Ia explosion, i.e. the WDs with a high C/O ratio usually have a lower central density at ignition, while those having the highest central density at ignition generally have a lower C/O ratio. The cooling time is mainly determined by the evolutionary age of secondaries, and the scatter of the cooling time decreases with the evolutionary age. Our results may indicate that WDs with a long cooling time have more uniform properties than those with a short cooling time, which may be helpful to explain why SNe Ia in elliptical galaxies have a more uniform maximum luminosity than those in spiral galaxies. ", "introduction": "% \\label{sect:1} As one of the most widely used distance indicators, type Ia supernovae (SNe Ia) show their importance in determining cosmological parameters, which resulted in the discovery of the accelerating expansion of the universe (\\citealt{RIE98}; \\citealt{PER99}). The result was exciting and suggested the presence of dark energy. At present, SNe Ia are proposed to be cosmological probes for testing the evolution of the dark energy equation of state with time and testing the evolutionary history of the universe (\\citealt{RIESS07}; \\citealt{KUZNETSOVA08}; \\citealt{HOWEL09}). They were even chosen to check the consistency of general relativity (\\citealt{ZHAOGB10}). When SNe Ia are applied as a distance indicator, the Phillips relation is adopted, which is a linear relation between the absolute magnitude of SNe Ia at maximum light and the magnitude drop of the B light curve during the first 15 days following the maximum (\\citealt{PHI93}). This relation implies that the brightness of SNe Ia is mainly determined by one parameter. It is generally agreed that the amount of $^{\\rm 56}$Ni formed during the supernova explosion dominates the maximum luminosity of SNe Ia (\\citealt{ARN82}), but the origin of the different amount of $^{\\rm 56}$Ni for different SNe Ia is still unclear (\\citealt{POD08}). Some numerical and synthetical results showed that metallicity has an effect on the final amount of $^{\\rm 56}$Ni, and thus the maximum luminosity (\\citealt{TIM03}; \\citealt{TRA05}; \\citealt{POD06}; \\citealt{BRAVO10}) and there do be some observational evidence of the correlation between the properties of SNe Ia and metallicity (\\citealt{BB93}; \\citealt{HAM96}; \\citealt{WAN97}; \\citealt{CAP97}; \\citealt{SHA02}). However, the metallicity seems not to have the ability to interpret the scatter of the maximum luminosity of SNe Ia (\\citealt{TIM03}; \\citealt{GALLAGHER08}; \\citealt{HOWEL09b}). \\citet{NOM99, NOM03} suggested that the ratio of carbon to oxygen (C/O) of a white dwarf at the moment of explosion is the dominant parameter for the Phillips relation. The higher the C/O, the larger the amount of nickel-56, and then the higher the maximum luminosity of SNe Ia. The C/O ratio is a function of the initial mass of the WD, which then is related to the progenitor system of the SNe Ia. By comparing theory and observations, the results of \\citet{MENGXC09} and \\citet{MENGXC10a} upheld this suggestion. \\citet{LESAFFRE06} carried out a systematic study of the sensitivity of ignition conditions for H-rich Chandra single degenerate exploders on various properties of the progenitors, and suggested that the central density of the WD at ignition may be the origin of the Phillips relation (see also \\citealt{POD08}). These authors noticed that the more massive and/or the cooler the CO WD is when accretion begins, the higher the central density is at ignition. The central density is then also related to the progenitor system. When one simulates the explosion of SNe Ia, the C/O and the central density are always set to be free parameters (\\citealt{ROPKE06}). It is thus interesting to analyze whether there is a correlation between the C/O and the central density. In addition, analyzing how the initial mass of the CO WD and its cooling time vary with the delay time is also an interesting task (\\citealt{GREGGIO10}). The purpose of this paper is to check these interesting problems. In section \\ref{sect:2}, we describe our model. We show the results in section \\ref{sect:3} and give discussions and conclusions in sections \\ref{sect:4}. ", "conclusions": "\\label{sect:4} In this paper, we do not find a correlation between the initial WD mass and its cooling time. Since the C/O ratio is a function of the initial WD mass and the central density for a WD with given initial mass is mainly determined by the cooling time (\\citealt{NOM99, NOM03}; \\citealt{LESAFFRE06}), our results may imply that the C/O ratio and the central density at ignition are free parameters when simulating an SNe Ia explosion. However, our results still provide a constraint when simulating these types of explosions, i.e. the cooling time of WDs with an initial mass less than 1 $M_{\\odot}$ is generally shorter than 1 Gyr, but it may be as long as 15 Gyr for WDs with an initial mass larger than 1 $M_{\\odot}$ (see Fig. \\ref{mwdcoolt}). Because a high initial WD mass means a lower C/O, and a massive WD and a long cooling time leads to a high central density at ignition (\\citealt{NOM99, NOM03}; \\citealt{LESAFFRE06}), the above result could imply that WDs with a high C/O usually have a lower central density at ignition, but those having the highest central density at ignition generally have a lower C/O. We also checked the effect of metallicity on the distributions of the initial WD mass and the cooling time by $Z=0.001$ (\\citealt{MENGXC10}) and no significant effect was found. Theory and observations did confirm that the effect of metallicity cannot explain the scatter in the maximum luminosity of SNe Ia (\\citealt{TIM03}; \\citealt{GALLAGHER08}; \\citealt{HOWEL09b}). Metallicity should, at most, be the secondary parameter for the Phillips relation, i.e. it is the origin for the scatter of the Phillips relation (\\citealt{POD06, POD08}). Furthermore, even regarding the secondary parameter, there is not a consensus. \\citet{MAZZZALI06} suggested that a variation of the relative content of ($^{\\rm 54}$Fe+$^{\\rm 58}$Ni) versus $^{\\rm 56}$Ni may be responsible for the observed scatter of the Phillips relation. However, \\citet{KASEN09} argued that the breaking of spherical symmetry is a critical factor in determining both the Phillips relation and the observed scatter around it. Then, the origin of both the Phillips relation and the scatter of the relation still should be investigated carefully. We found that the range of the initial WD mass decreases and the average WD mass increases with the cooling time, which is similar to the relation between the initial WD mass and the delay time found by \\citet{MENGXC10a}. In addition, the scatter of the cooling time also decreases with delay time. These results may indicate that the difference among the WDs with s short cooling time could be huge, but the properties of WDs with a long cooling time might be more uniform (see from Fig. \\ref{mwdcoolt} that all the WDs with a cooling time longer than several Gyr have a mass larger than 1.0 $M_{\\odot}$). Since a long cooling time is equivalent to a long delay time (see Fig. \\ref{coolage}), our results may imply that the properties of SNe Ia with long delay times might be more uniform than those with short delay times. It is widely known that there exists a scatter of the maximum luminosity of SNe Ia, and the scatter is affected by its environment. The most luminous SNe Ia always occur in spiral galaxies, but both spiral and elliptical galaxies are hosts for dimmer SNe Ia, which lead to a dimmer mean peak brightness in elliptical than in spiral galaxies, i.e. the maximum luminosity of SNe Ia hosted in elliptical galaxies is more uniform (\\citealt{HAM96}). These properties may be qualitatively interpreted by the model developed by \\citet{MENGXC10a}, at least if the C/O ratio is the origin of the Phillips relation. However, the quantitative study by \\citet{ROPKE06} showed that the C/O has only little impact on the amount of produced $^{\\rm 56}$Ni. It should be noticed that the results in \\citet{ROPKE06} are model dependent, and are much different from those of \\citet{NOM99, NOM03}. Considering that the central density is mainly determined by the initial mass of the WD and its cooling time, the results in Figs. \\ref{mwdcoolt} and \\ref{coolage}, i.e. the average values of the WD mass and the cooling time increase with the delay time, which means that the average of the central density increases with the delay times of SNe Ia. A high central density at ignition leads to a larger amount of the produced $^{\\rm 56}$Ni (\\citealt{ROPKE06}). The effect of the central density on the amount of $^{\\rm 56}$Ni seems then to be opposite with observations of \\citet{HAM96}. Furthermore, the variation of the amount of $^{\\rm 56}$Ni derived from the central density only amounts to about 7\\%, which can not be used to interpret the variation in the maximum luminosity of SNe Ia (\\citealt{ROPKE06}). Perhaps, the C/O ratio, the central density and the metallicity all contribute to the variation of the maximum luminosity of SNe Ia (\\citealt{ROPKE06b}). Then, which is the dominant parameter for the Phillips relation is still an open queation. In addition, as the cooling/delay time increases, the CO WDs become more and more degenerate, and even crystallization may occur (\\citealt{FONTAINE01}). What is the effect of the crystallization on the amount of $^{\\rm 56}$Ni should be an interesting problem. \\normalem" }, "1005/1005.3747_arXiv.txt": { "abstract": "{The blazar \\source has been previously detected by the VERITAS and MAGIC telescopes in the very high energies. The new detection of VERITAS from December 2008 to April 2009 proves that \\source is not static, but shows short-time variability.}{We show that the time variability may be explained in the context of a self-consistent synchrotron-self Compton model, while the long time observation do not necessarily require a time-resolved treatment.}{The kinetic equations for electrons and photons in a plasma blob are solved numerically including Fermi acceleration for electrons as well as synchrotron radiation and Compton scattering.}{The light curve observed by VERITAS can be reproduced in our model by assuming a changing level of electron injection compared to the constant state of \\source. The multiwavelength behaviour during an outburst becomes comprehensible by the model.}{The long time measurements of VERITAS are still explainable via a constant emission in the SSC context, but the short outbursts each require a time-resolved treatment.} ", "introduction": "Blazars are a special class of active galactic nuclei (AGN) exhibiting a spectral energy distribution (SED) that is strongly dominated by nonthermal emission across a wide range of wavelengths, from radio waves to gamma rays, and rapid, large-amplitude variability. The source of this emission is presumably the relativistic jet emitted at a narrow angle to the line of sight to the observer.\\\\In high-peaked BL Lac objects (HBLs) the SED shows a double hump structure as the most notable feature with the first hump in the UV- to X-ray regime and the second hump in the gamma-ray regime. Indeed, a substantial fraction of the known nearby HBLs have already been discovered with Cherenkov telescopes like H.E.S.S., MAGIC or VERITAS. The origin of the first hump is mostly undisputed: nonthermal, relativistic electrons in the jet are emitting synchrotron radiation. The origin of the second hump is still controversially debated. Up to now two kinds of models are discussed: leptonic \\citep[e.g.][]{maraschi92} and hadronic \\citep[e.g.][]{mannheim93} ones, which are mostly applied for other subclasses of blazars.\\\\Another important feature of AGNs in general and HBLs in particular is their strong variability. The dynamical timescale may range from minutes to years. This requires complex models, which obviously have to include time dependence, but this gives us also the chance to understand the mechanisms that drive AGNs. We will apply a self-consistent leptonic model to new data observed for the source \\source, because those are the ones favoured for HBLs.\\\\\\\\The source HBL \\source has been discovered as a candidate BL Lac object on the basis of its X-ray emission and has been identified with the X-ray source \\mbox{2A 1219+30.5} \\citep{wilson79,ledden81}. For the first time, \\source has been observed at VHE energies using the MAGIC telescope in January 2005 \\citep{magic1218} and later from VERITAS \\citep{veritas1218}. Coverage of the optical/X-ray regime is provided by BeppoSAX \\citep{beppo05} and SWIFT \\citep{swift07}, unfortunately the data are not always simultaneous. During the observations from December 2008 to April 2009 VERITAS also observed \\source showing a time-variability \\citep{veritas2010}. The observations from the MAGIC telescope have previously been modelled by \\citet{michl1218}. \\citet{veritas2010} claim that their new observations exhibiting variability challenge the previous models. We will show that a timedependent model using a self-consistent treatment of electron acceleration is able to model the new VERITAS data.\\\\\\\\We present the kinetic equation, which we solve numerically, describing the synchrotron-self Compton emission (Sect. \\ref{sec:model}). In Sect.~\\ref{sec:results} we apply our code to \\source, taking the VERITAS data into account and give a set of physical parameters for the most acceptable fit. Finally, we discuss our results in the light of particle acceleration theory and the multiwavelength features. ", "conclusions": "\\label{sec:discussion} Our results clearly show that the latest observations from the VERITAS telescope for \\source still agree with a constant (steady state) emission from a SSC model when averaged over a long observation period. This is due to the relatively moderate variability of \\source compared to the observation time.\\\\The variability may be well explained in the context of the self-consistent treatment of acceleration of electrons in the jet. We are aware that an outburst of the timescale of roughly five days as measured from \\source does not necessarily require a shock in jet model, which scales down to a few minutes depending on the SSC parameters \\citep{weidinger2010b}, but may also be explained as e.g. different accretion states. Nevertheless the fundamental statement remains the same: long time observation of slightly variable blazars will result in a steady state emission, while an average over a single outburst will, of course, result in a significantly different SED for the source. We are not yet able to rule out different emission models or even complex geometries of the emitting region. But we are able to model the influence of short outbursts of a source on the SED and the lightcurves in the different energy bands selfconsistently.\\\\ The VERITAS collaboration only shows an integrated spectrum for \\source, which is due to the low flux of the source and the photon index behaviour of the combined high-states. This integrated spectrum does not show strong variations with regard to the known low-state observed by MAGIC. Our model now predicts a clear change in the spectrum, which is indicated by the dashed line in Fig. 1, which shows the average over one outburst with a slight, currently not detectable spectral softening in the VHE range, while the synchrotron peak in the BeppoSAX/SWIFT regime remains spectrally constant. This situation changes for shorter and/or stronger outbursts of an overall timescale of hours, which will result in measurable spectral evolutions in all energy regimes when considered with the presented model. Furthermore the time-resolved SEDs during a flare are comprehensible with our model. Hence with better time-resolved spectra or/and better multiwavelength coverage it should be possible to prove this model, and if the model is indeed applicable it will be a good tool to investigate the whole SED during an outburst without having all energy regimes observationally covered.\\\\\\textit" }, "1005/1005.5482_arXiv.txt": { "abstract": "The global-scale interior magnetic field $\\BI$ needed to account for the Sun's observed differential rotation can be effective only if confined below the \\cz\\ in all latitudes including, most critically, the polar caps. Axisymmetric solutions are obtained to the nonlinear magnetohydrodynamic equations showing that such polar confinement can be brought about by a very weak downwelling flow $U\\sim10^{-5}$\\cms\\ over each pole. Such downwelling is consistent with the \\hsc\\ evidence. All three components of the magnetic field $\\B$ decay exponentially with altitude across a thin, laminar ``magnetic \\cl'' located at the bottom of the \\tc\\ and permeated by spiralling field lines. With realistic parameter values, the thickness of the \\cl\\ $\\sim10^{-3}$ of the Sun's radius, the thickness scale being the magnetic advection--diffusion scale $\\CLthick=\\eta/U$ where the magnetic (ohmic) diffusivity $\\eta\\approx 4.1\\times10^{2}$\\cmms. Alongside baroclinic effects and stable thermal stratification, the solutions take into account the stable compositional stratification of the \\HSL, if present as in today's Sun, and the small diffusivity of helium through hydrogen, $\\chi\\approx0.9\\times10^{1}$\\cmms. The small value of $\\chi$ relative to $\\eta$ produces a double boundary-layer structure in which a ``\\hsl'' of smaller vertical scale $(\\chi/\\eta)^{1/2}\\CLthick$ is sandwiched between the top of the \\HSL\\ and the rest of the \\cl. Solutions are obtained using both \\san\\ and purely numerical, finite-difference techniques. The \\chl\\ flows are magnetostrophic to excellent approximation. More precisely, the principal force balances are between Lorentz, Coriolis, pressure-gradient and buoyancy forces, with relative accelerations negligible to excellent approximation. Viscous forces are also negligible, even in the \\hsl\\ where shears are greatest. This is despite the kinematic viscosity being somewhat greater than $\\chi$. \\ We discuss how the \\cl s at each pole might fit into a global dynamical picture of the solar \\tc. That picture, in turn, suggests a new insight into the early Sun and into the longstanding enigma of solar lithium depletion. ", "introduction": "\\label{sec:intro} This paper analyses a new family of laminar magnetostrophic flows that may be important for confining the interior magnetic field $\\BI$ needed to explain the Sun's differential rotation. As illustrated in \\fig\\ref{fig:helioseismology}, the differential rotation observed within the \\cz\\ goes over into near-\\solid\\ rotation within the radiative, stably stratified interior, via a thin shear layer called the ``\\tc'' much of which is also stably stratified. The need for the interior field $\\BI$ has been argued elsewhere \\citep[][hereafter GM98]{McIntyre94, RudigerKitchatinov97, GM98}; the main arguments are briefly recalled below. The observational evidence together with many ideas about the \\tc\\ are reviewed and further referenced in a recent major compendium, \\emph{The Solar Tachocline} \\citep{Hughesetal07}, and further discussed in the second edition of Mestel's \\emph{Stellar Magnetism} \\citep{Mestel11}. By a ``confined'' $\\BI$ we mean a field most if not all of whose lines are contained beneath the \\cz, and held there against magnetic (ohmic) diffusion. Such confinement is well known to be necessary in order for the field to help enforce \\solid\\ rotation in the interior \\citep[\\eg][]{Ferraro37,MestelWeiss87, CharbonneauMacGregor93, MacGregorCharbonneau99}, and thereby keep the \\tc\\ thin. Confinement against magnetic diffusion requires fluid motion. So, besides magnetic effects, a realistic theory of confinement must take account of Coriolis effects, stable stratification, baroclinicity, and thermal relaxation. Without these effects we cannot correctly describe, for instance, the overall torque balance, which necessarily involves mean meridional circulations (MMCs) as well as Maxwell stresses. The first attempt at a tachocline theory was that of \\citet{SpiegelZahn92}. It included all the above effects except $\\BI$. \\ \\citet{RudigerKitchatinov97} included $\\BI$ but omitted the other effects. The first attempt to include all of them was that of GM98, in a line of investigation further developed by \\citet{GaraudGaraud08}. Meanwhile, the dynamical importance of compositional as well as thermal stratification \\citep[\\eg][]{Mestel53} was suggested for tachocline theories \\citep{McIntyre07}. In particular, the helium settling layer beneath the tachocline is nearly impermeable to MMCs because of the small diffusivity of helium through hydrogen. This near-impermeability of compositionally stratified regions has been called the ``\\muchoke'' \\citep{MestelMoss86}. The reality of the Sun's \\HSL\\ is strongly indicated both by standard solar-evolution models and by helioseismology \\citep[\\eg][]{ JCD-etal93, Ciacio-etal97, ElliottGough99, JCDThompson07}.\\footnote{ Also Christensen-Dalsgaard \\& Gough 2011, in preparation. } As will be seen, this combination of circumstances gives rise to some new and interesting fluid dynamics. The need for the interior field $\\BI$ arises from a well known difficulty with non-magnetic theories. They tend to spread the strong differential rotation of the convection zone down into the radiative interior. Although sometimes disputed, this is a robust and well-understood consequence of thermal relaxation, interacting with Coriolis effects and gyroscopically-pumped MMCs \\citep{Haynes-etal91,SpiegelZahn92,Elliott97,McIntyre07,GaraudBrummell08}. As shown by \\citeauthor{SpiegelZahn92} and confirmed by \\citeauthor{Elliott97}, this downward spreading or burrowing would have produced a \\tc\\ far thicker than observed. The accompanying MMC, acting throughout the Sun's lifetime, would also have prevented the \\HSL\\ from forming. To counter the burrowing tendency and to allow the interior to rotate \\solid ly, angular momentum has to be transported somehow from the low-latitude tachocline to the high-latitude tachocline. The non-magnetic horizontal eddy viscosity proposed for this purpose by \\citeauthor{SpiegelZahn92} is inconsistent with the properties of non-magnetic stratified turbulence known from many studies of the terrestrial atmosphere \\citep[][\\& refs.]{McIntyre94,McIntyre03}. Angular-momentum transport by internal gravity waves is a physically possible alternative \\citep[\\eg][\\& refs.]{Schatzman93,Zahn-etal97, RogersGlatzmaier06,CharbonnelTalon07}. However, it is highly improbable as the main mechanism because, by itself, it has no natural tendency to produce \\solid\\ rotation at all latitudes and depths \\citep[\\eg][]{PlumbMcEwan78}. A suitably-shaped magnetic field can, by contrast, naturally produce the required angular momentum transport, via the Alfv\\'enic elasticity of the field lines. A suitable shape is one in which the field lines link low latitudes to high latitudes within the \\tc. The simplest such shape --- simplest by virtue of its axisymmetry --- is that suggested schematically in \\fig\\ref{fig:sphere}, in which the linkage is via a time-averaged field whose lines thread the \\tc, forming the superficial part of a global-scale interior dipole stabilized by a deep toroidal field \\citep[\\eg][]{BraithwaiteSpruit04}. Such an interior dipole has a diffusive lifetime somewhat greater than the Sun's lifetime of around $4.5\\times10^9$\\yr. The dipole imposes an Alfv\\'enic ``Ferraro constraint'' on the interior. It is this constraint that helps to enforce the interior's \\solid\\ rotation \\citep[\\eg][]{Ferraro37,MestelWeiss87, MacGregorCharbonneau99}. \\begin{figure} \\centering \\subfigure[][]{ \\label{fig:helioseismology} \\psfrag{cz}{\\begin{tabular}{c}convection\\\\[-0.1em]zone\\end{tabular}} \\psfrag{rz}{\\begin{tabular}{c}radiative\\\\[-0.1em]interior\\end{tabular}} \\psfrag{tc}{tachocline} \\includegraphics[clip=true,bb=118 -12 614 510,height=6cm]{Schou-etal-bw.eps} } \\hspace{1cm} \\subfigure[][]{ \\label{fig:sphere} \\includegraphics[clip=true,bb=14 -152 878 1048,height=6cm]{spherebw.eps} } \\caption{\\footnotesize (a) The Sun's differential rotation deduced from helioseismic data using inverse methods \\citep[adapted from][]{Schou-etal98}. The radiative interior rotates approximately \\solid ly with angular velocity $\\OmegaI = 2.7\\times10^{-6}$\\sm, or 435\\,\\nHz. Within the \\cz, the angular velocity increases with colatitude through 350, 400, 450\\,\\nHz\\ (heavy contours) to a maximum just under 470\\,\\nHz\\ at the equator. \\newline (b) Schematic illustration showing the top of the radiative interior (inner sphere) and the time-averaged magnetic field threading the \\tc\\ just above. The cutaway outer sphere indicates the top of the \\tc, whose depth has been exaggerated. Poloidal magnetic field lines emerge from the interior in high latitudes and are wound up into their curved shapes by the \\tc's differential rotation, acting against turbulent eddy diffusion. A prograde torque is transmitted from low to high latitudes along these field lines. The slow polar and fast equatorial rotation are indicated by the darker shadings of the outer sphere. The dashed lines indicate the latitudes at which the rotation of the \\cz\\ matches that of the interior. } \\end{figure} The field lines shown in \\fig\\ref{fig:sphere} emerge from the interior (light-grey sphere) near the north pole and, after threading their way through the \\tc, re-enter the interior near the south pole. They return northward through an interior ``apple-core'' region, not shown, surrounding the rotation axis. It is crucial that the field lines emerging from the interior bend over toward the horizontal as they enter the \\tc. They must be prevented from extending upward through the polar cap, as occurs when magnetic diffusion dominates \\citep[\\eg][]{BraithwaiteSpruit04,BrunZahn06}. The curved shapes of the field lines in \\fig\\ref{fig:sphere} are evidently such as to transport angular momentum from low to high latitudes by means of persistent Alfv\\'enic torques, exactly as required to prevent the \\tc's MMCs from burrowing into the interior and thickening the \\tc. The time averaging envisaged in \\fig\\ref{fig:sphere} conceals a plethora of fast processes, including the 22-year dynamo cycle, convective overshoot, and other turbulent processes arising from various instabilities in the \\tc. All these are fast relative to the timescales on which the mean structure of the \\tc\\ is maintained, $\\sim10^5$\\yr\\ or more. We presume that the fast processes have two important consequences. The first is to produce a turbulent magnetic diffusivity that stops the field lines being wound up arbitrarily tightly by the shear in the \\tc, keeping the curved shapes shown.\\footnote{Of course the persistent angular momentum transport from the curved $\\B$ lines could be supplemented by equally persistent contributions coming directly from MHD-turbulent stresses \\citep[\\eg][]{Spruit02,GilmanCally07,ParfreyMenou07}. } The second important consequence is that, away from the poles, the field lines are held down, and held approximately horizontal, by turbulent ``\\mfp'' from the convective overshoot layer. The effectiveness of such flux pumping can be strongly argued from several lines of evidence, including three-dimensional direct numerical simulations, with varying emphasis on the role of turbulent anisotropy and of vertical gradients of density and turbulent intensity \\citep[\\eg][\\& refs.]{Tobias-etal01, KitchatinovRudiger06} \\citep[see also \\S3 of][for a historical review]{Weiss-etal04}. Near the poles it is less clear that \\mfp\\ will be effective in confining the field. At least its effectiveness for near-vertical magnetic fields has not, to our knowledge, been convincingly demonstrated. However, as argued for instance in GM98, there are good reasons in any case (\\S\\ref{sec:downwelling} below) to expect the \\tc's MMC near the poles to take the form of weak but persistent downwelling. This suggests that the field can, in any case, be confined in the polar caps through an advection--diffusion balance, the kind of balance argued for heuristically in GM98. The purpose of this paper is to show in detail, by solving an appropriate set of nonlinear magnetohydrodynamic equations, that such polar confinement by downwelling is indeed possible in a physically realistic model, applicable to the Sun both today and early in its main-sequence lifetime. A large family of axisymmetric nonlinear solutions showing polar confinement has been obtained using two different techniques. The first technique is \\san\\ in a sense to be explained, and the second is numerical on a \\mbox{2-dimensional} grid. The solutions are to be regarded as candidate solutions for possible flows in the real Sun, all showing confinement in the sense that the total magnetic field strength $|\\B|$ dies off exponentially with altitude, thanks to downward advection acting against upward diffusion. In this sense the poloidal and toroidal field components are both well confined. We call these flows ``\\cl s''. They are not to be confused with the \\tc\\ itself. Rather, they occupy relatively thin regions at the bottom of the \\tc\\ and are much more weakly sheared, with relatively long timescales $\\sim 10^5$\\yr. The detailed dynamics involves not only magnetic advection, stretching, twisting and diffusion but also a near-perfect balance between Lorentz, Coriolis, pressure-gradient and buoyancy forces (\\S\\S\\ref{sec:model-equations}ff.). Thus the \\chl\\ flows are magnetostrophic, like certain flows that have been studied in connection with models of the Earth's liquid core \\citep[\\eg][\\& refs.]{Kleeorinetal97}, though different in most other respects. For instance the latter flows are viscous but unstratified: buoyancy forces and thermal diffusion are absent. In the \\chl\\ flows studied here, by contrast, viscosity turns out to be wholly unimportant while buoyancy and thermal diffusion are crucial, along with magnetic diffusion. \\begin{figure} \\centering \\psfrag{U}[l][]{$\\mathbf{u}$} \\psfrag{B}[tl][B]{$\\mathbf{B}$} \\psfrag{W}[][t]{$_{\\phantom{\\ii}}\\OmegaI$} \\psfrag{TC}[][]{\\sc \\tc} \\psfrag{cl}[][]{\\begin{tabular}{c}confinement\\\\[-0.1em]layer\\end{tabular}} \\psfrag{sl}[][]{\\begin{tabular}{c}helium\\\\[-0.1em]sublayer\\end{tabular}} \\psfrag{HSL}[][]{\\sc\\begin{tabular}{c}helium\\\\[-0.1em]settling\\\\[-0.1em]layer\\end{tabular}} \\includegraphics[clip=true,bb=-205 150 818 675,width=10cm]{new-section.eps} \\caption{\\footnotesize The magnetic confinement layer near the north pole in a model for today's Sun. The field strength $|\\B|$ falls off exponentially with altitude $z$. \\ The toroidal components of $\\B$ and velocity $\\bu$ are not shown. The streamlines with arrows show the downwelling responsible for the confinement. If the downwelling were switched off, the field near the pole would diffuse and become nearly vertical, as illustrated for instance in \\citet{BrunZahn06}. Compositional stratification is indicated by shading. The plot is from a numerical solution; the corresponding \\san\\ solution looks almost identical. The horizontal and vertical axes are colatitudinal distance $r$ and altitude $z$ in units of $\\CLthick$, the advection--diffusion scale defined in (\\ref{eq:CLthickdef}). With typical parameter values, the scale $\\CLthick$ is of the order of a fraction of a megametre, $\\sim10^{-3}$ of the Sun's radius. } \\label{fig:section} \\end{figure} \\Fig\\ref{fig:section} gives a preview of a typical \\chl\\ flow, seen in vertical cross-section. It shows the poloidal velocity and magnetic field components from a numerical solution. The emerging magnetic field lines are bent over within the \\cl, as required to fit into the global picture sketched in \\fig\\ref{fig:sphere}. The magnetic field $\\B$ has a toroidal component, not shown in the figure, imparting spiral shapes to the three-dimensional field lines and providing the prograde Alfv\\'enic torque demanded by the global picture, in balance with a retrograde Coriolis torque on the equatorward flow. The vertical and colatitudinal distances in \\fig\\ref{fig:section} are shown in units of the magnetic advection--diffusion scale \\vspace{-0.15cm} \\begin{equation} \\CLthick \\definedas \\eta/U \\,, \\label{eq:CLthickdef} \\end{equation} say, where $U$ is the magnitude of the downwelling and $\\eta$ is the magnetic diffusivity. Throughout this paper, we assume that the \\chl\\ flow is laminar, and therefore use molecular or microscopic diffusivity values (\\S\\ref{sec:model-equations}). Issues of stability or instability lie beyond the scope of this paper but, close to the pole at least, there appears to be a strong case for stability, to be argued in a future paper, arising from the smallness of the scale $\\CLthick$. Under reasonable assumptions, $\\CLthick$ is only a fraction of a megametre, far smaller than the thickness of the overlying \\tc\\ which latter, by contrast, is probably unstable and indeed turbulent, as already mentioned \\citep[\\eg][]{Spruit02,GilmanCally07,ParfreyMenou07}. In the present-day Sun's \\HSL, the top of which corresponds to the shaded region in \\fig\\ref{fig:section}, a downward gradient of helium concentration reinforces the stable stratification due to the sub-adiabatic temperature gradient. Because the diffusivity of helium through hydrogen, $\\chi\\approx0.9\\times10^{1}$\\cmms, is much less than the magnetic diffusivity $\\eta\\approx4.1\\times10^2$\\cmms, the \\HSL\\ is nearly impermeable to the \\chl\\ flow. Helium advection and diffusion are comparable only in the extremely thin ``\\hsl'' marked in \\fig\\ref{fig:section}. In this and other respects, all the solutions in the present paper supersede those described in a first report on this work \\citep[][hereafter WM07]{WM07}. For instance, in WM07 we took $\\chi$ to be zero, implying a \\hsl\\ of vanishing thickness. We also took $\\nu$, the kinematic viscosity, to be zero and allowed a finite slip velocity at the top of the \\HSL, assuming that this slip velocity would in reality be resolved into a weak Ekman layer. However, the solutions presented here show that, on the contrary, no Ekman layer forms. The slip discontinuity is replaced by a smooth velocity profile across the \\hsl\\ and, as will be shown in \\S\\ref{sec:helium}, the flow stays essentially inviscid. The plan of the paper is as follows. In \\S\\ref{sec:downwelling} we summarize the reasons for expecting persistent downwelling over the poles. In \\S\\ref{sec:model-equations} we present the model equations and in \\S\\ref{sec:self-similar} the \\san\\ solutions. Those solutions rely on assuming a self-similar horizontal structure that is asymptotically valid in the limit of strong stable stratification. The same limit was taken in WM07. The validity of the strong-stratification limit is assessed in \\S\\S\\ref{sec:CL-scalings} and~\\ref{sec:helium}, which take a thorough look at the dynamical balances and scalings in the \\cl\\ and \\hsl\\ respectively. Strong stable stratification means that the thermal and compositional stratification surfaces are ``flat'', meaning gravitationally horizontal, to sufficient approximation in some region surrounding the poles, which for reasonable parameter values can be quite large in horizontal extent, up to tens of degrees of colatitude. Within the \\hsl, the low magnetic Reynolds number and flat geometry cause the momentum balance to take on the character of flow in a porous medium, as fluid pushes horizontally past the field lines. As already indicated, true viscous effects are negligible everywhere, even in the sublayer. Boundary conditions for the numerical solutions are discussed in \\S\\ref{sec:boundary-conditions}. The numerical solutions themselves are presented and discussed in \\S\\ref{sec:num-solutions}. They provide cross-checks with the \\san\\ solutions plus additional insights. In particular, they directly demonstrate the flatness of the stratification surfaces by solving the full equations, for finite stratification. The solutions allow the stratification surfaces to tilt as they may, but confirm that the departures from flatness are indeed small when the stratification is realistically strong. In \\fig\\ref{fig:section}, for instance, the departures from flatness are barely visible. In \\S\\ref{sec:comparison} we discuss a subtlety that arises when comparing the \\san\\ and numerical solutions in the upper part of the flow. The dynamical balances aloft become delicate as the Lorentz and Coriolis forces become vanishingly small. The effects of truncation error and other small effects thus complicate the comparison. However, this is something of an academic point because of our expectation that, in reality, the \\chl\\ solutions will need to be matched to a turbulent \\tc\\ aloft, a task that remains a challenge for the future. In \\S\\ref{sec:noHSL} we show that the presence of the \\HSL\\ is not crucial to our \\chl\\ model. The interior field $\\BI$ is sufficient by itself to turn the flow equatorward, and the field remains confined in much the same way. That result has relevance to the Sun's early main-sequence evolution. It explains for instance how the burrowing tendency could have been held in check from the start, allowing the \\HSL\\ to form. In the concluding discussion, \\S\\ref{sec:conclusions}, we consider the implications for early solar evolution and lithium depletion. ", "conclusions": "\\label{sec:conclusions} We cannot yet claim to have a complete \\tc\\ theory. Indeed, the \\cl\\ and \\hsl\\ form only two pieces of a complicated jigsaw puzzle. Other aspects of that jigsaw include the way in which the \\cl\\ matches upward to the relatively large negative shear in the bulk of the \\tc, and the way in which the baroclinic temperature anomalies induced by the \\tc's MMC fit into the perturbed global-scale heat flow. In particular, without putting the whole jigsaw together we cannot quantitatively predict the thickness of the \\tc. Nor can we predict the precise shapes of the vertical profiles of $\\B$ and $\\bu$ in the \\cl. Those profile shapes depend on matching to conditions not only aloft but also equatorward, where stratification surfaces and field lines extend into colatitudes outside the polar downwelling regions. However, the results obtained here give us the first fully consistent model of polar field confinement, as such, together with insight into how it could work in today's Sun. The results cover a wide range of possible downwelling values and interior field strengths (end of \\S\\ref{sec:CL-scalings}). We have also shown, in \\S\\ref{sec:noHSL}, how confinement could have worked in the early Sun. The dynamics is similar apart from the slightly deeper penetration of the MMC in the absence of the \\HSL\\ and \\hsl. We can use the resulting insights, alongside our well-established understanding of the gyroscopic pumping of MMCs, to say something new about the early Sun and the solar lithium-burning problem. Standard solar-evolution models predict surface lithium abundances higher than observed by a factor $\\sim10^2$ \\citep[\\eg][]{Vauclair-etal78}. The reason is that the standard models mix material down to the bottom of the \\cz\\ but no further. To destroy lithium, material from the \\cz\\ must be mixed or circulated to somewhat greater depths and therefore to somewhat higher temperatures, beyond those at the bottom of today's \\tc. However, there is no evidence of depletion of the \\cz's beryllium, which is destroyed at only moderately higher temperatures than lithium. Further discussion and references may be found in \\citet{JCDGoughThompson92}, and in \\citet[][chap.\\ 6]{Wood10}. Here we argue that a quantitative version of the scenario sketched in \\fig\\ref{fig:sphere} has promise as a way of circulating material to the required depth in the early Sun, and no further, thus making sense of the high beryllium as well as the low lithium abundance. As already mentioned in \\S\\ref{sec:intro}, the downwelling MMC in the polar \\tc\\ that makes field confinement possible can be regarded as due to a gyroscopically-pumped MMC trying to burrow downward, but held in check by its encounter with the interior field $\\BI$ and with the \\HSL, if present. If, in a \\te, we were to switch off the interior field $\\BI$, then the downwelling would spread or burrow to ever-increasing depths. The timescale for such burrowing is inversely proportional to $\\OmegaI^2$ \\citep[\\eg][\\eq(8.15)ff.]{McIntyre07}; one may think of rotational stiffness as strengthening the burrowing tendency. Now, because the early Sun rotated much faster than today, not only would there have been no \\HSL\\ but, also, the burrowing tendency would have been much stronger than today, tending to push the bottom of the \\tc\\ downward. This reopens the possibility conjectured in {GM98} that there might have been a ventilated ``polar pit'' in which most of the \\cz's lithium, though not too much of its beryllium, was burnt during the first gigayear or two of the Sun's main-sequence evolution. To take this further we again need to consider the way in which the \\cl\\ fits into the global picture. It is arguable that the bottom of the entire polar downwelling region is depressed relative to its surroundings, forming not so much a ``pit'' as a shallow ``frying pan'', too shallow to burn lithium in today's Sun but possibly just deep enough in the early Sun. \\begin{figure} \\centering \\includegraphics[width=7cm]{pit3.eps} \\caption{\\footnotesize Schematic drawing of the magnetic confinement layer and its immediate surroundings at the \\base\\ of the high-latitude \\tc. Close to the pole the interior magnetic field (solid lines) is confined by the downwelling MMC (dashed streamlines). The vertical scale has been greatly exaggerated.\\label{fig:surroundings}} \\end{figure} Here we need to distinguish the shape of the ventilated region from the shapes of the stratification surfaces, which latter must remain relatively flat, meaning close to the horizontal. \\Fig\\ref{fig:surroundings} sketches the way in which the \\cl\\ might fit into its surroundings near the bottom of the polar \\tc. The stratification surfaces are shown dotted. At the periphery of the polar downwelling region, the field lines (solid) spiral outward and upward from the \\cl\\ on their way to lower latitudes. They will tend to splay out, as well as slanting upward, as they emerge from the downwelling region. The MMC will similarly slant upward, flowing approximately along the field lines (dashed streamlines). This is because the splaying-out increases the magnetic Reynolds number beyond the order-unity values characteristic of the \\cl. Further out, the field lines must continue to rise through the \\tc\\ until they encounter the \\cz's overshoot layer, where they are held horizontal by turbulent \\mfp\\ as suggested in \\fig\\ref{fig:sphere}. On the way we must expect turbulent eddy fluxes to become increasingly important, decoupling the MMC's upwelling streamlines from the time-averaged field lines and leaving the upwelling free to spread over a wide range of latitudes, constrained only by mass conservation and global-scale heat flow. Such a picture applies equally well to today's Sun and to the early Sun, the main difference being that the ventilated polar region (unshaded in \\fig\\ref{fig:surroundings}) is likely to have been pushed deeper in the early Sun with its much faster rotation, stronger burrowing tendency, and global-scale $|\\BI|$ values only modestly larger. The ventilated polar regions could well have been deeper by many tens of megametres, as required to burn lithium.\\footnote{This deepening is additional to the deepening of the \\cz\\ itself, in the early Sun relative to today's Sun, amounting to several more tens of megametres according to standard solar models \\citep[e.g.][]{Ciacio-etal97}. } The early Sun would have started to form its \\HSL\\ just below these ventilated, lithium-destroying polar regions, \\ie\\ just below the polar \\cl s. Then, with the gradual diminution of $\\OmegaI$ through solar-wind braking, the \\cl s, marking the bottom of the ventilated regions, would have retreated upward, and the top of the \\HSL\\ would have followed them upward as new helium strata formed. Within the peripheral lightly-shaded region in \\fig\\ref{fig:surroundings}, into which the MMC does not penetrate, we suggest that ventilation is weak or nonexistent and that shear will be limited by the Ferraro constraint. The darker shading represents the top part of today's \\HSL. As the lightly-shaded region expands upward and outward beyond the immediate surroundings sketched in \\fig\\ref{fig:surroundings}, through the \\tc\\ toward the overshoot layer, we may surmise that small-scale MHD instabilities will kick in \\citep[\\eg][\\& refs.]{Spruit02,GilmanCally07,ParfreyMenou07}, breaking the Ferraro constraint and blurring the distinction between the shaded and unshaded regions as turbulent eddy fluxes increase. So a larger-scale picture of the ``lithium frying pan'' would show its upward-sloping lower boundary becoming increasingly porous and indistinct at greater colatitudes. The global \\tc\\ model that would be needed to test, and to begin to quantify, the foregoing speculations would have to describe \\smallskip \\begin{enumerate} \\item the precise way in which turbulent stresses in the convection zone and \\tc\\ gyroscopically pump the polar downwelling needed to confine $\\BI$ in polar latitudes; \\item the global-scale distribution of temperature and heat flow that fits in with the MMCs; \\item the turbulent \\mfp\\ by convective overshoot assumed to confine $\\BI$ in \\extrapolar\\ latitudes; \\item the extent to which the winding-up of the time-averaged toroidal field in \\extrapolar\\ latitudes (\\fig\\ref{fig:sphere}) is limited by turbulent eddy fluxes; \\item the reaction of the overlying turbulent layers to all of the above, especially the deficit in the \\cz's differential rotation governing the torques exerted from above, for instance via feedback on the strength of gyroscopic pumping of the MMC. \\end{enumerate} \\smallskip Progress on these formidable problems will depend on finding suitable ways to model the turbulent processes." }, "1005/1005.5161_arXiv.txt": { "abstract": "We reconsider non-minimal $\\lambda\\,\\phi^4$ chaotic inflation which includes the gravitational coupling term $\\xi\\,\\mathcal{R}\\,\\phi^2$, where $\\phi$ denotes a gauge singlet inflaton field and $\\mathcal{R}$ is the Ricci scalar. For $\\xi \\gg 1$ we require, following recent discussions, that the energy scale $\\lambda^{1/4} m_P / \\sqrt{\\xi}$ for inflation should not exceed the effective UV cut-off scale $m_P / \\xi$, where $m_P$ denotes the reduced Planck scale. The predictions for the tensor to scalar ratio $r$ and the scalar spectral index $n_s$ are found to lie within the WMAP 1-$\\sigma$ bounds for $10^{-12} \\lesssim \\lambda \\lesssim 10^{-4}$ and $10^{-3} \\lesssim \\xi \\lesssim 10^2$. In contrast, the corresponding predictions of minimal $\\lambda\\,\\phi^4$ chaotic inflation lie outside the WMAP 2-$\\sigma$ bounds. We also find that $r \\gtrsim 0.002$, provided the scalar spectral index $n_s \\geq 0.96$. In estimating the lower bound on $r$ we take into account possible modifications due to quantum corrections of the tree level inflationary potential. ", "introduction": " ", "conclusions": "" }, "1005/1005.3021_arXiv.txt": { "abstract": "The aim of this paper is to study the astrometric trajectory of microlensing events with an extended lens and/or source. We consider not only a dark lens but also a luminous lens as well. We find that the discontinuous finite-lens trajectories given by Takahashi will become continuous in the finite-source regime. The point lens (source) approximation alone gives an under (over)estimation of the astrometric signal when the size of the lens and source are not negligible. While the finiteness of the source is revealed when the lens transits the surface of the source, the finite-lens signal is most prominent when the lens is very close to the source. Astrometric microlensing towards the Galactic bulge, Small Magellanic Cloud and M31 are discussed, which indicate that the finite-lens effect is beyond the detection limit of current instruments. Nevertheless, it is possible to distinguish between self-lensing and halo lensing through a (non-)detection of the astrometric ellipse. We also consider the case where the lens is luminous itself, as has been observed where a lensing event was followed up with the \\textit{Hubble Space Telescope}. We show that the astrometric signal will be reduced in a luminous-lens scenario. The physical properties of the event, such as the lens-source flux ratio, the size of the lens and source nevertheless can be derived by fitting the astrometric trajectory. ", "introduction": "Most of the microlensing events detected to date are through photometric monitoring of the source flux. In this case, the information on the physical identity of the lens is reduced, because the only quantity one can retrieve from the light curve is the Einstein timescale $t_{\\mathrm{E}}$. $t_{\\mathrm{E}}$ is defined by the time required for the source to transit the angular Einstein radius $\\AERR$ of the lens \\citep{2000ApJ...542..785G}: \\begin{equation} t_{\\mathrm{E}} = \\frac{\\AERR}{|\\mbox{\\boldmath $\\mu_{rel}$}|},\\quad \\AERR = \\sqrt{k\\ML\\pi_{rel}},\\quad k\\equiv \\frac{4G}{c^2 \\mathrm{AU}}\\approx 8.14\\frac{\\mathrm{mas}}{M_{\\odot}}, \\label{eq.tE} \\end{equation} where \\mbox{\\boldmath $\\mu_{rel}$} is the relative lens-source proper motion, $\\ML$ is the mass of the lens, $\\pi_{rel}:= \\mathrm{AU}/(\\Dol^{-1}-\\Dos^{-1})$ is the relative lens-source parallax, $\\Dol$ and $\\Dos$ are distance to the lens and the source from the observer, respectively. Equation~(\\ref{eq.tE}) shows that the mass, distance, and velocity of the lens are degenerated into $t_{\\mathrm{E}}$. To better constrain the lens properties, \\cite{1995A&A...294..287H}, \\cite{1995ApJ...453...37W} and \\cite{1995AJ....110.1427M} thus suggested to use astrometric microlensing. That is, to measure the centroid displacement of the two images during the course of microlensing. Former studies have shown that the trajectory of the centroid displacement will trace out an ellipse, and the size of the ellipse is proportional to the angular Einstein radius. Therefore, one can determine $\\AERR$ through the observation of such astrometric ellipses and constrain the relative lens-source proper motion. \\cite{1992ApJ...392..442G} has shown that if one can further measure the microlens parallax $\\pi_{\\mathrm{E}}=\\sqrt{\\pi_{rel}/(k\\ML)}$ form the light-curve distortion induced by the orbital motion of the Earth, the lens mass $\\ML$ and the relative lens-source parallax $\\pi_{rel}$ can be determined without ambiguity: \\begin{equation} \\ML = \\frac{\\AERR}{k\\pi_{\\mathrm{E}}},\\quad \\pi_{rel}=\\pi_{\\mathrm{E}}\\AERR. \\label{eq.M} \\end{equation} The location of the lens can be derived as well if the distance to the source is well known, which is often the case towards the Galactic bulge and Magellanic Clouds. The typical value of the astrometric microlensing signal for a source in the Galactic bulge and a 0.5 $M_{\\odot}$ lens located half-way to the source is of order of 0.1 mas, which is much larger than the astrometric accuracy of upcoming space missions such as \\textit{Space Interferometry Mission} \\citep[\\textit{SIM}; ][]{1997SPIE.2871..504A}, \\textit{Global Astrometric Interferometer for Astrophysics} \\citep[\\textit{GAIA}; ][]{1994SPIE.2200..599L} and ground-based instruments, e.g. \\textit{Phase Referenced Imaging and Micro-arcsecond Astrometry} \\citep[\\textit{PRIMA}; ][]{1998SPIE.3350..807Q}. \\textit{GAIA} will survey the whole sky with sources brighter than 20 mag in \\textit{V} band. It is expected to reach an astrometric accuracy of 30 $\\mu$as (150 $\\mu$as) with $V <$ 12 ($V <$ 16) for a single measurement \\citep{2002MNRAS.331..649B} and an estimated detection of $\\approx$ 1000 events \\citep{2000ApJ...534..213D}. Unlike \\textit{GAIA}, which only scans the sky with a pre-determined pattern, \\textit{SIM} can point to selectable targets and thus tracks the ongoing microlensing event upon request. The expected accuracy of \\textit{SIM} is 5 $\\mu$as (20 $\\mu$as) for $V <$ 12 ($V < $ 16) with 1-hour integration time \\citep{2008SPIE.7013E.151G}. While \\textit{SIM} and \\textit{GAIA} are scheduled to launch in the next few years, \\textit{PRIMA} has already been installed on the Very Large Telescope Interferometer (VLTI) and aims at achieving 10-$\\mu$as accuracy level in 30 min integration time provided a reference star within 10 arcsec and a 200-m baseline \\citep{2008NewAR..52..199D}. In addition to the standard point-source point-lens (PSPL) microlensing, single-lens events revealing an extended source signal have also been observed photometrically \\citep[e.g.][]{1997ApJ...491..436A, 2004ApJ...617.1307J, 2004ApJ...603..139Y, 2006A&A...460..277C, 2009A&A...508..467B, 2009ApJ...703.2082Y, 2009arXiv0912.2312Z, 2010A&A...518A..51F}. \\cite{1998MNRAS.300.1041M} thus derived the astrometric trajectory of finite-source events with a point-lens (FSPL). On the other hand, \\cite{2003ApJ...595..418T} studied the centroid displacement of finite-lens effects but assuming a point-source (PSFL). Furthermore, \\cite{2002ApJ...579..430A} and \\cite{2009ApJ...695..200L} have investigated the combination of finite-source and finite-lens (FSFL) effects photometrically, but left aside the astrometric aspect. Since the FSFL light curve deviates from either the PSFL or FSPL, as shown by \\cite{2002ApJ...579..430A}, we are motivated to study the astrometric behaviour when both FS and FL effects are relevant. There are events where both the source and the lens are resolved by the \\textit{Hubble Space Telescope (HST)}. This implies that the lens can also be a star and implies a luminous-lens scenario rather than a dark lens \\citep[][]{2001Natur.414..617A, 2007ApJ...671..420K}, which is also the case for self-lensing. We thus consider the light contribution from the lens star and study the astrometric behaviour by allowing for a luminous lens. This paper is organized as follows. In \\textsection~2 we introduce the theory of astrometric microlensing. We take into account the FS effects either with a uniform surface brightness source or with a more general surface brightness profile in \\textsection~3. We further include a dark lens with finite size in \\textsection~4. One might expect not only shadowing but also light contribution from the lens as well. Therefore, we allow for a luminous lens in \\textsection~5. The aforementioned properties of the microlensing system can be estimated by fitting the formula in \\textsection~5. A discussion of possible events with sources located in the Galactic bulge, Small Magellanic Cloud (SMC) and M31 is presented in \\textsection~6 followed by a summary in \\textsection~7. \\\\ ", "conclusions": "We have studied the astrometric aspects of microlensing by simultaneously including the FS and FL effects. Our results show that the astrometric signal is underestimated or overestimated by assuming PL or PS, respectively. While the FS effect is prominent when the lens transits the surface of the source, the FL effect is revealed when the lens is very close to the source, which would be in the self-lensing regime. In the context of the self-lensing scenario, where a background star is lensed by a foreground star, the light contribution from the lens is in general not negligible. We thus consider the luminous-lens scenario, which attenuates the signal of the centroidal displacement. Astrometric trajectories with a source located in the Galactic bulge, SMC, and M31 are discussed, which show that $\\AERR$ of halo-lensing events is at least one order of magnitude larger than that of self-lensing in SMC and M31. Our results also indicate that the finiteness of the lens is more likely to be revealed in the self-lensing scenario towards distant source located in Magellanic Clouds or M31, although it is very difficult to distinguish between PL and FL with current instruments." }, "1005/1005.3810_arXiv.txt": { "abstract": "Dark energy is an important science driver of many upcoming large-scale surveys. With small, stable seeing and low thermal infrared background, Dome A, Antarctica, offers a unique opportunity for shedding light on fundamental questions about the universe. We show that a deep, high-resolution imaging survey of 10,000 square degrees in \\emph{ugrizyJH} bands can provide competitive constraints on dark energy equation of state parameters using type Ia supernovae, baryon acoustic oscillations, and weak lensing techniques. Such a survey may be partially achieved with a coordinated effort of the Kunlun Dark Universe Survey Telescope (KDUST) in \\emph{yJH} bands over 5000--10,000 deg$^2$ and the Large Synoptic Survey Telescope in \\emph{ugrizy} bands over the same area. Moreover, the joint survey can take advantage of the high-resolution imaging at Dome A to further tighten the constraints on dark energy and to measure dark matter properties with strong lensing as well as galaxy--galaxy weak lensing. ", "introduction": "Antarctic Plateau, especially the Kunlun Station under construction by the Polar Research Institute of China (PRIC), provides a unique opportunity for wide field astronomical surveys targeting cosmological studies. Astronomical site survey of Dome A, Antarctica was enabled by the International Polar Year (IPY) endorsed PANDA program~\\citep{Yang09} led by the PRIC and the Chinese Center for Antarctica Astronomy (CCAA). Several international teams contributed to this effort. In particular, the power and on-site laboratory system built by the University of New South Wales (BUN'S) has provided the platform for all the site survey instruments. In its two years' operation, the site survey effort proves practically all aspects of the theoretical expectations of the Dome A site for astronomical observations. Preliminary analyses show that the boundary layer of atmospheric turbulence to be around $10-20$ meters during the Antarctic winter~\\citep{ashley10}. Similar to the relatively better studied neighboring Dome C site~\\citep{fossat10}, the Dome A site may enjoy free atmospheric seeing conditions of about 0.3 arcsec seeing above this boundary layer, thus making it an ideal site for high angular resolution wide area surveys. The temperature at Dome A is around $-60$ to $-70 \\degree{C}$, making it the coldest spot on the surface of the Earth. This implies a very low thermal background emission in the thermal infrared. The Dome A site is thus also the best site for astronomical observations in the near infrared wavelength region. One other especially exciting property of the site is the lack of water vapors due to the high altitude and the low temperature. This implies that the site is also ideal for terahertz observations which have been impossible from any temperate sites on Earth. While quantitative site properties are still under analyses and longer term monitoring are still needed to firmly establish the astronomical potential of the Dome A site, there is no doubt that a survey project at Dome A can be highly complimentary to programs such as the Large Synoptic Survey Telescope\\footnote{See \\url{http://www.lsst.org/}.} \\citep[LSST,][]{lsst09}, the Joint Dark Energy Mission\\footnote{See \\url{http://jdem.gsfc.nasa.gov/}.}, and Euclid\\footnote{See \\url{http://sci.esa.int/euclid/}.}. For example, a survey at Dome A can provide near infrared data that compliments the deep optical band survey of the LSST; a deep survey in the near infrared combined with the optical data from LSST can reveal high redshift objects at $z\\sim10$, which are not detectable in the optical. The Kunlun Dark Universe Survey Telescope (KDUST) is a 6-to-8-meter wide-area survey telescope being designed by the CCAA. The preliminary design includes a $3\\times3$ square degree optical camera with $0''.15$ pixel, and an infrared camera of $1\\times1$ square degree at $0''.1$/pixel optimized for $1-3.5$ $\\mu$m surveys. One of the key science missions of KDUST is to investigate the mystery of the accelerated cosmic expansion \\citep{riess98,perlmutter99a} using multiple techniques. In this paper, we estimate how well an ideal 10,000 deg$^2$ \\emph{ugrizyJH} survey can constrain the dark energy equation of state (EOS) with weak lensing (WL), baryon acoustic oscillations (BAOs), and type Ia supernova (SNe) luminosity distances. These dark energy probes have different sensitivities to the cosmic expansion and structure growth as well as various systematic uncertainties in the observations, and hence are highly complementary to each other for constraining dark energy properties \\citep[e.g.,][]{knox05,zhan06d,zhan09}. The SNe technique relies on the standardizable candle of the SNe intrinsic luminosity \\citep{phillips93} to measure the luminosity distance, $D_{\\rm L}(z)$. Dark energy properties can then be inferred from the distance--redshift relation. The BAO technique utilizes the standard ruler of the baryon imprint on the matter (and hence galaxy) power spectrum \\citep{peebles70, bond84} to measure the angular diameter distance, $D_{\\rm A}(z)$, and, if the redshifts are sufficiently accurate, the Hubble parameter, $H(z)$ \\citep{eisenstein98, cooray01b, blake03, hu03b, linder03, seo03}. The WL technique has the advantage that it can measure both $D_{\\rm A}(z)$ from the lensing kernel and the growth factor of the large-scale structure $G(z)$ \\citep{hu99,huterer02b,refregier03,takada04,knox06b,zhan09}. Because of the excellent seeing condition and infrared accessibility at Dome A, KDUST has a number of advantages for the commonly used cosmological probes. For example, the signal-to-noise ratio for point sources is inversely proportional to the seeing. Thus, a 6 meter telescope at Dome A ($\\sim 0''.3$ median seeing in the optical) would be equivalent to a 14 meter telescope at a temperate site ($\\sim 0''.7$ seeing) for point-source observations with the same sky background level. An 8m KDUST could detect SNe out to redshift 3 in the $K_{\\rm dark}$ ($2.27$--$2.45\\mu$m, redward of $K$) band \\citep{kim2010}. Although high-z distances are not sensitive to conventional dark energy, they can be used to determine the mean curvature accurately, which, in turn, helps constrain dark energy EOS \\citep{linder05b,knox06c}. Moreover, even though dark energy is thought to be sub-dominant at high redshift, there is no direct evidence to prove one way or another. Measurements of SNe at $z > 2$ will provide crucial data for tests of early dark energy. Small and stable seeing is particularly helpful for WL. One could resolve more galaxies at the same surface brightness limit, which reduces the shape noise for shear measurements. Fine resolution helps measure the shape accurately and reduce the shear measurement systematic errors. In addition, deep \\emph{JH} photometry can track the 4000\\AA{} break of an elliptical galaxy to $z \\sim 3$ and improve photometric redshifts (\\phz{}s) as well as systematic uncertainties in the \\phz{} error distribution \\citep{abdalla08}, which has a large impact on WL constraints on the dark energy EOS \\citep{huterer06,ma06,zhan06d}. Adding \\emph{K} or $K_{\\rm dark}$ band will certainly improve galaxy \\phz{}s, especially at $z \\gtrsim 3$. However, currently planned multiband dark energy surveys use galaxies at $z \\lesssim 3$, and their concern is the confusion between $z\\lesssim 0.5$ ellipticals and $2 \\lesssim z \\lesssim 3.5$ star-forming galaxies, which is greatly mitigated by $u$ and $JH$ bands \\citep{abdalla08}. Another consideration is that Dome A is far more advantageous at $K_{\\rm dark}$ band than at $K$ band because of the low thermal background there, so $K_{\\rm dark}$ is likely to be chosen over $K$. This would leave a considerable gap in the wavelength coverage and reduce the already-small gain on \\phz{}s in the useful redshift range for dark energy investigations. Therefore, we do not discuss utilities of wavebands beyond $H$ in this paper. Nevertheless, the $K_{\\rm dark}$ band is crucial for a broad range of other sciences and will be an important aspect of the KDUST survey. This paper is organized as follows. Section~\\ref{sec:survey} discusses the survey plan of KDUST including that of its pathfinder in context of the LSST survey. We then consider a joint KDUST and LSST survey in \\autoref{sec:de} for constraining the dark energy EOS with BAO, WL, and type Ia SN techniques. The results are presented in both two-parameter space where the dark energy EOS is parameterized as $w(z)=w_0+w_az(1+z)^{-1}$ and in model-independent principle component space. The conclusion is drawn in \\autoref{sec:con}. ", "conclusions": "\\label{sec:con} Dome A offers a very competitive site for studying dark energy. Given the amount of resources required to build a large telescope and run a massive survey there, one must give the highest priority to programs that cannot be easily carried out elsewhere. Thus, a reasonable strategy is to focus on NIR imaging and collaborate with other surveys for optical data. Using LSST as an example, we show that a high-resolution 5000--10,000 deg$^2$ KDUST survey in \\emph{yJH} bands could improve LSST BAO+WL constraints on the dark energy EOS parameters $w_0$ and $w_a$ by reducing the \\phz{} and shear measurement systematics. A SNAP-like SN sample plus a large local and nearby SN sample from KDUST would further boost the DETF FOM by more than a factor of two. In addition to forecasts for the $w_0$--$w_a$ parametrization, we also apply a PCA approach to investigate the constraints on the dark energy EOS $w(z)$ in a model-independent way. We find that regarding the number of the constrained eigenmodes of $w(z)$, an ideal 10,000 deg$^2$ \\emph{ugrizyJH} survey, combined with \\emph{Planck}, can constrain 7 eigenmodes, while KDUST+LSST can allow us to constrain 3 more modes. We have not discussed dark energy probes such as strong lensing, cluster counting, and higher-order statistics of the same galaxy and shear data, which could further tighten the constraints on the dark energy EOS. Strong lensing constrains dark energy through the time delay effect as well as counting of strong lenses. It is also an excellent probe of dark matter halo structures and, hence, can be used to measure dark matter particle properties. With high-resolution imaging, one could extract more cosmological information from strong lensing observations. Therefore, Dome A could be particularly advantageous for strong lensing studies. Dark energy forecasts depend crucially on the assumed properties of the survey data, including all the systematics. Dome A has many advantages over other ground sites and has an environment close to that in space. Hence, we use well-studied LSST and SNAP as references to make crude estimates of the data for this investigation. Further work and detailed modeling are needed to give a more realistic assessment of the Dome A site for studying dark energy." }, "1005/1005.3217_arXiv.txt": { "abstract": "We analysed 866 observations of the neutron-star low-mass X-ray binary XTE J1701--462 during its 2006-2007 outburst. XTE J1701--462 is the only example so far of a source that during an outburst showed, beyond any doubt, spectral and timing characteristics both of the Z and atoll type. There are 707 RXTE observations ($\\sim 2.5$ Ms) of the source in the Z phase, and 159 in the atoll phase ($\\sim 0.5$ Ms). We found, respectively, pairs of kilohertz quasi-periodic oscillations (kHz QPOs) in 8 observations during the Z phase and single kHz QPO in 6 observations during the atoll phase. Using the shift-and-add technique we identified the QPO in the atoll phase as the lower kHz QPO. We found that the lower kHz QPO in the atoll phase has a significantly higher coherence and fractional rms amplitude than any of the kHz QPOs seen during the Z phase, and that in the same frequency range, atoll lower kHz QPOs show coherence and fractional rms amplitude, respectively, 2 and 3 times larger than the Z kHz QPOs. Out of the 707 observations in the Z phase, there is no single observation in which the kHz QPOs have a coherence or rms amplitude similar to those seen when XTE J1701--462 was in the atoll phase, even though the total exposure time was about 5 times longer in the Z than in the atoll phase. Since it is observed in the same source, the difference in QPO coherence and rms amplitude between the Z and atoll phase cannot be due to neutron-star mass, magnetic field, spin, inclination of the accretion disk, etc. If the QPO frequency is a function of the radius in the accretion disk in which it is produced, our results suggest that in XTE J1701--462 the coherence and rms amplitude are not uniquely related to this radius. Here we argue that this difference is instead due to a change in the properties of the accretion flow around the neutron star. Regardless of the precise mechanism, our result shows that effects other than the geometry of space time around the neutron star have a strong influence on the coherence and rms amplitude of the kHz QPOs, and therefore the coherence and rms amplitude of the kHz QPOs cannot be simply used to deduce the existence of the innermost stable circular orbit around a neutron star. ", "introduction": "It is now more than 13 years ago that kilohertz (kHz) quasi-periodic oscillations (QPOs) were discovered (\\citealt{v1996}; \\citealt{st1996}) in neutron star (NS) low-mass X-ray binary (LMXB) systems. Interest in this phenomenon remains high because of the close correspondence between kHz QPO timescales and typical dynamical timescales of matter orbiting close to a NS. For this reason, kHz QPOs are potential tools to probe general relativity in the strong-gravitational field regime \\citep{v2005}, and constrain the NS equation of state \\citep{mil1998}.\\\\ Since the launch of the Rossi X-Ray Timing Explorer (RXTE) in 1995, kHz QPOs have been detected in about 30 NS LMXBs (for a review see \\citealt{v2005}). Most of these sources show two simultaneous kHz QPOs, usually called the lower and the upper kHz QPO, with frequencies that can drift as a function of time in the range 250-1200 Hz (\\citealt{v2004}). Studies of these kHz QPOs show that QPO frequencies are related to other properties of the source; e.g. on short time-scales (within a day or less) QPO frequencies are well correlated with the intensity of the source, whereas on long time-scales this correlation breaks down and intensity-frequency diagrams show the so-called ``parallel tracks'' (\\citealt{m1999}). The frequencies of the kHz QPOs correlate also with the position of the source in the colour-colour diagram, and with parameters of spectral components used to describe the X-ray spectrum of these sources (\\citealt{w1997}, \\citealt{mv1999}, \\citealt{k1999}, \\citealt{ds2001}). Nevertheless, it is still unclear which physical parameters drive the QPO frequency, although there are indications that mass accretion rate, $\\dot{m}$, plays a key role \\citep*{mil1998}.\\\\ Several models have been proposed to explain the kHz QPOs (e.g., \\citealt{mlp1998}, \\citealt{sv1998}, \\citealt{a2003}), as well as the connection between high-frequency QPOs and other time variability usually present in power density spectra (for a review of variability at low frequencies see \\citealt{v2001}). Despite these efforts, there is still no single model that is able to explain in a self-consistent way all the QPO properties.\\\\ KHz QPOs are characterised by three parameters: centroid frequency $\\nu$, quality factor $Q=\\nu/FWHM$, where $FWHM$ is the full-width at half-maximum of the QPO, and fractional rms amplitude. Systematic analyses of these kHz QPO properties have been done for a large number of sources \\citep[e.g.][]{j2000,van2000,ds2001,mvf2001,h2002,van2002,dmv2003,b2005a,b2005b,a2005,m2006,bom2006}. Those studies show that, in each source the quality factor and the rms amplitude of the lower kHz QPO increase with the centroid frequency of the QPO until they reach a maximum value, after which they decrease as the frequency continues to increase (e.g. \\citealt*{mvf2001}, \\citealt*{dmv2003}, \\citealt{b2005b}; see \\citealt{m2006} for a compilation of results). The upper kHz QPO does not show the same trend; in this case the quality factor usually does not change with the centroid frequency while the rms amplitude stays more or less constant at lower frequencies and then decreases as the frequency increases (\\citealt{van2002}; \\citealt{van2003}; \\citealt{b2005a}; \\citealt{a2008}).\\\\ \\citet{b2005b} and \\citet{bom2006} interpreted the drop of the quality factor of the lower kHz QPO at high frequencies in the LMXBs 4U 1636--536 and 4U 1608--522 as a signature of the inner disk radius reaching the innermost stable circular orbit (ISCO), and starting from that assumption they estimated the mass and the radius of the compact object in these two systems. However, \\citet{m2006} argued against this idea and suggested that the drop of $Q$ and rms in individual sources might be related (at least in part) to changes of the properties of the accretion flow in these systems.\\\\ Following those results, here we investigate the properties of the kHz QPOs for the transient NS LMXB XTE J1701--462. This source was detected for the first time on 2006 January 18 with the All-Sky monitor on-board RXTE \\citep{r2006}. As reported by \\citet{h2007ate}, \\citet{l2009a}, \\citet{a2010} and \\citet{h2010}, this is the only source so far that showed both Z and atoll behaviour (for more details about the Z and atoll classes see \\citealt{hv1989}). The luminosity range covered by XTE J1701--462, from Eddington limit to quiescence, gives us a unique opportunity to study kHz QPO properties in different states and, more importantly, at different mass accretion rates in the same system, which could provide vital information to understand the origin and the mechanisms that drive the properties of these QPOs.\\\\ In section 2 we describe the observations and the data analysis, and in section 3 we present our results. In section 4 we discuss those results in the context of current ideas concerning the mechanisms behind the kHz QPOs in LMXBs, and in section 5 we summarise our conclusions. \\section[]{OBSERVATIONS AND DATA ANALYSIS} \\label{data} \\begin{table*} \\scalebox{0.87} \\centering \\begin{tabular}{c|||c|c|c|||c|c|c||} \\hline \\hline \\multicolumn{7}{c}{Z phase} \\\\ \\hline Obs ID & \\multicolumn{3}{c||}{$L_{\\ell}$} & \\multicolumn{3}{c||}{$L_{u}$}\\\\ \\hline & $Q_{\\ell}$ & rms$_{\\ell}\\%$ & $\\nu_{\\ell}$ (Hz) & $Q_{u}$ & rms$_{u}\\%$ & $\\nu_{u}$ (Hz) \\\\ \\hline 91442-01-07-09 & $35.3\\pm18.8$&$1.4\\pm0.2$&$642.7\\pm3.1$&$52.8\\pm38.8^{1}$&$<1.6$&$932.6\\pm7.9^{1}$ \\\\ \\hline 92405-01-01-02 & $22.6\\pm11.1$&$3.4\\pm0.5$&$615.1\\pm3.8$&$56.9\\pm28.8$&$2.9\\pm0.5$&$944.9\\pm3.7$\\\\ \\hline 92405-01-01-04 & $5.4\\pm2.8^{1}$&$< 3.5$&$502.4\\pm23.1^{1}$&$10\\pm3$&$3.1\\pm0.3$&$760.8\\pm6.4$ \\\\ \\hline 92405-01-02-03 & $11.9\\pm6.5$&$3.0\\pm0.5$&$623.9\\pm8.3$&$20.6\\pm11.1$&$2.7\\pm0.4$&$911.2\\pm8.4$\\\\ \\hline 92405-01-02-05 & $8.9\\pm6.9$ & $2.7\\pm0.4$& $595.9\\pm10.1$ & $6.2\\pm1.4$ & $4.1\\pm0.4$ & $850.5\\pm12.4$\\\\ \\hline 92405-01-03-05 &$8.5\\pm2.7$&$4.6\\pm0.5$&$612.5\\pm7.7$&$10.5\\pm3.5$&$4.6\\pm0.6$&$917.3\\pm10.6$\\\\ \\hline 92405-01-40-04 &$9.2\\pm3.2$&$2.8\\pm0.3$&$650.5\\pm7.1$&$11.1\\pm3.1$&$2.9\\pm0.3$&$911.0\\pm8.6$\\\\ \\hline 92405-01-40-05 &$8.1\\pm2.4$&$3.0\\pm0.3$&$637.9\\pm8.6$&$12.9\\pm3.5$&$3.0\\pm0.3$&$919.8\\pm7.6$\\\\ \\hline \\end{tabular} \\centering \\caption{Properties of the kHz QPOs detected in the Z phase of XTE J1701--462. Subscript letters $u$ and $\\ell$ denote lower and upper kHz QPOs, respectively. Parameters without errors represent 95\\% confidence level upper limits. All other errors reported represent 1$\\sigma$ confidence intervals. $^{1}$These parameters are calculated for kHz QPOs with a significance level lower than 3$\\sigma$ (see the text for precise values).} \\label{tab2} \\end{table*} We analysed all the public data of the LMXB XTE J1701--462 collected with the Proportional Counter Array (PCA) on board of RXTE (\\citealt{b1993}; \\citealt{ja2006}). There are 866 observations of this source in the RXTE archive, for a total exposure time of $\\sim$ 3 Ms. During these observations the source showed several type-I X-ray bursts that we excluded from our analysis (see \\citealt{l2009b} for a detailed analysis of the bursts). \\subsection{Timing analysis} To search for kHz QPOs, we created Leahy-normalised power density spectra using Event mode data with 125$\\mu s$ time resolution covering the full PCA energy band, nominally from 2 to 60 keV. We created Fourier power density spectra from 16-seconds data segments, using 1/4096 s time resolution such that the frequency range is defined from 0.0625 Hz to 2048 Hz. We removed detector drop-outs; no dead-time correction or subtraction of background contribution were done to calculate the power density spectra. We created one averaged power density spectrum for each observation that we visually inspected to search for the presence of QPOs with characteristic frequencies in the range from 200 Hz to 1200 Hz. We found kHz QPOs in 14 out of the 866 observations that we analysed.\\\\ Following \\citet{a2010} and \\citet{h2010}, we considered that XTE J1701--462 was in the Z phase from its discovery in January 2006 \\citep{r2006} until the end of April 2007 (as reported by \\citealt{a2010} and \\citealt{h2010}, no clear boundary between Z and atoll phase has been found which makes this date just an approximation), when it started to behave as an atoll source until it went into a quiescence state. Using this division, there are 707 observations ($\\sim 2.5$ Ms) of XTE J1701--462 in the Z phase, and 159 observations ($\\sim 0.5$ Ms) in the atoll phase. KHz QPOs were detected in individual observations only in the horizontal branch during the Z phase, and in the lower banana in the atoll phase of the outburst \\citep[see][]{h2010}. From 14 observations with kHz QPOs, 8 belong to the Z phase (ObsIDs are reported in Table \\ref{tab2}) and 6 to the atoll phase (observations: 93703-01-02-04, 93703-01-02-11, 93703-01-02-05, 93703-01-02-08, 93703-01-03-00, and 93703-01-03-02).\\\\ A quick analysis of the observations showed clear differences in the properties of the QPOs between Z and atoll phases (see \\S3 for a detailed discussion). The QPOs were always weaker and broader in the Z than in the atoll phase. It is well known that the frequency of the kHz QPOs can change over tens of Hz in time intervals of a few hundred seconds \\citep[e.g][]{b1996}, and this can artificially broaden the QPO in the averaged power spectrum of long observations. Therefore, for each observation in which we found a kHz QPO, we divided the observation in smaller intervals to check if a significant QPO was still present, and whether the QPO frequency was changing. We found that in the Z phase it was not possible to detect a significant kHz QPO in power spectra of intervals shorter than a full observation: In all observations in the Z phase with kHz QPOs the QPOs were weak and broad over short time intervals. Therefore, for the rest of the analysis, for the Z phase we report QPO properties for the average power density spectrum of each observation. In the atoll phase, the kHz QPOs were significantly detected on time intervals shorter than a full observation, and in most cases the frequency was changing in time. Therefore, for the atoll phase, we decided to average power density spectra according to the frequency of the QPO. To do this we produced power density spectra of 16s of data, and averaged up to 15 of these power spectra to get a significant detection of the kHz QPO from which we measured the QPO frequency as a function of time. Finally we averaged power density spectra such that the frequency of the QPO was within a range of 10 Hz to 60 Hz (frequency intervals are reported in Table \\ref{tab1}). We shifted the frequency scale of all these power density spectra in order to align the frequencies of the QPOs to a constant value and then averaged the power density spectra to create one power density spectrum per selection (see \\citealt{m1998}). We fitted each Z and atoll average power density spectrum in the frequency range 200-1200 Hz using a constant to model the Poisson noise plus one Lorentzian to model the kHz QPO. It must be clarified that, for the atoll phase, the centroid frequencies of the QPOs reported in Table \\ref{tab1} are the mean frequencies within the interval selected, and the errors associated are the standard deviations of the selection.\\\\ In order to label the QPOs we used the standard convention introduced by \\citet{bpv2002} where each QPO is denoted with the letter \\emph{L} with a subscript that specifies the category. In this particular case we were interested in kHz QPOs so we used $L_{\\ell}$ and $L_{u}$ to identify the lower and the upper kHz QPO, respectively. Following this criterion all the characteristics associated with one QPO have the same label.\\\\ We estimated the fractional rms amplitude of QPOs in the atoll phase in different energy bands. For each observation with QPOs we first created power density spectra in 5 different energy intervals: 2-3 keV, 3-6 keV, 6-11 keV, 11-16 keV and 16-25 keV (we stopped at 25 keV because of the lack of sensitivity of the detector above that energy). We shifted and added the power density spectra creating one single power density spectrum for each energy band and we fitted these power density spectra as the atoll power density spectra previously described. To calculate the fractional rms amplitude we calculated the integral power of the Lorentzian and we renormalised it using the source and background count rate (see \\citealt{v1997}). \\\\ \\subsection{Spectral analysis} We calculated X-ray colours and intensity of the source using the Standard 2 mode data. We defined a hard colour as the count rate ratio in the energy bands 9.7-16.0 keV and 6.0-9.7 keV, and the intensity of the source as the count rate in the energy band 2.0-16.0 keV. To obtain the exact count rate in each of these bands we interpolated in channel space.\\\\ To correct for the gain changes and differences in the effective area between the proportional counter units (PCUs) as well as differences due to changes in the channel to energy conversion of the PCUs as a function of time, we normalised our colour and intensity by the Crab Nebula values obtained close in time to our observations and per PCU (see \\citealt{k2004} and \\citealt{a2008} for details). Finally we averaged the normalised colours and intensities per PCU every 16 seconds using all available PCUs.\\\\ We also calculated the source luminosity for all 14 observations containing kHz QPOs. In order to do so, we first created a light curve for each observation, we then checked for X-ray bursts and detector drop-outs, and if any were present we eventually excluded them from the analysis. Using Standard 2 data we extracted energy spectra following the procedures described in the RXTE web page, and we added a 0.6\\% systematic error in quadrature to each channel. We fitted the energy spectra in Xspec in the energy range from 3 keV to 22 keV using a model consisting of a blackbody and a multi-colour blackbody for the Z observations, while we added also a broken power-law with the break energy fixed at 20 keV to fit the atoll observations \\citep[see][]{l2009a}. These model also include absorption from the interstellar medium toward the source and when necessary we added a Gaussian emission line at $\\sim$ 6.5 keV. The reduced $\\chi^{2}$ of our fits range from 0.6 to 1.1 (for 36 d.o.f.). From the best-fitting model we calculated the unabsorbed flux in the energy range 2-50 keV, setting the N$_{H}$ to zero. and creating an artificial response function for the full energy range 2 keV to 50 keV. We then estimated the luminosity assuming a distance $d= 8.8$ Kpc \\citep{l2009b} .\\\\ \\begin{table} \\centering \\begin{tabular}{c||c|c|c||c|} \\hline \\hline \\multicolumn{5}{c}{atoll phase}\\\\ \\hline Interval &Hz & $Q$& rms$\\%$ & $\\nu$(Hz) \\\\ \\hline 1 & 600-660 & $69.9\\pm17.1$ & $9.8\\pm0.9$ & $640.9\\pm13.5$\\\\ \\hline 2 & 660-700 & $59.6\\pm15.6$ & $10\\pm1$ & $673.3\\pm14.5$\\\\ \\hline 3 & 700-750 & $98.2\\pm14.8$ & $9.2\\pm0.5$ & $716.8\\pm6.7$\\\\ \\hline 4 & 750-780 & $67.1\\pm19.2$ & $11.8\\pm1.2$ & $772.2\\pm4.2$\\\\ \\hline 5 & 780-800 & $106.9\\pm9.7$ & $9.9\\pm0.3$ & $793.9\\pm5.4$\\\\ \\hline 6 & 800-820 & $150.3\\pm20.9$ & $8.7\\pm0.4$ & $811.5\\pm6.6$\\\\ \\hline 7 & 820-830 & $114.9\\pm12.7$ & $9.3\\pm0.4$ & $827.3\\pm2.6$\\\\ \\hline 8 & 830-840 & $93.4\\pm8.7$ & $8.9\\pm0.3$ & $835.8\\pm2.6$\\\\ \\hline 9 & 840-850 & $100.6\\pm13.9$ & $7.9\\pm0.4$ & $846.2\\pm2.9$\\\\ \\hline 10 & 850-950 & $123.9\\pm36.8$ & $6.6\\pm0.7$ & $854.8\\pm4.3$\\\\ \\hline \\end{tabular} \\caption{Properties of the kHz QPOs detected in the atoll phase of XTE J1701--462. Column 2 shows the frequency selections used to create the intervals (see text for details). Column 3, 4 and 5 show the quality factor, the fractional rms amplitude and the frequencies of the kHz QPOs, respectively. All errors represent 1$\\sigma$ confidence intervals. } \\label{tab1} \\end{table} \\section[]{RESULTS} \\label{res} In Table \\ref{tab2} and Table \\ref{tab1} we report the quality factor $Q$, fractional rms amplitude and frequency of the kHz QPOs detected in all the RXTE observations available for XTE J1701--462, from January 2006 to August 2007.\\\\ In Table \\ref{tab2} we show the properties of the kHz QPOs during the Z phase. The significance level of these QPOs ranges between 3.2$\\sigma$ and 6$\\sigma$, except for $L_{u}$ in observation 91442-01-07-09 and $L_{\\ell}$ in observation 92405-01-01-04 which have, respectively, significances of 2.4$\\sigma$ and 2.3$\\sigma$. (The significances of kHz QPOs are given as the ratio of the integral of the power of the Lorentzian used to fit the QPO divided by the negative error of the power. As shown by \\citealt{bou2010}, this probably underestimates the true significance of the QPOs). For these two kHz QPOs we report upper limits. During the Z phase the lower and upper kHz QPOs show similar quality factors, on average around 15; also the fractional rms amplitude of the two QPOs is similar, between $\\sim$ 1.4\\% and 4.5\\%.\\\\ In Table \\ref{tab1} we report the properties of the 10 QPOs detected in the frequency-selected intervals in the atoll phase. These QPOs have significances between 5$\\sigma$ and 15$\\sigma$. The QPO frequency varies from 640 Hz to 860 Hz; the quality factor changes from about 60 up to 150. The fractional rms amplitude ranges from about 7\\% to 12\\%.\\\\ \\begin{figure} \\begin{center} \\includegraphics[width=80mm]{figure1.eps} \\caption{Left panel shows the frequency of the QPOs in the atoll phase of XTE J1701--462 as a function of the hard colour of the source. Right panel show the frequency of the QPOs as a function of the intensity of the source.} \\label{hard} \\end{center} \\end{figure} \\subsection{QPO identification} While for the Z phase it is easy to label the QPOs, for the atoll phase the identification is not straightforward since there we found just one QPO in each observation. A quick inspection of the values in Table \\ref{tab1} shows similar QPO properties between different intervals, which indicates that in all cases we are probably dealing with the same QPO. To progress further we compared $Q$ and rms values with those of other LMXBs where two simultaneous kHz QPOs have been studied (\\citealt{w1997}, \\citealt{w1998}, \\citealt{j1998}, \\citealt{b2005a}, \\citealt{b2005b}, \\citealt*{bom2006}); the coherence and rms amplitude are similar to those of the lower kHz QPO in other atoll sources.\\\\ According to \\citet{b2007} and \\citet{m1999}, lower and upper kHz QPOs in atoll sources follow two different tracks in the frequency-hardness diagram. Lower kHz QPOs are found when the source has low hard colour and the centroid frequency does not seem to be correlated with hard colour (as the frequency changes the hard colour changes slightly in a restricted range). On the contrary, upper kHz QPOs are found where the source has high hard colour that decrease as the QPO frequency increases \\citep[see Figure 3 in][]{b2007}. In Figure \\ref{hard} we plot the QPO frequency as a function of the hard colour and intensity of the source. All points are concentrated within a narrow hard colour range, from 0.7 to 0.75, while the frequency ranges from 640 Hz to 860 Hz. The fact that the hard colour changes within a restricted range while the frequency moves in a range of about 200 Hz resembles what \\citet{b2007} show concerning the lower kHz QPO behaviour in the atoll source 4U 1636--53. This further supports our identification of the kHz QPOs in the atoll phase as lower kHz QPOs.\\\\ In order to understand the frequency-hardness diagram, we plot in the right panel of Figure \\ref{hard} the QPO frequency as a function of the source intensity. It is apparent from the graph that the data are divided into two different tracks characterised by intensity values which differ by a factor of $\\sim 2$. That trend resembles the so-called ``parallel tracks'' phenomenon which has been seen in many LMXBs \\citep[see e.g.][]{m1999}. Although the points in Figure \\ref{hard} are the result of frequency selection, it turns out that all the QPOs we combined to make the intervals from 1 to 4 and from 5 to 10 on Table \\ref{tab1} are concentrated in a time interval, respectively, of about 1 day and less than half a day. Moreover, those two groups are separated by more or less 4 days, which means the points in the right panel of Figure \\ref{hard} can in fact reflect the ``parallel tracks'' phenomenon.\\\\ \\begin{figure} \\begin{center} \\includegraphics[width=80mm]{figure2.eps} \\caption{Power density spectrum of all the atoll observation where QPOs are detected. This power spectrum is the result of the shift-and-add method. Two kHz QPOs are visible, the one at lower frequency (30 $\\sigma$ significant) was the one originally used ti shift and add the power spectra, while the one at higher frequencies (3.1$\\sigma$ significant, single trial) appears as result of the application of the method. } \\label{double} \\end{center} \\end{figure} We notice that QPOs at the highest intensities correspond to the ones with the highest frequencies in the frequency-hardness diagram, while QPOs at the lowest intensities correspond to those with the lowest frequencies in the frequency-hardness diagram. This can be interpreted as ``parallel tracks'' in the frequency-hardness diagram. A similar trend was seen in the LMXB 4U 1636--53 (see Figure 2 in \\citealt*{dmv2003}).\\\\ To search for a second (possibly weaker) kHz QPO in the atoll phase we apply the shift-and-add method \\citep{m1998}. As we described in section \\ref{data}, first we fit the centroid frequency of the kHz QPO in each individual atoll power density spectrum, then we shift all the power spectra such that the kHz QPO frequencies are aligned at the same value. Finally we average all the data into a single power density spectrum which we fit with a constant plus Lorentzian component for each QPO. This method corrects the frequency drift in time of the kHz QPOs, increasing their signal to noise ratio and their significance. If the second kHz QPO is at a (more or less) fixed distance from the first, this method adds all the data in a way that increases the significance of the second QPO, which may then become detectable.\\\\ The result of this procedure is presented in Figure \\ref{double}, which shows two simultaneous kHz QPOs, the one at lower frequency is the kHz QPO which we already knew, while the second one becomes detectable as a result of the shift-and-add method. The significance level of the second, upper, kHz QPO is 3.1$\\sigma$, which implies a marginal detection. Note however that the frequency difference between the kHz QPOs is $258\\pm13$ Hz, which is consistent with the value in the Z phase (see Table \\ref{tab2} and \\citealt{h2007}) and therefore there are no other trials involved in estimating this significance. As a result of this analysis, taking into account all the caveats, we find a pair of simultaneous kHz QPOs in the atoll phase of the XTE J1701--462 which strengthens our previous suggestion that the strong kHz QPOs detected in this phase are always the lower kHz QPO.\\\\ As described in section \\ref{data}, we calculate the fractional rms amplitude of the atoll QPOs in 5 different energy bands, 2-3 keV, 3-6 keV, 6-11 keV, 11-16 keV and 16-25 keV, respectively. We find fractional rms amplitudes of less than 10.9\\%, $6.9\\pm0.3$, $11.4\\pm0.2$, $17.5\\pm1.2$ and less than 21.3\\%, respectively (upper limits are given at 95\\% confidence). Those results, as already noticed in other sources (see Berger et al. 1996), show that the strength of the variability increases as the energy increases.\\\\ \\subsection{Amplitude and coherence of the kHz QPOs in the Z and atoll phases} Figure \\ref{Q_nu} shows the values of the QPO quality factor as a function of frequency. Black points represent measurements in the atoll phase while grey points are measurements in the Z phase. For the Z phase, empty and filled square symbols represent, respectively, lower and upper kHz QPOs. From the plot we notice a few interesting aspects: 1) QPOs in the atoll phase are on average 10 times more coherent than in the Z phase; 2) in the atoll phase $Q$ increases as the frequency increases, reaching a maximum value of about 150 at $\\sim$ 810 Hz; within the errors the behaviour is consistent with what has been observed in other sources \\citep[see][]{b2005a}. No such trend seems to be present in the Z phase; 3) comparing lower kHz QPOs in both phases it is clear that in the Z phase the lower kHz QPO appears at lower frequencies than in the atoll phase; 4) there is an overlap in frequency around 640 Hz between Z and atoll QPOs with a significant mismatch in $Q$.\\\\ Figure \\ref{r_nu} shows the QPO rms fractional amplitude as a function of frequency. As in Figure \\ref{Q_nu}, black points denote measurements in the atoll phase, while grey points are QPOs in the Z phase. The rms amplitude of the lower kHz QPO in the atoll phase remains more or less constant around 10\\% as the QPO frequency increases from 640 Hz to 780 Hz and then drops rapidly at $\\sim$ 800 Hz. There is no evidence of a similar trend for the rms of the lower or upper kHz QPOs in the Z phase. As already noticed for the quality factor, the rms amplitude values show a clear difference between the atoll and Z phases: The lower kHz QPOs amplitude in the atoll phase is on average a factor of 2 higher than in the Z phase; this difference is also apparent in the region where atoll and Z QPOs overlap in frequency between 620 Hz and 660 Hz.\\\\ We test whether we may have missed any significant kHz QPO in our analysis. We calculate upper limits to the fractional rms amplitude in frequency ranges where we did not find kHz QPOs. To do so we take two power density spectra, one for each source phase, where no kHz QPOs were found. We fit those power density spectra using a model consisting of a constant to fit the Poissonian noise, and a Lorentzian to fit the QPO with fixed values for the centroid frequency and the quality factor $Q$. We fit the data and we estimate the upper limit of the amplitude of the Lorentzian using $\\Delta \\chi^{2}=2.7$, which corresponds to 95\\% confidence level. We repeat the same procedure shifting gradually the frequency of the Lorentzian to higher values until we cover the frequency range of interest.\\\\ For the analysis of atoll observation we use power density spectra of 256s, comparable to the time intervals over which we detected kHz QPOs in the atoll phase. We use two different quality factors, $Q=50$ and $Q=20$, which are comparable to the values we found for the kHz QPOs at low frequencies (see Table \\ref{tab1}). We find upper limits to the fractional rms amplitude that vary within the range 2.7\\%-5.2\\% and within the range 3\\%-7\\%, respectively for $Q=20$ and $Q=50$. From these values it appears that the rms amplitude of the lower kHz QPO in the atoll phase should decrease at frequencies lower than 640 Hz. This would be in agreement with the typical rms-frequency trend observed in most of the LMXBs (\\citealt*{mvf2001}; \\citealt{van2000, van2002, van2003}; \\citealt{ds2001}; \\citealt*{dmv2003}; \\citealt{b2005a}).\\\\ For the Z observation, we estimate the upper limits to the amplitude of a kHz QPO with typical atoll properties within short time intervals to check whether we could have missed such a QPO during the analysis. We calculate the upper limits in the frequency range 650-940 Hz assuming a kHz QPO with $Q=100$ and setting 3 values of the integration time, 128, 256 and 512 seconds. We find that the upper limits to the fractional rms amplitude vary between 1\\% and 4\\% for 128 seconds of integration time, between 0.8\\% and 4\\% for 256 seconds, and between 0.4\\% and 3.4\\% for 512 seconds. From this we conclude that if a kHz QPO with $Q=100$ and fractional rms amplitude similar to what we found for the atoll kHz QPOs was present in the Z observations, we would have detected it significantly in intervals as short as 128s.\\\\ ", "conclusions": "The transient source XTE J1701--462 is so far the only accreting neutron-star X-ray binary that changed from a Z into an atoll source during an outburst. Here we found that the properties of the kHz QPOs in XTE J1701--462 in the atoll and Z phases are significantly different: At approximately the same frequency (about 640 Hz), the coherence of the kHz QPOs is a factor of 2 larger, and the rms amplitude is a factor of 3 larger, in the atoll than in the Z phase (see Figure 3 and Figure 4, respectively). Furthermore, out of 707 observations in the Z phase, there is no single one in which the kHz QPOs have a coherence or rms amplitude similar to those seen when XTE J1701--462 was in the atoll phase, even though the total exposure time was about 5 times longer in the Z than in the atoll phase. If the kHz QPOs reflect the motion of matter in the accretion disk, and the QPO centroid frequency is related to the radius in the accretion disk where this motion takes place, our results show that, at least in XTE J1701--462 there is no unique relation between the radius of the disk at which the QPO is produced and the coherence and rms amplitude of the QPO.\\\\ It was already known that the kHz QPOs are broader and weaker in Z than in atoll sources (see M\\'endez 2006 for a compilation), but this is the first time that the same trend is observed in a single source. This result conclusively excludes things like neutron-star mass, magnetic field, spin, inclination of the accretion disk, etc., as the cause of this trend, since these parameters cannot change on time scales of one and half year. As we discuss below, the most likely reason for the difference of QPO coherence and rms amplitude between the Z and atoll phase in XTE J1701--462 is a change in the properties of the accretion flow around the neutron star where the QPOs are produced. Despite the the lack of information about the precise mechanism, our result shows that effects other than the geometry of space time around the neutron star have a strong effect on the QPO properties. If, as suggested by \\citet{b2005a, b2005c}, the ISCO affects the coherence and rms amplitude of the kHz QPOs, our result shows that there are other mechanisms that should also be taken into account to explain the trend seen in the data. For instance, in XTE J1701--462 the coherence and rms amplitude differ by a factor of $\\sim 3$ to 10 between the Z and atoll phases similar, for instance, to the change of coherence and rms amplitude that has been proposed to be due to the ISCO in 4U 1636--53 \\citet{b2005b}.\\\\ \\citet{m2006} found that the the behaviour of the coherence and rms amplitude of the kHz QPOs in individual sources is similar to the behaviour of the maximum coherence and maximum rms of the kHz QPOs as a function of the luminosity in a sample of 12 NS-LMXBs. It is interesting to test to what extent this luminosity relation holds for the Z and atoll phases of XTE J1701--462, which are well separated in luminosity. In Figure \\ref{max} we combine the data points from the Z and atoll phases of XTE J1701--462 with those of the 12 NS-LMXBs studied by \\citet{m2006}.\\\\ \\begin{figure*} \\begin{center} \\includegraphics[angle=-90,width=120mm]{figure3.eps} \\caption{Quality factor $Q$ as a function of the QPO centroid frequency for XTE J1701--462. Black points represent QPOs in the atoll phase, and grey points represent QPOs in the Z phase. Empty and filled squares indicate lower and upper kHz QPOs, respectively. The two values of the coherence at frequencies $\\sim$500 Hz and $\\sim$930 Hz correspond to kHz QPOs with significances 2.3$\\sigma$ and 2.4$\\sigma$, respectively (see text).} \\label{Q_nu} \\end{center} \\end{figure*} \\begin{figure*} \\begin{center} \\includegraphics[angle=-90,width=120mm]{figure4.eps} \\caption{Factional rms amplitude as a function of the QPO centroid frequency for XTE J1701--462. Black points represent QPOs in the atoll phase, and grey points represent QPOs in the Z phase. Empty and filled squares indicate lower and upper kHz QPOs, respectively. Arrows correspond to 95\\% confidence level upper limits.} \\label{r_nu} \\end{center} \\end{figure*} The upper and the lower panels show respectively the maximum quality factor and the maximum rms amplitude of the lower kHz QPO for 13 sources as a function of the luminosity of the source in the 2-50 keV range, normalised by the Eddington luminosity, L$_{Edd}=2.5\\,\\times10^{38}$ erg s$^{-1}$, corresponding to a neutron star of 1.9 M$_{\\odot}$ accreting gas with cosmic abundance. As it is apparent from the plots, XTE J1701--462 follows the trends already traced by the other sources; moreover the coherence of the kHz QPOs in the Z phase fill the gap between the atoll sources in the left-hand side and the Z sources in the right-hand side of the graph, strengthening the correlation. (Notice, however that the luminosity of XTE J1701--462 could be uncertain by up to a factor of $\\sim$ 2, \\citealt{l2009a}).\\\\ Besides differences of the kHz QPOs properties between Z and atoll sources, Figure \\ref{max} shows also a trend of both $Q$ and rms amplitude within the atoll sources. This is noticeable in the lower panel of Figure 5 of \\citet{m2006}, where the maximum coherence and maximum rms amplitude of the lower kHz QPO are plotted vs. each other. From that plot it is apparent that the two quantities both in Z and atoll sources follow the same correlation \\citep[see][]{m2006}. This suggests that there is a single mechanism behind this trend. Our observations of XTE J170--462 in the Z and atoll phase are in line with this.\\\\ We can qualitatively explain how it is possible to find high rms amplitudes of kHz QPOs at energies where the contribution of the disk is insignificant. As reported by \\citet{b1996}, the rms amplitude of the lower kHz QPO in the LMXB 4U 1608--52 increases with energy up to 20\\% (fraction of the total emitted flux) at energies around 30 keV, significantly above the energy range where the accretion disk contributes to the X-ray spectrum (see also \\citealt*{mvf2001}; \\citealt{g2003}; \\citealt{a2008}). If the kHz QPOs are produced in the accretion disk, this must imply the presence of a mechanism that amplifies the variability at different energy bands. Some mechanisms have been proposed for this: \\citet{lm1998} found that oscillations of the density and temperature of the Comptonizing medium can reproduce the rms amplitude behaviour of the lower kHz QPO in the atoll source 4U 1608--52. \\citet*{g2003}, from the analysis of the Z source GX340+0, suggest that QPOs are related to the contribution of the boundary layer emission to the total source emission. \\citet{gr2005} found that as $\\dot{m}$ increases from the horizontal branch to the flaring branch along the Z-shaped track in the colour-colour diagram of GX 340+0, the contribution of the boundary layer decreases. If we combine the results of \\citet*{g2003} and \\citet{gr2005}, and we apply them to the case of XTE J1701--462, we would expect the boundary-layer contribution to be stronger in the atoll phase than the in Z phase. To verify this, we used the spectral analysis of XTE J1701--462 done by \\citet{l2009a}; their Figures 14 and 15 show the spectral fitting results, respectively, of the atoll and the Z phase. From their Figure 14 we notice that in the atoll phase the blackbody (BB) component used by \\citet{l2009a} to fit the boundary-layer emission becomes dominant, while the emission from the disk (fitted with a multi colour disk blackbody, MCD) becomes negligible. Comparing this with their Figure 15, we found that the fractional contribution of the boundary-layer emission in the atoll phase is higher than in the Z phase, which is in agreement with the results of \\citet*{g2003}, and can also explain the fact that we found stronger kHz QPOs in the atoll phase than in the Z phase. If this is correct, we should expect that most of the variability concentrates in the energy band where the boundary-layer emission peaks. From Figure 14 of \\citet{l2009a}, the BB temperature is about 2 keV, which implies that the peak of the boundary-layer emission is at about 10 keV. According to the results shown in the previous section, the atoll fractional rms amplitude increases as the energy increases, up to about 20\\% in the energy band 16-25 keV. We test whether the fractional rms amplitude we found above 10 keV is compatible with the fractional contribution of the boundary layer emission to the total emission of the source. From the spectral fits shown by \\citet[][see their Figure 12 panel ``atoll SS'']{l2009a}, the BB contributes more than 30\\% of the total emission in the 10-25 keV band, which means that the picture where the amplitude and coherence of the kHz QPOs are driven by the boundary layer is still consistent at those energies. Further studies should address the fact the kHz QPOs are only detected over narrow regions in the colour-colour diagram.\\\\ \\begin{figure} \\begin{center} {\\includegraphics[width=60mm,angle=-90]{figure5.eps}} \\caption{Upper panel: the maximum fractional rms amplitude of the lower kHz QPO for a sample of 12 sources (filled circles, see Table 1 \\citealt{m2006} for the source names) plus XTE J1701--462 as a function of the luminosity of the source. The diamond and the triangle represent, respectively, measurements in the atoll and the Z phase. Lower panel: the maximum quality factor of the lower kHz QPO for the same sources mentioned above as a function of the luminosity of the source. The symbols are the same as in the upper panel. The luminosity is in units of the Eddington luminosity for a 1.9 M$_{\\odot}$ neutron star.} \\label{max} \\end{center} \\end{figure} Adding all these together, we suggest a possible scenario to describe the behaviour of the properties of the kHz QPO. Mathematically a Lorentzian (which is usually used to model quasi-periodic oscillations) is the Fourier transform of an exponentially damped sinusoid signal, $A\\,e^{-t/\\tau} sin (2\\pi \\nu t+\\phi)$, where $A$ is the amplitude of the signal, $\\tau$ is the life time of the variability, $\\nu$ is the frequency of the oscillation, and $\\phi$ is an arbitrary phase. Starting from this mathematical expression we can build a qualitative mechanism to describe the behaviour of the quality factor and the rms of the kHz QPO. Most of the models proposed to explain the kHz QPOs, consider the disk as the most probable location where those QPOs are generated (e.g. \\citealt*{mlp1998}). Under this assumption, the above expression could represent the oscillator that generates the QPO in the disk and sets its frequency $\\nu$. Since the rms amplitude of kHz QPOs depends on energy, we should introduce an energy-dependent amplification factor $B(E)$ which should be physically related to the mechanism which generates high-energy photons. For example, this could be related to the properties of the Comptonizing medium \\citep{lm1998} or to the contribution of the boundary-layer emission to the total emission of the source \\citep*{g2003, gr2005}. The life time $\\tau$ drives the QPO width, and as the amplitude, also this could be energy dependent. Additionally, we should consider that each process that amplifies the variability could also modify the life time of the variability. Just to give an example, if the QPO is created in the disk, and later on the QPO photons interact with matter in the corona, the final width of the QPO will be the result from the combination of the two process. In our model this would imply a lifetime for the QPO $1/\\tau_{tot}=(\\tau_{disk}+\\tau_{cor})/(\\tau_{disk}*\\tau_{cor})$. With these considerations, the modulated flux in the kHz QPO could be described by the expression $B(E)*A\\,e^{-t/\\tau_{tot}}\\,sin(2\\pi \\nu t+\\phi)$. This simple expression describes the rms amplitude and coherence of the QPO in terms of the quantities $B(E)$ and $\\tau_{tot}$, which in turn could depend on the properties of the boundary layer. The above expression would reproduce the behaviour of the rms amplitude and coherence if, for instance, $B(E)$ and $\\tau_{tot}$ depended upon mass accretion rate, $\\dot m$, such that $B(E)$ and $\\tau_{tot}$ decreased as $\\dot m$ increased. The previous description does not resolve however the issue of which processes are involved, or what the key mechanisms are that create or amplify the QPO signal. Nevertheless, this qualitative explanation provides a starting point to build a realistic model to explain the kHz QPO properties.\\\\ For most of the considerations in this paper we assumed that the only kHz QPO in the atoll phase is the lower kHz QPO. Although it is very unlikely (see section \\ref{res}), if it instead was the upper kHz QPO, one would have to explain the fact that in the frequency range, from 740 Hz to 860 Hz, the atoll kHz QPOs show $Q$ and rms values significantly different than those of the Z phase kHz QPOs (Figures \\ref{Q_nu} and \\ref{r_nu}).\\\\ From the INTEGRAL catalog (INTEGRAL general reference catalog, Version 30) we found no other sources within 1 square degree of the position of XTE J1701--462. The closest source reported (about 2.3 degrees away from XTE J1701--462) is the Gamma-ray source 2EGS J1653-4604 that does not show X-ray emission. \\citet{kr2006} observed XTE J1701--462 in outburst with the Chandra's High-Resolution Camera (HRC-S) in timing mode and they did not find evidence of other sources in the HRC-S field of view ($6' \\times 90'$). Swift observations with the X-ray telescope as well as XMM Newton observations of XTE J1701--462 did not show other sources, respectively within $23.6' \\times 23.6'$ and $30' \\times 30'$. Finally we checked the RXTE Galactic Bulge scans archive and we found that with a spatial accuracy of $15'$ no sources have been detected in a 1 square degree area from the XTE 1701-462 position. The evidence just discussed indicates that most likely the outburst observed by RXTE was due to a single source, XTE J1701--462, switching from Z to atoll.\\\\ Another interesting aspect to mention is the mismatch of the frequency of the lower kHz QPOs between the two different phases of the source. It is clear from Tables \\ref{tab2} and \\ref{tab1} and Figures \\ref{Q_nu} and \\ref{r_nu} that Z phase lower kHz QPOs are in the range $500-660$ Hz, while in the atoll phase the lower kHz QPOs are in the frequency range $640-860$ Hz. Although a similar effect has been observed when one compares other atoll and Z sources (see \\citealt{m2006} for a compilation) , we have now shown this effect in a single source. Studying the energy spectra in different parts of the colour-colour diagram, \\citet{l2009a} found that in the Z phase the disk is truncated far from the NS surface and its inner radius could be set by the local Eddington limit, while in the atoll phase the disk extends closer to the NS. Different sizes of the inner radius of the accretion disk could explain the different range of kHz QPO frequencies. However we can not discard the presence of other lower kHz QPOs in the Z phase at frequencies higher than what we found. From the upper limits for the Z phase reported in section \\ref{res}, it is apparent that we are not sensitive enough to detect broad and weak QPOs at high frequencies in that phase." }, "1005/1005.2950_arXiv.txt": { "abstract": "{} {We report new Very Long Baseline Array (VLBA) polarimetric observations of the Compact Steep-Spectrum (CSS) sources 3C\\,119, 3C\\,318, and 3C\\,343 at 5 and 8.4~GHz. } {By using multifrequency VLBA observations we have derived milliarcsecond-resolution images of the total intensity, polarisation, and rotation measure ($RM$) distributions. } { The CSS source 3C\\,119, associated with a possible quasar, has source rest-frame $RM$ values up to $\\sim$10200 rad m$^{-2}$ in a region which coincides with a change in the direction of the inner jet. This component is located $\\sim$325 pc from the core, which is variable and has a peaked radio spectrum. In the case of 3C\\,318, associated with a galaxy, a rest-frame $RM$ of $\\sim$3030 rad m$^{-2}$ has been estimated for the brightest component which contributes almost all of the polarised emission. Further, two more extended components have been detected, clearly showing ``wiggles'' in the jet towards the southern side of the source. The CSS source 3C\\,343 contains two peaks of emission and a curved jet embedded in more diffuse emission. It exhibits complex field directions near the emission peaks, which indicate rest-frame $RM$ values in excess of $\\approx$6000 rad m$^{-2}$. The locations of the cores in 3C\\,318 and 3C\\,343 are not clear. } {The available data on mas-scale rest-frame $RM$ estimates for CSS sources show that these have a wide range of values extending up to $\\sim$40000 rad m$^{-2}$ in the central region of OQ172, and could be located at projected distances from the core of up to $\\sim$1600 pc, as in 3C\\,43 where this feature has a rest-frame $RM$ of $\\sim$14000 rad m$^{-2}$. $RM$ estimates for cores in core-dominated radio sources indicate that in addition to responding to an overall density gradient of the magneto-ionic medium, geometry, orientation and modes of fuelling may also play a significant role. In addition to these effects, the high values of $RM$ in CSS sources are possibly due to dense clouds of gas interacting with the radio jets. The observed distortions in the radio structures of many CSS sources are consistent with this interpretation. } ", "introduction": "The number of Compact Steep-Spectrum (CSS) sources with detailed polarimetric information available at milliarcsecond resolution is still small. Polarised radio emission from CSS radio galaxies is either very weak or below the detection limits at centimetre wavelengths. In contrast, CSS quasars show linear polarisation percentages of up to 10\\% above 1\\,GHz \\citep [ ] [and references therein] {Rossetti08}. We have conducted a series of observations of CSS sources having significantly polarised emission and high values of rotation measure ({\\it RM}) using the Very Long Baseline Array (VLBA). CSS objects are {\\it young} radio sources, with ages $<10^{{\\rm 3-5}}$\\,yr. They have linear sizes $\\leq 20$\\,kpc \\footnote {$H_0=71\\,{\\rm km}\\, {\\rm s}^{-1}\\, {\\rm Mpc}^{-1}, \\Omega_{\\rm m}=0.27, \\Omega_{\\rm vac}=0.73$} and steep high-frequency radio spectra ($\\alpha >0.5$; ${\\rm S}_{\\nu}\\propto\\nu^{-\\alpha}$). Being sub-galactic in size, CSS sources reside deep within their host galaxies. Therefore, Faraday rotation effects are to be expected when their polarised synchrotron emission is observed through the host galaxy magneto-ionic interstellar medium (ISM). The comparison of polarised emission over a range of wavelengths is an important diagnostic of the physical conditions within and around these compact radio sources (see Cotton et al. 2003c for an overview). Existing sub-arcsec polarimetry has provided evidence in favour of the interaction of components of CSSs with dense clouds of gas, as seen for example in the CSS quasar 3C\\,147 \\citep {Junor99a}. Results for the first two CSS quasars observed in our on-going program, \\object{B$0548+165$} and \\object{B$1524-136$}, are available in \\citet{Mantovani02}, while those for \\object{3C\\,43} (B$0127+233$) are to be found in \\citet{Mantovani03}. More recently, the results for \\object{3C\\,147} (\\object{B$0538+498$}) have been presented by ~\\citet{Rossetti09}. These sources have all been imaged with milliarcsecond resolution by means of full-Stokes VLBA observations. In this paper, we report on multi-frequency VLBA, plus a single Very Large Array (VLA) antenna, polarisation observations at 5 and 8.4\\,GHz for \\object{3C\\,119} (\\object{B$0429+415$}), \\object{3C\\,318} (\\object{B$1517+204$}), and \\object{3C\\,343} (\\object{B$1634+628$}). In Section \\ref{sec:observation} we summarise the observations and data processing. Section \\ref{sec:sources} describes the new information obtained on the structural and polarisation properties of \\object{3C\\,119, 3C\\,318, and 3C\\,343}. Discussion and conclusions are presented in Sections \\ref{sec:discussion} and \\ref{sec:conclusions} respectively. ", "conclusions": "\\label{sec:conclusions} We have presented multi-frequency VLBA polarisation observations of three CSS sources, namely 3C\\,119, 3C\\,318 and 3C\\,343 to estimate their {\\it RM} values. The radio source 3C\\,119 is associated with a possible quasar, and its $RM$ in the source rest frame has been found to be as high as $\\sim$10200 rad m$^{-2}$ in a region which coincides with a change in direction of the inner jet. This component is located at a projected distance of $\\sim$325 pc from the core, which is almost point-like, variable, has a peaked radio spectrum and is at best marginally polarised. 3C\\,318 is associated with a radio galaxy and its rest frame {\\it RM} has been found to reach a maximum of $\\sim$3030 rad m$^{-2}$ for the brightest component which contributes almost all of the polarised emission. These observations are more sensitive than those of Spencer et al. (1991) and have detected two more extended components, which trace ``wiggles'' in the jet towards the southern side of the source. Of the three, the CSS source 3C\\,343 has perhaps the most complex structure. It contains two peaks of emission and a curved jet embedded in more diffuse emission. It exhibits complex field directions near the peaks of emission, which indicate rest frame $RM$ values in excess of $\\approx$4000 rad m$^{-2}$. The varying sensitivity for the different frequencies and the complex field patterns near the peaks of emission make it difficult to construct a reliable {\\it RM} image for this source. We have compiled the available data on mas-scale $RM$ estimates for CSS sources. These show a wide range of values with indications of a low {\\it RM} for 3C\\,287 which needs to be confirmed from observations with a larger number of frequencies, to values as high as $\\approx$40000 rad m$^{-2}$ in the central region of OQ172 (Udomprasert et al. 1997). The components with high {\\it RM} can also occur at considerable distances from the core, e.g. in 3C\\,43 where the component with an {\\it RM} of $\\sim$14000 rad m$^{-2}$ is located at a projected distance of $\\sim$1600 pc (Cotton et al. 2003a). {\\it RM} estimates for flat-spectrum cores in largely core-dominated radio sources appear to increase with frequency (see O'Sullivan \\& Gabuzda 2009), suggesting that as one probes deeper into the core or unresolved base of the jet, one samples regions of higher density and/or magnetic field in the magneto-ionic medium. On larger scales, the jet {\\it RM}s tend to be low as these objects are observed along a line of sight where the magneto-ionic medium may have been swept out by the relativistic jets (e.g. Saikia et al. 1998; Taylor 2000; Zavala \\& Taylor 2004). The CSS objects for which {\\it RM} values have been estimated are almost entirely quasars. While the effects of an overall density gradient in the magneto-ionic medium, along with effects of geometry, orientation {\\bf and modes of fuelling of the AGN} , are likely to play a significant role, the high $RM$ values in many of these CSS sources appear to be due to dense clouds of gas interacting with the radio jets. Usually, they also exhibit large structural bends and distortions, consistent with the possibility of jet-cloud interactions in the interstellar medium of the host galaxy. Some of these gas clouds may also be responsible for fuelling the AGN activity." }, "1005/1005.5447.txt": { "abstract": "A new fully quantum method describing penetration of packet from internal well outside with its tunneling through the barrier of arbitrary shape used in problems of quantum cosmology, is presented. The method allows to determine amplitudes of wave function, penetrability $T_{\\rm bar}$ and reflection $R_{\\rm bar}$ relatively the barrier (accuracy of the method: $|T_{\\rm bar}+R_{\\rm bar}-1| < 1 \\cdot 10^{-15}$), coefficient of penetration (i.~e. probability of the packet to penetrate from the internal well outside with its tunneling), coefficient of oscillations (describing oscillating behavior of the packet inside the internal well). % Using the method, evolution of universe in the closed Friedmann--Robertson--Walker model with quantization in presence of positive cosmological constant, radiation and component of generalize Chaplygin gas is studied. It is established (for the first time): (1) oscillating dependence of the penetrability on localization of start of the packet; (2) presence of resonant values of energy of radiation $E_{\\rm rad}$, at which the coefficient of penetration increases strongly. % From analysis of these results it follows: (1) necessity to introduce initial condition into both non-stationary, and stationary quantum models; (2) presence of some definite values for the scale factor $a$, where start of expansion of universe is the most probable; (3) during expansion of universe in the initial stage its radius is changed not continuously, but passes consequently through definite discrete values and tends to continuous spectrum in latter time. % % (4) \u00f1\u00eb\u00e5\u00e4\u00f1\u00f2\u00e2\u00e8\u00e5\u00ec \u00ea\u00e2\u00e0\u00ed\u00f2\u00ee\u00e2\u00e0\u00ed\u00e8\u00ff \u00ea\u00ee\u00f1\u00ec\u00ee\u00eb\u00ee\u00e3\u00e8\u00f7\u00e5\u00f1\u00ea\u00ee\u00e9 \u00ec\u00ee\u00e4\u00e5\u00eb\u00e8 \u00ff\u00e2\u00eb\u00ff\u00e5\u00f2\u00f1\u00ff \u00e4\u00e8\u00f1\u00ea\u00f0\u00e5\u00f2\u00ed\u00ee\u00f1\u00f2\u00fc \u00ef\u00f0\u00ee\u00f1\u00f2\u00f0\u00e0\u00ed\u00f1\u00f2\u00e2\u00e0-\u00e2\u00f0\u00e5\u00ec\u00e5\u00ed\u00e8, \u00ef\u00f0\u00ee\u00ff\u00e2\u00eb\u00ff\u00fe\u00f9\u00e0\u00ff\u00f1\u00ff \u00ed\u00e0\u00e8\u00e1\u00ee\u00eb\u00e5\u00e5 % \u00f1\u00e8\u00eb\u00fc\u00ed\u00ee \u00e2 \u00ef\u00e5\u00f0\u00e2\u00ee\u00e9 \u00f1\u00f2\u00e0\u00e4\u00e8\u00e8 \u00f0\u00e0\u00f1\u00f8\u00e8\u00f0\u00e5\u00ed\u00e8\u00ff \u00e2\u00f1\u00e5\u00eb\u00e5\u00ed\u00ed\u00ee\u00e9. ", "introduction": "} If we analyzed existed variety of quantum models which describe formation of the universe and its subsequent evolution in the first stage, we should come to conclusion that the semiclassical approach for description of tunneling and determination of wave function is the most prevailing today. This approach forms a basis, which props up both models with the Feynman formalism of path integrals in multidimensional space-time, developed by the Cambridge group and other researchers, called the \\emph{``Hartle--Hawking method''} (for example, see Ref.~\\cite{Hartle.1983.PRD}), and methods based on direct consideration of tunneling in 4-dimensional Euclidian space-time called the \\emph{``Vilenkin method''} (for example, see Refs.~\\cite{Vilenkin.1982.PLB,Vilenkin.1983.PRD,Vilenkin.1984.PRD,Vilenkin.1986.PRD,Vilenkin.1988.PRD,% Vilenkin.1994.PRD,Vilenkin.1995,Vilenkin.1997.PRD.Comments,Vilenkin.2003.PRD}). The models in 4-dimensional space-time directed on description of inflation, on study of fluctuations of vacuum with inclusion of massive fields (for example, see~Refs.\\cite{Bouhmadi-Lopez.2002.PRD}), multidimensional cosmological models (for example, see~Refs.\\cite{Kolb.1990,Carugno.1996.PRD}), variety of string models (for example, see~Refs.\\cite{Brustein.2006.PRD,Kobakhidze.2007.EPJC}) have mainly such a semiclassical grounds. To date, this basis is supposed to be sufficiently reliable and give interested quantum characteristics of the universe with good accuracy. Such a point of view so has been taken root and is prevailing, that, in spite of almost 40 years of researches in quantum cosmology (see~the first papers~\\cite{DeWitt.1967,Wheeler.1968}), the papers devoted on more tiny and deeper study of possible quantum nature of the universe, its formation and evolution on the first stage, can be found enough rarely (for example, see \\cite{Rubakov.2002.PRD,AcacioDeBarros.2007.PRD,Monerat.2007.PRD}, also~\\cite{Yurov.2004.TMP,Garcia.2006.IJTP}). However, one should ask whether the penetrability determined according to the semiclassical theory by a shape of the barrier solely between two turning points, gives exhaustive answer and the best estimations of rates of evolution of universe. % \\vspace{2mm} (1) If, despite such widespread confidence in the semiclassical approach, we still want to check it, we immediately will miss some of the parameters. For example, it would seem, one can use a test of $T + R = 1$ (where $T$ and $R$ are penetrability and reflection relatively the barrier in the cosmological problem), where usually we have no doubt. However, one should recall that in quantum mechanics, the semiclassical approximation neglects the reflected waves completely (see~\\cite{Landau.v3.1989}, eq.~(46.10), p.~205, also p.~221--222) and, therefore, to compare the calculated penetrability $T$ is just not with anything. % \\vspace{2mm} (2) If we still wanted to determine the reflection coefficient, then we should increase the order of approximation of the semiclassical method (in order to take into account the decreasing component of the wave function on the background of increasing one in the region of tunneling), and just stumble on the next problem --- presence of a non-zero interference between the incident and reflected waves in the radial task. Now the criterion $T + R = 1$ for testing is not satisfied and it needs to take into account the third component $M$ of interference in addition (see~\\cite{Maydanyuk.2010.IJMPD}). In particular, at unsuccessfully chosen separation of the exactly known full wave function on the incident and reflected waves (but the semiclassical approaches have no needed apparatus for such analysis), the interference component can increase without limit, and be substantially larger than the penetrability and reflection. After the appearance of such an arbitrariness, the penetrability and reflection can freely exceed the unit and increase without limit. In what is now the general meaning of the penetrability? % \\vspace{2mm} (3) We shall give only some easy examples from quantum mechanics. (i) If we consider two-dimensional penetration of the packet through the simplest rectangle barrier (with finite size), we shall see that the penetrability is directly dependent on direction of tunneling of the packet. So, the penetrability is not a single value but the function. (ii) If to consider one-dimensional tunneling of the packet through the simplest rectangular barrier, one can obtain ``interference picture'' of its amplitude in the transmitted region, which is depended on time and space coordinates and is an exact analytical solution. Of course, the stationary part of such a result coincides exactly with well known stationary solutions \\cite{Maydanyuk.2003.PhD-thesis}. From the arguments above the impression could appear that the penetrability defined solely by the shape of the barrier between two turning points is nothing more than prevailing simplified understanding, while for more accurate and deep analysis we need in the strong basis. % \\vspace{2mm} (4) Advance of the semiclassical approach is in simplicity of formula of the penetrability based on determination of the outgoing wave in the asymptotic region. A \\emph{tunneling boundary condition} \\cite{Vilenkin.1995,Vilenkin.1994.PRD} seems to be natural and clear, where the wave function should represent an outgoing wave at large scale factor $a$. However, whether is such a wave free? In contrast to problems of quantum atomic and nuclear physics, in cosmology we deals with potentials, which modules only increase with increasing the scale factor $a$ (also their gradients increase, which have sense of force acting on the wave) and, therefore, we have nothing mutual with a free propagation of the wave in the asymptotic region. Now it is unclear to which combination two Airy functions should be combined at turning point, in order to obtain the proper outgoing wave. It turns out that instead of the free wave in the asymptotic (missing in problems of quantum cosmology), we should be able to work with the waves propagating inside strong fields (see~\\cite{Maydanyuk.2010.IJMPD}). % \\vspace{2mm} These problems violate (destroy) the basis of the semiclassical models, and now statements about reliability of the semiclassical models are transformed into the question of `` faith'' in them, though widespread \\cite{Maydanyuk.2010.IJMPD}. The semiclassical approach could be compared to \\emph{``black box''} in which deeper and more detailed information about the dynamics of the universe is hide. More importantly, in such a black box those missing elements are hidden, without which it is impossible to combine everything together to obtain self-consistent formalism of quantum description of the formation of the universe and its evolution in the first stage. In order to clarify these questions, we have developed a new fully quantum method presented in this paper. % \\vspace{2mm} This paper is organized so. In Sec.~2 a new fully quantum method for description of the formation of the universe and its evolution in the first stage on the basis of a packet, which penetrates from the internal potential well outside by tunneling through the barrier, is presented. A main advance of this method is determination of characteristics of the packet with high accuracy (without implication of the semiclassical approximations). The formalism of the method has been developed relatively the barrier of arbitrary shape, that makes the method universal. In Sec.~3 the method is applied for solution of the problem of evolution of the packet in FRW-model with radiation and Chaplygin gas. The penetrability and reflection relatively the barrier are calculated. We propose new characteristics, more adequately determining the probability of formation of the universe and its subsequent expansion. An accuracy of the method is demonstrated, achieving to $|T_{\\rm bar}+R_{\\rm bar}-1| < 1 \\cdot 10^{-15}$ inside whole under-barrier range of the energy of radiation ($M = 0$, author has not yet found competitive approaches, achieving such a precision), stability in calculations for the obtained results is shown. A special attention in analysis is devoted to study of the initial conditions. On such a basis, for the first time oscillatory dependence of the penetrability on the localization of start of the packet and resonant levels of the energy of radiation $E_(\\rm rad)$ (where the penetration extremely increases) are opened (which are hidden in the semiclassical picture). In finishing, in Sec.~\\ref{sec.conclusions} conclusions of the obtained results are formulated. % ******************************************************************************************************************* % ******************************************************************************************************************* ", "conclusions": "} The new method for determination of probability of penetration of the packet from the internal well outside with its tunneling through one-dimensional barrier of arbitrary shape used in problems of quantum cosmology, is presented. Note the following: % \\begin{enumerate} \\item The method is further development of approach of multiple internal reflections (see Refs.~\\cite{Maydanyuk.2000.UPJ,Maydanyuk.2002.JPS,Maydanyuk.2002.PAST,Maydanyuk.2003.PhD-thesis,% Maydanyuk.2006.FPL,Maydanyuk.arXiv:0805.4165,Maydanyuk.arXiv:0906.4739}, also Refs.~\\cite{Fermor.1966.AJPIA,McVoy.1967.RMPHA,Anderson.1989.AJPIA,Esposito.2003.PRE}), where a process of tunneling of the packet through the barrier is considered consequently by steps of its propagation relatively to each boundary of the barrier. The method is fully quantum, allows to determine amplitudes of wave function, coefficients of penetrability and reflection relatively the barrier. For the first time \\emph{exact} analytical solutions for amplitudes of the wave function, penetrability $T$ and reflection $R$ for the barrier, composed from arbitrary number $n$ of rectangular potential steps, are found. At limit $n \\to \\infty$ these solutions could be considered as exact limits for potential with interested barrier and internal well of arbitrary shape. \\item Accuracy of the method in determination of penetrability $T_{\\rm bar}$ and reflection $R_{\\rm bar}$ is: $|T_{\\rm bar}+R_{\\rm bar}-1| < 1 \\cdot 10^{-15}$ (see Tabl.~\\ref{table.model_Monerat.1}). Author has not found other methods achieving such accuracy in similar problems of quantum physics (with possible exception of some selected cases of exactly solvable barriers which could be obtained by methods of supersymmery). \\item On the basis of the method the probability of penetration of the packet from the internal well outside with its tunneling through the barrier of arbitrary shape called \\emph{coefficient of penetration}, is determined. It succeeds to separate that coefficient explicitly on the penetrability and new coefficient, which characterizes oscillating behavior of the packet inside the internal well and is called \\emph{coefficient of oscillations}. That found for the first time formula seems to be fully quantum analogue of the semiclassical formula of $\\Gamma$ width of decay in quasistationary state proposed by Gurvitz and K\\\"{a}lbermann in Ref.~\\cite{Gurvitz.1987.PRL} (here, the coefficient of oscillations is fully quantum analogue for the semiclassical $F$ factor of formation and the coefficient of penetration is analogue for the semiclassical $\\Gamma$ width). \\end{enumerate} This method has been applied for study of properties of the packet, describing evolution of universe on the first stage in the closed Friedmann--Robertson--Walker model with quantization in the presence of the positive cosmological constant, radiation and component of generalize Chaplygin gas with potential chosen from \\cite{Monerat.2007.PRD}. Let us formulate main results obtained for the first time: % \\begin{enumerate} \\item For the same chosen energy of radiation $E_{\\rm rad}$ the penetrability of the barrier is changed visibly in dependence on the position of the starting point $R_{\\rm start}$ inside the internal well, where the packet begins to propagate (see Fig.~\\ref{fig.model_Monerat.2}): the penetrability has oscillating behavior, difference between its minimums and maximums is minimal at $R_{\\rm start}$ in the center of the well, with increasing $R_{\\rm start}$ this difference increases achieving maximum close to the turning point. The behavior of the coefficients of reflection, oscillations and penetration turns out to be similar. Coincidence (up to the first 15 digits) of the amplitudes of the wave function obtained by such a method, with corresponding amplitudes obtained in the standard approach of quantum mechanics (see App.~\\ref{sec.app.1}) at different energies $E_{\\rm rad}$ confirms that this result does not depend on a choice of the fully quantum method applied for calculations. Such a peculiarity is shown in the fully quantum non-stationary and stationary considerations and it is hidden after imposing the semiclassical restrictions. \\item In non-stationary and stationary considerations the penetrability of the barrier should be connected with initial condition localizing start of the packet. Note that possible introduction of the initial condition into known stationary semiclassical models could change their results. \\item The penetrability is changed visibly, if to take the external tail of the barrier into account. For example, for the barrier (\\ref{eq.model.2.2}) with parameters $A=0.001$ and $B=0.001$ (see Fig.~\\ref{fig.model_Monerat.1}) at the energy of radiation $E_{\\rm rad} = 223$ the penetrability is changed up to 2 times (see Fig.~\\ref{fig.model_Monerat.3}). If to increase the external boundary $a_{\\rm max}$, all amplitudes and coefficients are convergent in calculations that confirms efficiency of the developed method. \\item Dependence of the coefficient of penetration on the energy of radiation has oscillating behavior: here peaks are clearly shown, localized approximately on the same distances, between which smooth minimums (wells) stable in calculations are observed (see Fig.~\\ref{fig.model_Monerat.4}). By other words, for the first time in the fully quantum approach we have obtained clear and stable picture of resonances, which indicate on presence of some quasistationary states. Here, with increasing of the energy of radiation the penetrability is changed monotonously and it describes a general tendency of behavior of the coefficient of penetration, while the coefficient of oscillations gives peaks. Now a reason of existence of resonances becomes clear --- oscillations of the packet inside the internal well give them. For example, for the barrier (\\ref{eq.model.2.2}) with parameters $A=0.001$ and $B=0.001$ I establish 134 such resonant levels inside range $E_{\\rm rad}$ = 0--223 (see Tabl.~\\ref{table.model_Monerat.2},~\\ref{table.model_Monerat.3}). \\item The dependence of the penetrability on the starting point has clear established maximums and minimums. On such a basis one can suppose that the most probable start of the packet (describing start of expansion of the universe) is in one point of such maximums. This allows to predict some definite initial values of the scale factor, at which the universe begins to expand (such initial data is direct result of quantization of the classical cosmological model). \\item Modulus of the wave function both in the internal, and in the external regions has clear established own maximums and minimums~\\cite{Maydanyuk.2008.EPJC,Maydanyuk.2010.IJMPD}. This indicates on such values of the scale factor, at which ``appearance'' of the universe will be more or less probable. By other words, radius of the universe during its expansion is changed not continuously, but passes consequently through some definite discrete values connected with these maximums. It follows that after quantization space-time of universe on the first stage of its expansion seems to be rather discrete than continuous. According to results~\\cite{Maydanyuk.2008.EPJC,Maydanyuk.2010.IJMPD}, difference between maximums and minimums with increasing of the scale factor $a$ is slowly smoothed and we obtain obvious for us continuous structure of the space-time at latter times. Discontinuity of space-time is direct result of quantization of cosmological model, which is shown the most strongly on the first stage of expansion and disappeared after imposition of the semiclassical approximations. \\end{enumerate} % ******************************************************************************************************************* % ******************************************************************************************************************* \\appendix" }, "1005/1005.0036_arXiv.txt": { "abstract": "\\hspace{\\parindent} {\\HII} regions are the birth places of stars, and as such they provide the best measure of current star formation rates (SFRs) in galaxies. The close proximity of the Magellanic Clouds allows us to probe the nature of these star forming regions at small spatial scales. To study the {\\HII} regions, we compute the bolometric infrared flux, or total infrared (TIR), by integrating the flux from 8 to 500~$\\mu$m. The TIR provides a measure of the obscured star formation because the UV photons from hot young stars are absorbed by dust and re-emitted across the mid-to-far-infrared (IR) spectrum. We aim to determine the \\textit{monochromatic} IR band that most accurately traces the TIR and produces an accurate obscured SFR over large spatial scales. We present the spatial analysis, via aperture/annulus photometry, of 16 Large Magellanic Cloud (LMC) and 16 Small Magellanic Cloud (SMC) {\\HII} region complexes using the \\textit{Spitzer Space Telescope's} IRAC (3.6, 4.5, 8~$\\mu$m) and MIPS (24, 70, 160~$\\mu$m) bands. Ultraviolet rocket data (1500 and 1900 \\AA) and SHASSA {\\Ha} data are also included. All data are convolved to the MIPS 160~$\\mu$m resolution (40 arcsec full width at half-maximum), and apertures have a minimum radius of $35\\arcsec$. The IRAC, MIPS, UV, and {\\Ha} spatial analysis are compared with the spatial analysis of the TIR. We find that nearly all of the LMC and SMC {\\HII} region spectral energy distributions (SEDs) peak around 70~$\\mu$m at all radii, from $\\sim10$ to $\\sim400$ pc from the central ionizing sources. As a result, we find the following: the sizes of {\\HII} regions as probed by 70~$\\mu$m is approximately equal to the sizes as probed by TIR ($\\approx70$ pc in radius); the radial profile of the 70~$\\mu$m flux, normalized by TIR, is constant at all radii (70~$\\mu\\mbox{m}\\sim0.45~\\mbox{TIR}$); the $1\\sigma$ standard deviation of the 70~$\\mu$m fluxes, normalized by TIR, is a lower fraction of the mean ($0.05-0.12$ out to $\\sim220$ pc) than the normalized 8, 24, and 160~$\\mu$m normalized fluxes ($0.12-0.52$); and these results are the same for the LMC and the SMC. From these results, we argue that 70~$\\mu$m is the most suitable IR band to use as a monochromatic obscured star formation indicator because it most accurately reproduces the TIR of {\\HII} regions in the LMC and SMC and over large spatial scales. We also explore the general trends of the 8, 24, 70, and 160~$\\mu$m bands in the LMC and SMC {\\HII} region SEDs, radial surface brightness profiles, sizes, and normalized (by TIR) radial flux profiles. We derive an obscured SFR equation that is modified from the literature to use 70~$\\mu$m luminosity, $\\mbox{SFR}~(M_{\\odot}~\\mbox{yr}^{-1}) = 9.7(0.7)\\times10^{-44}~L_{70}~(\\mbox{ergs}~\\mbox{s}^{-1})$, which is applicable from 10 to 300 pc distance from the center of an {\\HII} region. We include an analysis of the spatial variations around {\\HII} regions between the obscured star formation indicators given by the IR and the unobscured star formation indicators given by UV and {\\Ha}. We compute obscured and unobscured SFRs using equations from the literature and examine the spatial variations of the SFRs around {\\HII} regions. ", "introduction": "\\hspace{\\parindent} {\\HII} regions are locations of active or recent star formation where the extreme ultraviolet (UV) radiation from massive OB stars ionizes the surrounding gas \\citep[for a good review see chapter 5 of][and references therein]{tiel05}. It is common to consider {\\HII} regions as being large complexes of overlapping individual {\\HII} regions where the sizes and components are not determined solely by the classical Str\\\"{o}mgren radius of an individual star but by an inclusion of all of the physical regimes in the interstellar medium (ISM) affected by the far-UV photons of the central sources. These {\\HII} region components include the central ionizing OB stars, and the surrounding photodissociation regions (PDRs) and molecular clouds, where the chemistry and the heating are driven by the UV photons from nearby stars \\citep{tiel05,rela09,wats08}. Thus, {\\HII} regions in this context are really {\\HII} complexes, or more generally, star forming regions. The observed properties of these star forming regions, including size and shape, depend on the physical traits such as the number of young hot ionizing stars, density of neutral gas, abundance of dust, star formation history, and prior supernovae \\citep[e.g., see][]{hodg74,tiel05,walb02,snid09,harr09}. These physical traits are often correlated, and the use of data at many wavelengths to observe {\\HII} regions in different galactic environments is required to piece them together. The nature of the hot central OB stars can be studied using UV \\citep[e.g.,][]{smit87,mart05} or nebular emission lines, such as {\\Ha} \\citep[e.g.,][]{heni56,hodg83,gaus01}. The {\\Ha} luminosity function of a galaxy is frequently used to determine many of the physical properties of {\\HII} regions, including the number of ionizing stars, evolutionary effects, and possible environmental effects \\citep{kenn89,oey98}. {\\HII} region size distributions are directly related to the luminosity functions and relate to the numbers of nebula of a given size for a galaxy\\citep[e.g.,][]{van81,oey03}. The PDRs can be studied by many methods including the [\\OI] and [\\CII] infrared (IR) cooling lines and CO radio data \\citep[e.g.,][]{kauf99}. The molecular clouds are typically studied using radio observations of molecular rotational lines \\citep[e.g.,][]{cohe88,genz91,fuku08}. Present star formation rates (SFRs) are calculated using a tracer of the UV photons from the young massive stars and spectral synthesis models \\citep[see][]{kenn98}. {\\HII} regions will have some fraction of their UV photons obscured by dust and some fraction unobscured. For unobscured SFRs, the ionizing photons can be directly observed via UV observations or recombination lines such as {\\Ha}. For obscured {\\HII} regions, bolometric IR observations of dust (i.e., the total infrared (TIR)) can be used to recover the extinguished UV photons. This is because the dust absorption cross section is highly peaked in the UV, and the re-emitted flux is in the broad spectral range from the mid-to-far-IR \\citep{kenn98}. Because the TIR around {\\HII} regions accounts for all of the extincted UV photons, the TIR is expected to be the single best indicator of SFR obscured by dust. Combining UV, optical, and IR observations of {\\HII} regions across a whole galaxy allows us to probe the current SFR of that galaxy. The Large Magellanic Cloud (LMC) and Small Magellanic Cloud (SMC) are ideal natural laboratories for studying star forming regions and their effects on the ISM. The close proximities of the LMC and SMC, at $\\sim52$ kpc \\citep{szew08} and $\\sim60$ kpc \\citep{hild05}, respectively, allow for detailed star formation studies down to parsec or sub-parsec scales depending on wavelength. Any broad study of {\\HII} regions across the LMC and SMC can take advantage of many multiwavelength observations, including the rocket UV data from \\citet{smit87}, the Southern {\\Ha} Sky Survey Atlas (SHASSA) {\\Ha} data from \\citet{gaus01}, and the \\textit{Spitzer Space Telescope} IR data from \\citet{meix06} and {\\gord}. Furthermore, there are many past optical {\\HII} region surveys that catalog the sources of {\\Ha} in the LMC and SMC \\citep[see][]{heni56,davi76,bica95}. Because of the observational advantages, many researchers rely on a single-band star formation indicator (i.e., UV, {\\Ha}, {\\Pa}, 8~$\\mu$m, 24~$\\mu$m, etc). There are complications in the UV and optical lines in that they can be greatly impacted by extinction, thus, requiring extinction corrections. Another complication is that the observed UV photons can come from stars of various ages ($< 100$ Myr) and will greatly depend on galaxy type (i.e., quiescent spiral galaxies, starbursts, etc.; \\citep{calz05}). The 8 and 24~$\\mu$m IR band emission will likely depend on the environment of the host galaxy because the abundance of the aromatics/small grains that give rise to their emission depend upon the metallicity and ionizing radiation present \\citep[K. D. Gordon et al. 2010, in preparation;][]{drai07,drai07b}. Work done by \\citet{calz07} and \\citet{dale05} indicate that 8 $\\mu$m makes for a poor star formation indicator due to large variability of emission in galaxies with respect to spectral energy distribution (SED) shape and metallicity. \\citet{calz07} also note, along with \\citet{calz05}, that a star formation indicator using 24~$\\mu$m by itself can vary from galaxy to galaxy. \\citet{dale05} claim that SFRs calculated from 24~$\\mu$m emission may be off by a factor of 5 due to variations of 24~$\\mu$m flux with respect to SEDs observed across nearby galaxies. To compensate for the extinction effects in UV and optical nebular emission lines, many researchers are now measuring SFRs via a combination of obscured (TIR, 8~$\\mu$m, 24~$\\mu$m) and unobscured (UV, {\\Ha}, {\\Pa}) star formation indicators \\citep[e.g.,][]{calz07,kenn07,thil07,rela09,kenn09}. However, the noted differences in 8 and 24~$\\mu$m emission, relative to host galaxy properties, may still introduce uncertainties to the calculated SFRs when applying them across large galaxy samples. In their analysis of 33 galaxies from the \\textit{Spitzer} SINGS sample, \\citet{calz07} claim that a combination of {\\Ha} and 24~$\\mu$m gives the most robust SFR using a procedure similar to the \\citet{gord00} ``flux ratio method''\\footnote{The ``flux ratio method'' uses UV and IR fluxes of galaxies to derive extinction-corrected UV luminosities \\citep{gord00}}. Their calibration has a caveat in that it is useful for actively star forming galaxies where the energy output is dominated by young stellar populations \\citep{calz07}. \\citet{kenn09} analyze SFRs of nearby galaxies derived by combining {\\Ha} with 8~$\\mu$m, 24~$\\mu$m, and the TIR. They find that linear combinations of {\\Ha} and TIR provide for the most robust SFRs. There is little work done on investigating the efficacy of using 70 or 160~$\\mu$m as star formation indicators. \\citet{dale05} claim that the 70~$\\mu$m emission may make a good monochromatic obscured star formation indicator because the 70 to 160~$\\mu$m ratio correlates well with local SFRs. A \\textit{Spitzer} analysis of far-IR compact sources in the LMC \\citep{jacc10} and SMC \\citep{jacc10b}, including compact {\\HII} regions, find that the bolometric correction to 70 $\\mu$m is modest, due to a typical dust temperature of 40~K. In a study of dwarf irregular galaxies, \\citet{walt07} find a good correlation between the brightest 70~$\\mu$m regions and optical tracers of star formation, albeit, with some galaxies contributing significant 70~$\\mu$m diffuse emission at large radii. Which of the \\textit{Spitzer} IR bands most accurately reproduces the TIR over a large spatial scale? Employing aperture/annulus photometry, we analyze 16 LMC and 16 SMC {\\HII} region complexes using the \\textit{Spitzer} Infrared Array Camera (IRAC) and Multiband Imaging Photometer (MIPS) bands. We determine that the MIPS 70~$\\mu$m band provides for the most accurate monochromatic obscured star formation indicator based on an analysis of the {\\HII} region complex SEDs, sizes, and radial monochromatic IR fluxes (normalized by the TIR). We include an analysis of the spatial distribution of the unobscured star formation indicators, UV and {\\Ha}, relative to the IR obscured star formation indicators. We modify an established TIR SFR recipe from \\citet{kenn98} to derive a new monochromatic obscured SFR equation using the 70~$\\mu$m luminosity. In Section~\\ref{sample}, we list the LMC and SMC {\\HII} region complexes sampled in this work. In Section~\\ref{obs}, we explain the IR, UV, and {\\Ha} observations and data reduction. The analysis of the multiwavelength photometry is discussed in Section~\\ref{analysis}. The basic results of the photometry, SEDs, and radial profiles, are discussed in Section~\\ref{results} as well as a discussion of our calculation of the TIR. Discussions of {\\HII} region normalized radial SEDs, {\\HII} complex sizes, normalized radial profiles, and SFRs are presented in Section~\\ref{disc}. In this section we also present our derived 70~$\\mu$m obscured SFR equation. We finish with some concluding statements in Section~\\ref{conc}. ", "conclusions": "\\label{conc} \\hspace{\\parindent}We performed UV, {\\Ha}, and IR aperture/annulus photometry of 16 LMC and 16 SMC {\\HII} regions at the common resolution of the MIPS 160~$\\mu$m $40\\arcsec$ {\\fwhm} PSF. With the aperture photometry, we computed IR SEDs using the total cumulative \\textit{Spitzer} IRAC (3.6, 4.5, 8~$\\mu$m) and MIPS (24, 70, 160~$\\mu$m) fluxes of the LMC, SMC, and each {\\HII} region. The cumulative SED of the LMC is brighter than the SMC, and their shapes are consistent with IR SEDs of the Magellanic Clouds in the literature. The SMC {\\HII} regions have cumulative fluxes that are weaker, by about an order of magnitude, than the LMC {\\HII} regions. We argue that this is expected from what is known about the LMC and SMC {\\HII} region luminosity functions. We compute the IR bolometric flux (TIR) for each {\\HII} region aperture/annulus by integrating under the linear extrapolation of the 8, 24, 70, and 160~$\\mu$m fluxes. We include the input from the Rayleigh--Jeans tail of a computed dust MBB by integrating in steps of 0.25~\\AA, from 160 to $500~\\mu$m. Using annulus photometry of the {\\HII} regions, at distances of $\\sim10$ up to $\\sim400$ pc from their cores, we plot the radial 8~$\\mu$m, 24~$\\mu$m, 70~$\\mu$m, 160~$\\mu$m and TIR surface brightness profiles of the 32 {\\HII} regions. The surface brightness profiles peak near the center of the {\\HII} regions for all IR bands and TIR. We argue that resolution effects may play a role in reducing the SMC {\\HII} region surface brightness profiles relative to the LMC {\\HII} region surface brightness profiles. The surface brightness profiles are used to compute the sizes of each {\\HII} region, using the TIR, 8~$\\mu$m, 24~$\\mu$m, 70~$\\mu$m, and 160~$\\mu$m emission. The typical sizes measured have large scatter and are $\\approx75$ pc for 8~$\\mu$m, $\\approx50$ pc for 24~$\\mu$m, $\\approx70$ pc for both 70~$\\mu$m and TIR, and $\\approx90$ pc for 160~$\\mu$m. The large scatter for each size measurement likely stems from sampling different parts of the LMC and SMC luminosity functions. We compute normalized (by TIR) IR SEDs for each {\\HII} region annulus. The radial SEDs nearly all peak around 70~$\\mu$m at all radii for every {\\HII} region, out to $\\sim400$ pc. The result of this is that it gives four favorable reasons to choose 70~$\\mu$m as a monochromatic obscured star formation indicator, and they are (1) 70~$\\mu$m emission most closely traces the size of {\\HII} regions as found using the TIR ($\\approx70$ pc in radius); (2) 70~$\\mu$m flux, normalized by TIR, remains nearly constant with radius ($L_{70}\\approx45\\%~L_{\\mbox{TIR}}$) from $\\sim10$ to 400 pc; (3) 70~$\\mu$m flux, normalized by TIR, has the smallest fractional error ($0.05-0.12$ out to 220 pc); and (4) 70 $\\mu$m flux, normalized by TIR, does not systematically differ between the LMC and SMC. Radial SFR density plots, using the obscured indicator (TIR) and unobscured indicators (UV, {\\Ha}), are computed using the SFR recipes of \\citet{kenn98}. The computed SFRs for all indicators peak near the center of each {\\HII} region. The SFRs for the LMC are larger than for the SMC due to the larger luminosities of the LMC {\\HII} regions. Within the limits of our resolution, the spatial similarities between UV, {\\Ha}, and 24~$\\mu$m are similar to results in the literature. A modified version of the obscured SFR equation from \\citet{kenn98} is created that depends on a conversion factor from TIR to a monochromatic IR band. The conversion factors for the 8, 24, and 160~$\\mu$m bands all have large uncertainties and are dependent on radius and/or host galaxy. The conversion factor for 70~$\\mu$m is constant, within the low uncertainties, at distances from 10 to 300 pc, and between the LMC and SMC. We produce a final modified obscured SFR equation using a single averaged 70~$\\mu$m conversion constant. This \\citet{kenn98} modified equation is applicable for Magellanic-like {\\HII} regions and for aperture sizes of $10-300$ pc radius. The LMC and SMC have different environmental properties than those galaxies of earlier Hubble types. The dust emission observed around Magellanic Cloud {\\HII} regions may not be indicative of dust emission around {\\HII} regions in other galaxies. The analysis in this work would be greatly benefited by including {\\HII} regions in a larger sample of galaxies. By analyzing the 30~Doradus IR properties, we found possible caveats in using 70~$\\mu$m as an obscured star formation indicator. If {\\HII} regions in a particular galaxy have, on average, similar fluxes as 30~Doradus, then the 70~$\\mu$m emission, relative to the TIR, will exhibit more scatter. However, we do not expect this to be the case because 30~Doradus is on the extreme high end of {\\HII} region {\\Ha}-derived luminosity functions \\citep{kenn89}. Another general caveat to mention is that this work applies directly to {\\HII} regions and their immediate surroundings. It is general practice for those studying SFRs over entire galaxies to use the SFR recipes calculated from {\\HII} regions, despite the added uncertainties of going from a local star forming region to an entire galaxy. The results from this work, applied to entire galaxies, will be subject to similar uncertainties. The recent launch of the \\textit{HERSCHEL} telescope will greatly increase our understanding of star formation in galaxies. At 3.5~m, not only is \\textit{HERSCHEL} the largest space telescope ever launched, but the PACS photometer, with the ability to observe at 75~$\\mu$m, 110~$\\mu$m, or 175~$\\mu$m, will provide a measure of {\\HII} region SEDs at wavelengths near their peak emissions. The SPIRE camera will observe at 250~$\\mu$m, 350~$\\mu$m, and 500~$\\mu$m (albeit with a lower resolution than PACS) and can provide a better constraint on the Rayleigh--Jeans side of the dust blackbody emission. NASA's 2.5~m \\textit{SOPHIA} telescope will also have a suite of high-resolution instruments capable of observing star forming regions throughout the IR, including near the expected peaks of the {\\HII} region SEDs. This work is based on observations made with the \\textit{Spitzer Space Telescope}, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under NASA contract 1407. The SAGE-LMC project has been supported by NASA/Spitzer grant 1275598 and Meixner's efforts have had additional support from NASA NAG5-12595." }, "1005/1005.2519_arXiv.txt": { "abstract": "{The Polaris Flare cloud region contains a great deal of extended emission. It is at high declination and high Galactic latitude. It was previously seen strongly in IRAS Cirrus emission at 100 microns. We have detected it with both PACS and SPIRE on {\\it Herschel}. We see filamentary and low-level structure. We identify the five densest cores within this structure. We present the results of a temperature, mass and density analysis of these cores. We compare their observed masses to their virial masses, and see that in all cases the observed masses lie close to the lower end of the range of estimated virial masses. Therefore, we cannot say whether they are gravitationally bound prestellar cores. Nevertheless, these are the best candidates to be potential prestellar cores in the Polaris cloud region.} ", "introduction": "In this paper we present observations, performed with the ESA {\\it Herschel} Space Observatory (Pilbratt et al 2010), of the Polaris Flare region. In particular we use the large collecting area and powerful science payload of {\\it Herschel} to perform imaging photometry using the PACS (Poglitsch et al 2010) and SPIRE (Griffin et al 2010) instruments. These observations were carried out as part of the guaranteed-time key programme to map most of the Gould Belt star-forming regions with {\\it Herschel} (Andr\\'e et al 2010). The Polaris Flare was first detected in HI as a spur of gas that appears to rise more than 30$^\\circ$ out of the Galactic Plane. This region is an area rich in IRAS cirrus emission (e.g. Low et al 1984), and is sometimes known as the Polaris Cirrus Cloud. It was mapped in CO by Heithausen \\& Thaddeus (1990). On the large scale this cloud appears to merge with the Cepheus Flare cloud (e.g. Kirk et al 2009), and both clouds extend to high Galactic latitude. \\begin{figure*} \\centering{\\includegraphics[width=\\textwidth]{14618fg1.eps}} \\caption{\\label{5bands}The densest part of the Polaris Flare region at some of the observed wavebands. Upper row: 160$\\mu$m from PACS, and 250$\\mu$m and 350$\\mu$m from SPIRE. Lower row: False-colour image (where 160$\\mu$m is shown in blue, 250$\\mu$m is shown in green, and 350$\\mu$m is shown in red), column density map (where red is $<$4, blue is 4--8, and yellow is $>$8 $\\times$ 10$^{21}$ cm$^{-2}$), and colour temperature map (where blue is 10--11~K and yellow is 12--13~K). The contour levels on the column density map start at 4 $\\times$ 10$^{21}$ cm$^{-2}$, and the interval between successive contours is 1.5 $\\times$ 10$^{21}$ cm$^{-2}$. The same contours are repeated on the temperature map for ease of location. Five sources are seen above a column density of 4 $\\times$ 10$^{21}$ cm$^{-2}$. These are labelled cores 1--5 (in order of increasing R.A.) on the last two panels and are discussed in the text. The loop (loop 1) discussed in the text (containing cores 4 \\& 5) is clearly visible in all images. The reddest features on the false-colour image are the coldest, and the loop shows up clearly as redder than the surroundings. Likewise in the temperature map, the loop shows up as blue, indicating that it is the coldest feature on the map. The position of the IRAS source (IRAS~01432+8725) is marked with a star on the last two panels (adjacent to core 4).} \\end{figure*} One of the denser regions in the cloud is known as molecular cloud 123.5+24.9, or MCLD~123.5+24.9 (e.g. Bensch et al 2003) -- hereafter MCLD~123 -- at a distance of 150~pc (Bensch et al 2003). It shows strong extended IRAS 100-$\\mu$m emission and is generally believed to be gravitationally unbound with a mass of $\\sim$18--32~M$_\\odot$ (Grossmann et al 1990; Bensch et al 2003). A CO study by Falgarone et al (1998) revealed a curved filament in MCLD~123 in $^{13}$CO and C$^{18}$O -- both in the J=2--1 transition. This filament is also apparent in some narrow velocity channels in the same transition of $^{12}$CO (Falgarone et al 1998). There is one IRAS source in the region, IRAS~01432+8725. This is listed in the IRAS catalogue as having a flux density at 100~$\\mu$m of 2.88~Jy, but only upper limits at the other IRAS wavebands. There is also one {\\it Spitzer} source that was only detected at a wavelength of 24~$\\mu$m at coordinates R.A. (2000) $=$ 01$^{\\rm h}$ 58$^{\\rm m}$ 27.5$^{\\rm s}$, Dec. (2000) $=$ $+$87\\degr\\ 40\\arcmin\\ 07\\arcsec\\ . It has a peak flux density at 24~$\\mu$m of 1.3~mJy/beam, where the {\\it Spitzer} beam at this wavelength is 7~arcsec. This detection lies in a {\\it Spitzer} calibration field in an unpublished archival dataset (AOR~33136386). ", "conclusions": "We have presented {\\it Herschel} data of the Polaris Flare dark cloud region, and in particular the region MCLD~123. We found a great deal of extended emission at wavelengths from 70 to 500~$\\mu$m with both PACS and SPIRE. We noted some filamentary and low-level structure. We identified the five densest cores within this structure. We carried out a temperature, mass and density analysis of the cores. We compared their observed masses to their virial masses, and found that the observed masses are on the lower limit of the range of their estimated virial masses, and thus we cannot say for certain whether they are gravitationally bound." }, "1005/1005.2175_arXiv.txt": { "abstract": "Certain oscillatory features in the primordial scalar power spectrum are known to provide a better fit to the outliers in the cosmic microwave background data near the multipole moments of $\\ell=22$ and $40$. These features are usually generated by introducing a step in the popular, quadratic potential describing the canonical scalar field. Such a model will be ruled out, if the tensors remain undetected at a level corresponding to a tensor-to-scalar ratio of, say, $r\\simeq 0.1$. In this work, in addition to the popular quadratic potential, we investigate the effects of the step in a small field model and a tachyon model. With possible applications to future datasets (such as PLANCK) in mind, we evaluate the tensor power spectrum exactly, and include its contribution in our analysis. We compare the models with the WMAP (five as well as seven-year), the QUaD and the ACBAR data. As expected, a step at a particular location and of a suitable magnitude and width is found to improve the fit to the outliers (near $\\ell=22$ and $40$) in all these cases. We point out that, if the tensors prove to be small (say, $r\\lesssim 0.01$), the quadratic potential and the tachyon model will cease to be viable, and more attention will need to be paid to examples such as the small field models. ", "introduction": "Inflation continues to remain the most promising paradigm for describing the origin of the perturbations in the early universe. It has been performing remarkably well against the observational data, and the challenge before the other competing scenarios is to match the simplicity and efficiency of inflation. Many models of inflation lead to an epoch of slow roll lasting for, say, $50$-$60$ e-folds, as is required to resolve the horizon problem. It is well known that slow roll inflation leads to a featureless, nearly scale invariant, power law, primordial scalar spectrum. Such a spectrum, along with the assumption of a spatially flat, concordant $\\Lambda$CDM background cosmological model, provides a good fit to the recent observations of the anisotropies in the Cosmic Microwave Background (CMB) by different missions such as the Wilkinson Microwave Anisotropy Probe (WMAP)~\\cite{wmap-5,wmap-7}, the QUEST at DASI (QUaD)~\\cite{quad-2009}, and the Arcminute Cosmology Bolometer Array Receiver (ACBAR)~\\cite{acbar-2008}. The efficacy of the inflationary scenario also seems to be responsible for an important drawback. Though, as a paradigm, inflation can be considered to be a success, it would be fair to say that we are rather far from converging on a specific model or even a class of models of inflation. There exist a plethora of inflationary models that remain consistent with the data. We mentioned above that a nearly scale invariant, power law, scalar spectrum fits the observations of the anisotropies in the CMB quite well. However, there exist a few data points at the lower multipoles---notably, at the quadrapole ($\\ell=2$) and near the multipole moments of $\\ell=22$ and $40$---which lie outside the cosmic variance associated with the power law primordial spectrum. Needless to add, statistically, a few outliers in a thousand or so data points can always be expected. These outliers were noticed in the WMAP first-year data, and they continue to be present even in the most recent, seven-year data, making them unlikely to be artifacts of data analysis. It is possible that they actually indicate certain non-trivial inflationary dynamics. In that case, these outliers are important from the phenomenological perspective of attempting to constrain the models from the data, because only a more restricted class of inflationary models can be expected to provide an improved fit to these outliers. Therefore, it is a worthwhile exercise to systematically explore models that lead to specific deviations from the standard power law, inflationary perturbation spectrum, and also provide an improved fit to the data. Various efforts towards a model independent reconstruction of the primordial spectrum from the observed pattern of the CMB anisotropies seem to indicate the presence of certain features in the spectrum~\\cite{rc}. (However, we should add that there also exist other views on the possibility of features in the primordial spectrum; in this context, see, for example, Refs.~\\cite{rc-wf}.) In particular, a burst of oscillations in the primordial spectrum seems to provide a better fit to the CMB angular power spectrum near the multipole moments of $\\ell=22$ and $40$. Generating these oscillations requires a short period of deviation from slow roll inflation~\\cite{starobinsky-1992,dvorkin-2010}, and such a departure has often been achieved by introducing a small step in the popular, quadratic potential describing the canonical scalar field (see Refs.~\\cite{adams-2001,wmap-1,covi-2006-2007,mortonson-2009}; for a discussion on other models, see, for instance, Refs.~\\cite{joy-2008-2009,jain-2009}). At the cost of three additional parameters which characterize the location, the height and the width of the step, it has been found that this model provides a considerably better fit to the CMB data with the least squares parameter $\\chi_{\\rm eff}^{2}$ typically improving by about $7$, when compared to the nearly scale invariant spectrum that would have resulted in the absence of the step~\\cite{covi-2006-2007,mortonson-2009}. But, such a chaotic inflation model leads to a reasonable amount of tensors, and these models will be ruled out if tensors are not detected corresponding to a tensor-to-scalar ratio of, say, $r\\simeq 0.1$. Our aims in this paper are twofold. Firstly, we wish to examine whether, with the introduction of a step, other inflationary models too perform equally well against the CMB data, as the quadratic potential does. Secondly, we would also like to consider a model that leads to a tensor-to-scalar ratio of $r<0.1$, so that suitable alternative models exist if the tensor contribution turns out to be smaller. Motivated by these considerations, apart from revisiting the popular quadratic potential, we shall investigate the effects of the step in a small field model~(in this context, see, for example, Ref.~\\cite{efstathiou-2006}) and a tachyon model~\\cite{steer-2004}. Also, with possible applications to future datasets in mind (such as the ongoing PLANCK mission~\\cite{planck}), we shall evaluate the tensor power spectrum exactly, and include its contribution in our analysis. We shall compare the models with the CMB data from the WMAP, the QUaD and the ACBAR missions. We shall consider the five as well as the seven-year WMAP data~\\cite{wmap-5,wmap-7}, the QUaD June 2009 data~\\cite{quad-2009} and the ACBAR 2008 data~\\cite{acbar-2008} to arrive at the observational constraints on the inflationary parameters. We find that, as one may expect, a step at a suitable location and of a certain magnitude and width improves the fit to the outliers (near $\\ell=22$ and $40$) in all the cases. We point out that, if the amplitude of the tensors prove to be small, the quadratic potential and the tachyon model will become inviable, and we will have to turn our attention to examples such as the small field models. The remainder of this paper is organized as follows. In the following section, we shall outline the different inflationary models that we shall be considering in this work. In Section~\\ref{sec:method}, we shall discuss the methodology that we adopt for comparing the inflationary models with the data, the datasets that we use for our analysis, and the priors on the various parameters that we work with. In Section~\\ref{sec:results}, we shall present the results of our comparison of the theoretical CMB angular power spectra that arise from the various models with the WMAP five-year as well as seven-year data, the QUaD and the ACBAR data. We shall tabulate the best fit values that we obtain on the background cosmological parameters and the parameters describing the inflationary models. We shall also illustrate the constraints that we arrive at on the parameters describing the step in the case of the small field model. Further, we shall explicitly show that the models with the step perform better against the data because of the fact that they lead to an improvement in the fit to the outliers around $\\ell=22$ and $40$. In Section~\\ref{sec:sps}, we shall illustrate the scalar power spectra and the CMB angular power spectra corresponding to the best fit values of the parameters of some of the models that we consider. Finally, in Section~\\ref{sec:summary}, we shall close with a brief summary, and a few comments on certain implications of our results. Note that we shall work in units such that $\\hbar=c=(8\\,\\pi\\,G)=1$. Moreover, we shall assume the background cosmological model to be the standard, spatially flat, $\\Lambda$CDM model. ", "conclusions": "\\label{sec:summary} In this work, we have investigated the effects of introducing a step in certain inflationary models. In addition to revisiting the case of the quadratic potential that has been considered earlier, we have studied the effects of the step in a small field model and a tachyon model. The introduction of the step leads to a small deviation from slow roll inflation, which results in a burst of oscillations in the scalar power spectrum. These oscillations, in turn, leave their imprints as specific features in the CMB angular power spectrum. Actually, we have also evaluated the tensor power spectrum exactly, and have included it in our analysis. We believe that this is a timely effort considering the fact that results from, say, the ongoing PLANCK mission might necessitate such an analysis. Upon comparing the inflationary models with the WMAP, the QUaD and the ACBAR data, we find that, with the step, all the models lead to an improvement in $\\chi_{\\rm eff}^{2}$ by about $7$-$9$ over the smooth, nearly scale invariant, slow roll spectrum, at the expense of three additional parameters describing the location, the height and the width of the step in the inflaton potential. The output of the WMAP likelihood code and a plot of the difference in $\\chi_{\\rm eff}^{2}$ with and without the step clearly illustrate that the improvement occurs because of a better fit to the data at the lower multipoles due to the presence of the step. Evidently, if future observations indicate that the amplitude of the tensors are rather small, then the quadratic potential and the tachyon model will be ruled out, while a suitable small field model with a step can be expected to perform well against the data. The introduction of the step in an inflationary model can possibly be viewed as an abrupt change in a potential parameter~\\cite{adams-2001}. But, it has to be admitted that it is rather ad-hoc, and one needs to explore the generation of features and a resulting improvement in the fit in better motivated inflationary models. Two field models offer such a possibility. For instance, with suitably chosen parameters, the two field models can easily lead to a brief departure from slow roll inflation (in this context, see, for example, Refs.~\\cite{joy-2008-2009,cline-2003,hunt-2004-2007}). However, iso-curvature perturbations arise whenever more than one field is involved~\\cite{gordon-2000,tsujikawa-2003-2005}, and they need to be carefully taken into account when comparing these models with the data. It will be a worthwhile effort to systematically explore the two field models, including the effects due to the iso-curvature perturbations, in an attempt to fit the outliers near the multipole moments of $\\ell=22$ and $40$. Over the last few years, it has been recognized that primordial non-Gaussianity can act as a powerful observational tool that can help us discriminate further between the various inflationary models. For example, it has been shown that slow roll inflation driven by the canonical scalar fields leads only to a small amount of non-Gaussianity~\\cite{maldacena-2003}. However, recent analysis of the CMB data seem to suggest that the extent of primordial non-Gaussianity may possibly be large (see, for instance, Refs.~\\cite{wmap-7,smith-2009}). It is known that models which lead to features, such as the ones we have considered here, also generate a reasonably large non-Gaussianity (see, for example, Refs.~\\cite{chen-2007-2008}). While the different models that we have considered in this work lead to virtually the same scalar power spectrum and almost the same extent of improvement in the fit (i.e. with the introduction of the step) to the CMB data, it is important to examine whether they lead to the same extent of non-Gaussianity as well. We are currently investigating such issues." }, "1005/1005.2944_arXiv.txt": { "abstract": "{ We present NLTE Li abundances for 88 stars in the metallicity range -3.5 $<$ [Fe/H] $<$ -1.0. The effective temperatures are based on the infrared flux method with improved $E(B-V)$ values obtained mostly from interstellar \\ion{Na}{i}\\,D lines. The Li abundances were derived through MARCS models and high-quality UVES+VLT, HIRES+Keck and FIES+NOT spectra, and complemented with reliable equivalent widths from the literature. The less-depleted stars with [Fe/H]$<-2.5$ and [Fe/H] $>-2.5$ fall into two well-defined plateaus of $A_{\\rm Li}$ $= 2.18 \\; (\\sigma =0.04)$ and $A_{\\rm Li}$ $=2.27 \\; (\\sigma =0.05)$, respectively. We show that the two plateaus are flat, unlike previous claims for a steep monotonic decrease in Li abundances with decreasing metallicities. At all metallicities we uncover a fine-structure in the Li abundances of Spite plateau stars, which we trace to Li depletion that depends on both metallicity and mass. Models including atomic diffusion and turbulent mixing seem to reproduce the observed Li depletion assuming a primordial Li abundance $A_{\\rm Li}$ = 2.64, which agrees well with current predictions ($A_{\\rm Li}$ = 2.72) from standard Big Bang nucleosynthesis. Adopting the Kurucz overshooting model atmospheres increases the Li abundance by +0.08 dex to $A_{\\rm Li}$ = 2.72, which perfectly agrees with BBN+WMAP. } ", "introduction": "One of the most important discoveries in the study of the chemical composition of stars was made in 1982 by Monique and Fran\\c cois Spite, who found an essentially constant Li abundance in warm metal-poor stars (Spite \\& Spite 1982), a result interpreted as a relic of primordial nucleosynthesis. Due to its cosmological significance, there have been many studies devoted to Li in metal-poor field stars (e.g., Ryan et al. 1999; Mel\\'endez \\& Ram{\\'{\\i}}rez 2004, hereafter MR04; Boesgaard et al. 2005, hereafter B05; Charbonnel \\& Primas 2005; Asplund et al. 2006, hereafter A06; Shi et al. 2007, hereafter S07; Bonifacio et al. 2007, hereafter B07; Hosford et al. 2009; Aoki et al. 2009; Sbordone et al. 2010, hereafter S10), with observed Li abundances at the lowest metallicities ([Fe/H] $\\sim-$3) from $A_{\\rm Li}$ = 1.94 (B07) to $A_{\\rm Li}$ = 2.37 (MR04). A primordial Li abundance of $A_{\\rm Li}$ = 2.72$_{-0.06}^{+0.05}$ is predicted (Cyburt et al. 2008; see also Steigman 2009; Coc \\& Vangioni 2010) with the theory of big bang nucleosynthesis (BBN) and the baryon density obtained from WMAP data (Dunkley et al. 2009), which is a factor of 2-6 times higher than the Li abundance inferred from halo stars. There have been many theoretical studies on non-standard BBN trying to explain the cosmological Li discrepancy by exploring the frontiers of new physics (e.g. Coc et al. 2009; Iocco et al. 2009; Jedamzik \\& Pospelov 2009; Kohri \\& Santoso 2009). Alternatively, the Li problem could be explained by a reduction of the original Li stellar abundance due to internal processes (i.e., by stellar depletion). In particular, stellar models including atomic diffusion and mixing can deplete a significant fraction of the initial Li content (Richard et al. 2005; Piau 2008), although such models depend on largely unconstrained free parameters. On the other hand, it is not easy to reconcile the lack of observed abundance scatter in the Spite plateau with substantial Li depletion (e.g. Ryan et al. 1999; A06). Due to the uncertainties in the Li abundances and to the small samples available, only limited comparisons of models of Li depletion with stars in a broad range of mass and metallicities have been performed (e.g., Pinsonneault et al. 2002; B05). The observed scatter in derived Li abundances in previous studies of metal-poor stars can be as low as 0.03 dex (e.g. Ryan et al. 1999; A06), fully consistent with the expected observational errors. Yet, for faint stars observational errors as high as 0.2 dex have been reported (e.g., Aoki et al. 2009). In order to provide meaningful comparisons with stellar depletion models, precise Li abundances for a large sample of stars are needed. Here we present such a study for the first time for a large sample of metal-poor stars (-3.5 $<$ [Fe/H] $<$ -1.0) with masses in a relatively broad mass range (0.6-0.9 M$_\\odot$). ", "conclusions": "\\subsection{The \\teff cutoff of the Spite plateau} Despite the fact that Li depletion depends on mass (e.g. Pinsonneault et al. 1992), this variable has been ignored by most previous studies. Usually a cutoff in \\teff is imposed to exclude severely Li-depleted stars in the Spite plateau, with a wide range of adopted cutoffs, such as 5500 K (Spite \\& Spite 1982), 5700 K (B05), 6000 K (MR04; S07) and $\\sim$6100 K for stars with [Fe/H] $< -2.5$ (Hosford et al. 2009). At a given mass, the \\teff of metal-poor stars has a strong metallicity-dependence (e.g. Demarque et al. 2004). As shown in Figs. 11-12 of M06, the \\teff of turnoff stars increases for decreasing metallicities. Hence, a metallicity-independent cutoff in \\teff may be an inadequate way to exclude low-mass Li-depleted stars from the Spite plateau. As show in Fig. 2, where $A_{\\rm Li}$ in different metallicity bins is shown as a function of \\tsin, stars with lower \\teff in a given metallicity regime are typically the stars with the lowest Li abundances, an effect that can be seen even in the sample stars with the lowest metallicities ([Fe/H] $\\sim -3$). This is ultimately so because the coolest stars are typically the least massive, and therefore have been more depleted in Li (see Sect. 4.3). In Fig. 3 we show the Li abundance for cutoffs = 5700 K (open circles), 6100 K (filled squares) and 6350 K (filled triangles). Using a hotter cutoff is useful to eliminate the most Li-depleted stars at low metallicities, but it removes from the Spite plateau stars with [Fe/H] $>$ -2. Imposing a hotter \\teff cutoff at low metallicities and a cooler cutoff at high metallicities eliminates the most Li-depleted stars at low metallicities, but keeps the most metal-rich stars in the Spite plateau. We propose such a metallicity-dependent cutoff below. \\subsection{Two flat Spite plateaus} Giving the shortcomings of a constant \\teff cutoff, we propose an empirical cutoff of \\teff $=$ 5850 - 180$\\times$[Fe/H]. The stars above this cutoff are shown as stars in the bottom panel of Fig. 3. Our empirical cutoff excludes only the most severely Li-depleted stars, i.e., the stars that remain in the Spite plateau may still be affected by depletion. The less Li-depleted stars in the bottom panel of Fig. 3 show two well-defined groups separated at [Fe/H] $\\sim-2.5$ (as shown below, this break represents a real discontinuity), which have essentially zero slopes (within the error bars) and very low star-to-star scatter in their Li abundances. The first group has $-2.5 \\leq$ [Fe/H] $< -1.0$ and $<$A$_{\\rm Li}$$>_1$ = 2.272 ($\\sigma$=0.051) dex and a slope of 0.018$\\pm$0.026, i.e., flat within the uncertainties. The second group is more metal poor ([Fe/H] $< -2.5)$ and has $<$A$_{\\rm Li}$$>_2$ = 2.184 dex ($\\sigma = 0.036$) dex. The slope of this second group is also zero ($-$0.008$\\pm$0.037). Adopting a more conservative exponential cutoff obtained from $Y^2$ isochrones (Demarque et al. 2004), which for a 0.79 M$_\\odot$ star can be fit by \\teff = 6698 $-$2173 $\\times$ $e^{\\rm [Fe/H]/1.021}$, we would also recover a flat Spite plateau, although only stars with [Fe/H] $> -$2.5 are left using this more restrictive cut-off. Thus, the flatness of the Spite plateau is independent of applying a linear or an exponential cutoff. Adopting a constant cutoff in \\teff we also find flat plateaus. For example adopting a cutoff of \\teff $>$ 6100 K (filled squares in Fig. 3) we find in the most metal-rich plateau ([Fe/H] $\\geq -2.5$) no trend between Li and [Fe/H] (slope = 0.019$\\pm$0.025, Spearman rank correlation coefficient $r_{\\rm Spearman}$=0.1 and a probability of 0.48 (i.e., 48\\%) of a correlation arising by pure chance for [Fe/H] $\\geq -$2.5), while for the most metal-poor plateau ([Fe/H] $< -$2.5) we also do not find any trend within the errors (slope = 0.058$\\pm$0.072, $r_{\\rm Spearman}$=0.2 and 41\\% probability of a spurious correlation). Using a hotter cutoff (\\teff $>$ 6350 K, filled triangles in Fig. 3) we obtain also two flat plateaus with slope = -0.040$\\pm$0.063 ($r_{\\rm Spearman}$ = -0.2, probability = 60\\%) for [Fe/H] $\\geq -$2.5 and slope = 0.008$\\pm$0.035 ($r_{\\rm Spearman}$ = 0.1, probability = 68\\%) for [Fe/H] $< -$2.5. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{breakli.eps}} \\caption{(a) filled circles show the slope/$\\sigma$ of $A_{\\rm Li}$ vs. [Fe/H] in the range X$_{\\rm min}$ $<$ [Fe/H] $<-1.0$, where X$_{\\rm min}$ is in the interval [-3.5,-1.25]. For all values of X$_{\\rm min}$ $\\geq -2.5$, the slope is insignificant ($<<$ 3$\\sigma$); only when stars with [Fe/H] $< -2.5$ are included, a slope is forced in $A_{\\rm Li}$ vs. [Fe/H]. Open circles show the slope/$\\sigma$ for the range $-3.25<$ [Fe/H] $<$ X$_{\\rm max}$, where $-$3.0 $\\leq$ X$_{\\rm max}$ $\\leq -$1.0. For X$_{\\rm max}$ $< -2.5$ the slope is negligible ($<<$ 3$\\sigma$), but when stars with [Fe/H] $\\geq -2.5$ are included, a slope is generated. (b) The correlation coefficient $r_{\\rm Spearman}$ is shown by triangles, with filled and open symbols with similar meaning as in panel (a). The probabilities of a correlation between $A_{\\rm Li}$ and [Fe/H] by pure chance, associated to the filled and open triangles, are shown by solid and dashed lines, respectively. No meaningful correlation (probability $<<$ 1) exists for either [Fe/H] $< -$2.5 or [Fe/H] $\\geq -$2.5. } \\label{libreak} \\end{figure} In Fig. 4 we demonstrate that the break at [Fe/H] $\\sim -2.5$ is statistically significant. In panel (a) the slope/$\\sigma$ in the $A_{\\rm Li}$ vs. [Fe/H] plot for the range X$_{\\rm min}$ $<$ [Fe/H] $<-1.0$ are shown as filled circles, where X$_{\\rm min}$ varies within [-3.50,-1.25]. For X$_{\\rm min}$ $\\geq -2.5$, the slope is insignificant ($<<$ 3$\\sigma$), and only when stars with [Fe/H] $< -2.5$ are included a measurable slope is forced in the $A_{\\rm Li}$ vs. [Fe/H] relation. The opposite test is shown by open circles, where we show slope/$\\sigma$ for the range $-$3.25 $<$ [Fe/H] $<$ X$_{\\rm max}$, where X$_{\\rm max}$ changes from [$-$3.00,$-$1.00]. For X$_{\\rm max}$ $< -2.5$ the slope is negligible ($<<$ 3$\\sigma$), and only when stars with [Fe/H] $\\geq -2.5$ are included, a slope is produced between $A_{\\rm Li}$ and [Fe/H]. The correlation coefficient $r_{\\rm Spearman}$ and the probability of a correlation between $A_{\\rm Li}$ and [Fe/H] by pure chance are shown in panel (b). Again, this plot shows that no correlation between $A_{\\rm Li}$ and [Fe/H] exists {\\it within} the two groups ($-3.5<$ [Fe/H] $<-2.5$ and $-2.5\\geq$ [Fe/H] $\\geq-1.0$), and that only when stars {\\it between} the two groups are mixed, significant (probability $\\sim$ 0) correlations of $A_{\\rm Li}$ with [Fe/H] are generated. Systematically lower E(B-V) values (by $\\sim$0.03 mags) in the most metal-poor plateau ([Fe/H] $< -$2.5) could produce a more Li-depleted plateau. However, those E(B-V) values (Fig.~1) are actually slightly higher ($\\sim$0.007 mags, i.e., $\\sim$33 K) than those for [Fe/H] $\\geq -$2.5 (mainly due to a high number of unreddened nearby more metal-rich stars), thus not explaining the existence of two different plateaus. Previous claims of a steep monotonic decrease in Li abundance with decreasing metallicity (e.g. Ryan et al. 1999; A06; B07; Hosford et al. 2009) are probably due to the mix of stars from the two different groups, forcing a monotonic dependence with metallicity. Our large sample of homogeneous and precise Li abundances that covers a broad metallicity range ($-3.5 <$ [Fe/H] $< -1.0$) does not support these claims. Nevertheless, a hint of two different groups in the Spite plateau was already found by A06, who found a change in the slope of the Spite plateau at [Fe/H] $\\approx$ $-$2.2. Also, in the combined A06+B07 sample (Fig. 7 of B07), there are two different groups: stars with [Fe/H] $\\gtrsim -2.6$ have $A_{\\rm Li}$ $>$ 2.2, while stars with [Fe/H] $\\lesssim -2.6$ have $A_{\\rm Li}$ $<$ 2.2. Although in the study by MR04 a flat Spite plateau is found in the range $-3.4 <$ [Fe/H] $< -1$, this is due to the overestimation of \\teff below [Fe/H] $< -2.5$, thus overestimating $A_{\\rm Li}$ at low metallicities and forcing a flat plateau from [Fe/H] = $-3.4$ to $-$1. \\subsection{Correlation between Li and mass} Models of Li depletion predict that the least massive stars are the most depleted in Li, but due to the limitations of previous samples, these predictions have not been thoroughly tested at different metallicity regimes in metal-poor stars.\\footnote{Except for the work of B05, who provide comparisons at different metallicities but for cool severely Li-depleted dwarfs, i.e. probably of lower mass than most stars shown in Fig. 5} In Fig. 5 we show our Li abundances as a function of stellar mass for different metallicity ranges. As can be seen, the Li plateau stars have a clear dependence with mass for all metallicity regimes. Excluding stars with mass $<$ 0.7M$_\\odot$ (including those stars will result in even stronger correlations), linear fits result in slopes of 6, 3, 2, 2 dex M$_\\odot^{-1}$ for stars in the metallicity ranges $-1.3 <$ [Fe/H] $< -1.0$, $-1.6 <$ [Fe/H] $< -1.3$, $-2.5 <$ [Fe/H] $< -1.6$, and $-3.5 <$ [Fe/H] $< -2.5$. The slopes are significant at the 8, 2, 5, 1 $\\sigma$ level, respectively. The correlation coefficient $r_{\\rm Spearman}$ is 0.9, 0.6, 0.6, 0.3, and the probability of a correlation between $A_{\\rm Li}$ and mass arising by pure chance is very small: 5$\\times 10^{-5}$, 3$\\times 10^{-2}$, 1$\\times 10^{-3}$, and 1.3$\\times 10^{-1}$, for stars in the same metallicity ranges as above. Thus, the correlations of $A_{\\rm Li}$ and mass in different metallicity regimes are very significant. Recently, Gonz\\'alez Hern\\'andez et al. (2008, hereafter G08) have studied the metal-poor ([Fe/H] $\\sim-$3.5) double-lined spectroscopic binary BPS CS22876-032, providing thus crucial data to test our Li-mass trend. With our method and the stellar parameters of G08, we obtain a mass ratio of 0.89, very close to their value (0.911$\\pm$0.022) obtained from an orbital solution. For the primary we obtain $M_A$ = 0.776 M$_\\odot$, and adopting the mass ratio of G08, $M_B$ = 0.707 M$_\\odot$ is obtained for the secondary. The LTE Li abundances were taken from G08 and corrected for NLTE effects ($\\sim-$0.05 dex). The components of the binary are shown as circles in Fig. 5, nicely following the trend of the most metal-poor stars. Including this binary in our sample would strengthen the Li-mass correlation of stars with [Fe/H] $< -$2.5. A slope = 3 dex M$_\\odot^{-1}$ significant at the 3$\\sigma$ level is obtained, with $r_{\\rm Spearman}$ = 0.5 and a low probability (2$\\times 10^{-2}$) of the trend being spurious. While mass-dependent Li depletion is expected from standard models of stellar evolution, this should only occur at significantly lower masses than considered here. These stellar models only predict very minor $^7$Li depletion ($\\la 0.02$\\,dex) for metal-poor turn-off stars (e.g. Pinsonneault et al. 1992), which is far from sufficient to explain the $\\sim$0.5\\,dex discrepancy between the observed Li abundance and predictions from BBN+WMAP. Bridging this gap would thus require invoking stellar models that include additional processes normally not accounted for, such as rotationally-induced mixing or diffusion. In Fig. 5 we confront the predictions of Richard et al. (2005) with our inferred stellar masses and Li abundances. The models include the effects of atomic diffusion, radiative acceleration, and gravitational settling, but moderated by a parameterized turbulent mixing (T6.0, T6.09, and T6.25, where higher numbers mean higher turbulence); so far only the predictions for [Fe/H]=$-$2.3 are available for different turbulent mixing models. The agreement is very good when adopting a turbulent model of T6.25 (see Richard et al. for the meaning of this notation) and an initial $A_{\\rm Li} = 2.64$; had the Kurucz convective overshooting models been adopted, the required initial abundance to explain our observational data would correspond to $A_{\\rm Li} = 2.72$. Two other weaker turbulence models that produce smaller overall Li depletions are also shown in Fig. 5, but they are less successful in reproducing the observed Li abundance pattern. Our best-fitting turbulence model (T6.25) is different from that (T6.0) required to explain the abundance pattern in the globular cluster NGC\\,6397 at similar metallicity (Korn et al. 2006; Lind et al. 2009b). Another problem with adopting this high turbulence is that the expected corresponding $^6$Li depletion would amount to $>1.6$\\,dex and thus imply an initial $^6$Li abundance at least as high as the primordial $^7$Li abundance if the claimed $^6$Li detections in some halo stars (A06; AM08) are real. Our results imply that the Li abundances observed in Li plateau stars have been depleted from their original values and therefore do not represent the primordial Li abundance. It appears that the observed Li abundances in metal-poor stars can be reasonably well reconciled with the predictions from standard Big Bang nucleosynthesis (e.g. Cyburt et al. 2008) by means of stellar evolution models that include Li depletion through diffusion and turbulent mixing (Richard et al. 2005). We caution however that although encouraging, our results should not be viewed as proof of the Richard et al. models until the free parameters required for the stellar modelling are better understood from physical principles. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{masslimodel.eps}} \\caption{Li abundances as a function of stellar mass in different metallicity ranges. The metal-poor ([Fe/H] $\\sim-$3.5) binary BPS CS22876-032 (G08) is represented by open circles, nicely fitting the Li-mass trend. Models at [Fe/H] = $-2.3$ including diffusion and T6.0 (short dashed line), T6.09 (dotted line) and T6.25 (solid line) turbulence (Richard et al. 2005) are shown. The models were rescaled to an initial $A_{\\rm Li}$=2.64 (long dashed line) and by $\\Delta M = +0.05$\\,M$_\\odot$. } \\label{massli} \\end{figure}" }, "1005/1005.0346_arXiv.txt": { "abstract": "Using a 3D GCM, we create dynamical model atmospheres of a representative transiting giant exoplanet, HD 209458b. We post-process these atmospheres with an opacity code to obtain transit radius spectra during the primary transit. Using a spectral atmosphere code, we integrate over the face of the planet seen by an observer at various orbital phases and calculate light curves as a function of wavelength and for different photometric bands. The products of this study are generic predictions for the phase variations of a zero-eccentricity giant planet's transit spectrum and of its light curves. We find that for these models the temporal variations in all quantities and the ingress/egress contrasts in the transit radii are small ($< 1.0$\\%). Moreover, we determine that the day/night contrasts and phase shifts of the brightness peaks relative to the ephemeris are functions of photometric band. The $J$, $H$, and $K$ bands are shifted most, while the IRAC bands are shifted least. Therefore, we verify that the magnitude of the downwind shift in the planetary ``hot spot\" due to equatorial winds is strongly wavelength-dependent. The phase and wavelength dependence of light curves, and the associated day/night contrasts, can be used to constrain the circulation regime of irradiated giant planets and to probe different pressure levels of a hot Jupiter atmosphere. We posit that though our calculations focus on models of HD 209458b similar calculations for other transiting hot Jupiters in low-eccentricity orbits should yield transit spectra and light curves of a similar character. ", "introduction": "\\label{intro} Circa April 2010, more than 70 exoplanets had been discovered transiting their primary stars\\footnote{see J. Schneider's Extrasolar Planet Encyclopaedia at http://exoplanet.eu, the Geneva Search Programme at http://exoplanets.eu, and the Carnegie/California compilation at http://exoplanets.org}. Transits are useful in the characterization of exoplanets because they break the the mass/orbital-inclination degeneracy of radial-velocity measurements. Moreover, a planet's radius can be directly measured, yielding mass (\\mp) and radius (\\rp) correlations with which theorists can extract useful information about bulk composition and structure and can attempt to fit radius evolution models (Guillot et al. 1996; Burrows et al. 2000; Brown et al. 2001; Hubbard et al. 2001; Baraffe et al. 2003; Chabrier et al. 2004; Charbonneau et al. 2007). However, since atmospheric opacity is a function of wavelength, a planet's transit radius is also a function of wavelength (Seager \\& Sasselov 2000; Brown 2001; Spiegel et al. 2007; Fortney et al. 2003,2010). The variation with wavelength in the measured radius (actually an impact parameter) provides an ersatz spectrum directly related to the planet's atmospheric composition near the terminator. This radius spectrum, unlike the planet's spectrum at secondary eclipse, is more dependent upon composition than upon the temperature profile, and so provides complementary information to a planet's own direct emissions. The latter can be measured during secondary eclipse, but also during the traverse by the planet of its orbit as it traces out its phase light curve. The transit radius spectrum and the planet emission spectrum as a function of phase can, therefore, together help constrain planet properties (Fortney et al. 2006,2010). However, the stellar irradiation of such an exoplanet severely breaks what would otherwise be quasi-spherical symmetry, producing large day-night contrasts in thermal structure, zonal flows and banding, and violent atmospheric dynamics (Showman \\& Guillot 2002; Guillot \\& Showman 2002; Cho et al. 2003,2008; Burkert et al. 2005; Cooper \\& Showman 2005,2006; Showman et al. 2008,2009; Showman, Cho, \\& Menou 2010; Dobbs-Dixon \\& Lin 2008; Dobbs-Dixon, Cumming, \\& Lin 2010; Rauscher et al. 2007,2008; Menou \\& Rauscher 2009; Rauscher \\& Menou 2010). Importantly, such behavior may have photometric and spectral signatures. It is the exploration of such signatures that motivates this paper. The state-of-the-art in the 3D modeling of exoplanet atmospheres is still evolving and has not yet reached a level of maturity where detailed predictions for each known exoplanet are robust. Hence, in this paper we focus on a few generic ideas and conjectures that emerge from our modeling efforts. The procedure we have pursued is the following: First, using a 3D general circulation model (GCM) we derive dynamic model atmospheres at various epochs after they have achieved a steady state. Second, we post-process these 3D model atmospheres with a spectral atmosphere code to obtain transit radius spectra at various epochs during ingress, egress, and total transit. We assume equilibrium molecular compositions and solar metallicity. Third, using the same 3D model atmosphere, we integrate over the ``visible\" disk and calculate phase light curves as a function of wavelength and for various standard photometric bands. The results are generic predictions which, though not expected to be quantitatively precise and constraining on any particular giant exoplanet, nevertheless contain qualitatively interesting features that should inform future measurements. For specificity, we focus on two models of HD 209458b (Charbonneau et al. 2000; Henry et al. 2000), one with and one without an ``extra absorber\" and a thermal inversion (Hubeny, Burrows, \\& Sudarsky 2003; Burrows et al. 2007; Fortney et sl. 2008; Spiegel et al. 2009; Knutson et al. 2008), but suggest our qualitative results are generic beyond this planet and this modeling paradigm (see also Fortney et al. 2006,2010). In \\S\\ref{techniques}, we summarize our methodology and techniques. Then, in \\S\\ref{describe} we describe the 3D models and various of their salient characteristics. We go on in \\S\\ref{wave_rad} to present our results for the wavelength-dependent transit radius and ingress-egress asymmetries and in \\S\\ref{phase} we turn to a full discussion of the derived light curves. This section contains not only wavelength-dependent planet-star flux ratios as a function of phase, but the phase variation of several photometric planet band fluxes. In \\S\\ref{conclusions}, we summarize our general conclusions. ", "conclusions": "\\label{conclusions} Using a 3D GCM with Newtonian cooling and day- and nightside equilibrium temperature profiles based on two 1D spectral models (with and without a thermal inversion and a hot upper atmosphere), we created dynamical models of the transiting giant exoplanet HD 209458b. We post-processed these 3D model atmospheres with a detailed opacity code to obtain transit radius spectra at various epochs during the primary transit. Then, using a spectral atmosphere code, we integrated over the face of the planet seen by an observer at various orbital phases and calculated light curves as a function of wavelength and for nine different photometric bands from $R$ ($\\sim$0.6 \\mic) through IRAC 4 ($\\sim$8 \\mic). The products of this study are generic predictions for the character of the expected phase variations both of a giant planet's transit spectrum and of its light curves. Since our GCM employed Newtonian cooling (and not radiative transfer using opacities that corresponded with those used in the post-processing), the calculations are slightly inconsistent. Nevertheless, the results have interesting features that should inform future measurements. We found that the temporal variations in the derived integral quantities for this model suite are small (less than 1\\%) and that the ingress/egress contrasts due to zonal flows, while also small ($< 0.5 - 1.0$ \\%), are most manifest due to the scale-height asymmetries near the terminators and in the water bands. While models we have generated with and without the methane and carbon monoxide bands indicate that ingress/egress differences in the CO and CH$_4$ bands might be diagnostic of both carbon chemistry and abundance asymmetries due to zonal flows near the terminators, the differential effects due to different carbon-species abundances at the different terminators are small (less than a few tenths of a percent). A similar conclusion was reached by Fortney et al. (2010), who nevertheless found a slightly larger quantitative effect. Since the transit radius of HD 209458b has been measured in the optical, it should not be allowed to differ when comparing models with and without an upper-atmosphere absorber and the predicted radii at other wavelengths must be determined relative to it. When this is done, the predicted radii in the near- and mid-infrared are very different (by as much as 0.05 \\rj), though our current models do a poor job of fitting all the transit radius data for HD 209458b simultaneously (e.g., Knutson et al. 2007; Beaulieu et al. 2010; Richardson et al. 2006). The reason for this discrepancy is currently unknown. We determined that the angular phase shifts of the brightness peaks relative to the orbital ephemeris are functions of photometric band. The $J$, $H$, and $K$ bands are shifted most, by as much as $\\sim$45$^{\\circ}$, while the IRAC bands are shifted least. This is because different bands are spectrally formed at different pressure depths and the zonal winds that advect heat are depth-dependent. The IRAC band photospheres are generally at altitude, while those in the near infrared are at deeper pressures near and above $\\sim$0.1 bars. Therefore, the question of the magnitude of the downwind shift in the planetary ``hot spot\" due to equatorial winds is nuanced, depending upon the waveband in which measurements are made. This wavelength dependence of the phase shift in the brightness light curve, and the associated day/night contrasts, can in principle be used to constrain the circulation regimes of irradiated giant planets and to probe different pressure levels of a hot Jupiter atmosphere. Though our calculations focused on models of HD 209458b, similar calculations for other transiting hot Jupiters in low-eccentricity orbits should yield transit spectra and light curves of a similar character. As the subject of comparative exoplanetology matures, and JWST comes online, such observational manifestations of global circulation may well become possible for a wide range of irradiated planets. Our models have been constructed to help prepare the way, however imperfectly, for that era." }, "1005/1005.4639.txt": { "abstract": "We investigate whether or not the decadal and multi-decadal climate oscillations have an astronomical origin. Several global surface temperature records since 1850 and records deduced from the orbits of the planets present very similar power spectra. Eleven frequencies with period between 5 and 100 years closely correspond in the two records. Among them, large climate oscillations with peak-to-trough amplitude of about 0.1 $^oC$ and 0.25 $^oC$, and periods of about 20 and 60 years, respectively, are synchronized to the orbital periods of Jupiter and Saturn. Schwabe and Hale solar cycles are also visible in the temperature records. A 9.1-year cycle is synchronized to the Moon's orbital cycles. A phenomenological model based on these astronomical cycles can be used to well reconstruct the temperature oscillations since 1850 and to make partial forecasts for the 21$^{st}$ century. It is found that at least 60\\% of the global warming observed since 1970 has been induced by the combined effect of the above natural climate oscillations. The partial forecast indicates that climate may stabilize or cool until 2030-2040. Possible physical mechanisms are qualitatively discussed with an emphasis on the phenomenon of collective synchronization of coupled oscillators.\\\\ \\\\ Please cite this article as: Scafetta, N., Empirical evidence for a celestial origin of the climate oscillations and its implications. \\emph{Journal of Atmospheric and Solar-Terrestrial Physics} (2010), doi:10.1016/j.jastp.2010.04.015 ", "introduction": "Milankovic [1941] theorized that variations in eccentricity, axial tilt, and precession of the orbit of the Earth determine climate patterns such as the 100,000 year ice age cycles of the Quaternary glaciation over the last few million years. The variation of the orbital parameters of the Earth is due to the gravitational perturbations induced by the other planets of the solar system, primarily Jupiter and Saturn. Over a much longer time scale the cosmic-ray flux record well correlates with the warm and ice periods of the Phanerozoic during the last 600 million years: the cosmic-ray flux oscillations are likely due to the changing galactic environment of the solar system as it crosses the spiral arms of the Milky Way [Shaviv, 2003, 2008; Shaviv and Veizer, 2003; Svensmark, 2007]. Over millennial and secular time scales several authors have found that variations in total solar irradiance and variations in solar modulated cosmic-ray flux well correlate with climate changes: see for example: Eddy, 1976; Hoyt and Schatten, 1997; White \\emph{et al.}, 1997; van Loon and Labitzke, 2000; Bond \\emph{et al.}, 2001; Kerr, 2001; Douglass and Clader, 2002; Kirkby, 2007; Scafetta and West, 2005, 2007, 2008; Shaviv, 2008; Eichler \\emph{et al.}, 2009; Soon, 2009; Meehl \\emph{et. al.}, 2009; Scafetta 2009, 2010. Also the annual cycle has an evident astronomical origin. The above results suggest that the dominant drivers of the climate oscillations have a celestial origin. Therefore, it is legitimate to investigate whether the climate oscillations with a time scale between 1 and 100 years, can be interpreted in astronomical terms too. Global surface temperature has risen [Brohan \\emph{et al.}, 2006] by about 0.8 $^oC$ and 0.5 $^oC$ since 1900 and 1970, respectively. Humans may have partially contributed to this global warming through greenhouse gas (GHG) emissions [IPCC, 2007]. For instance, the IPCC claims that more than 90\\% of the observed warming since 1900 and practically 100\\% of the observed warming since 1970 have had an anthropogenic origin (see figure 9.5 in IPCC, AR4-WG1). The latter conclusion derives merely from the fact that climate models referenced by the IPCC cannot explain the warming occurred since 1970 with any known natural mechanism. Therefore, several scientists have concluded that this warming has been caused by anthropogenic GHG emissions that greatly increased during this same period. This theory is known as the \\emph{anthropogenic global warming theory} (AGWT). However, the anthropogenic GHG emissions have increased monotonically since 1850 while the global temperature record did not. Several oscillations are seen in the data since 1850, including a global cooling since 2002: see Figure 1. If these climate oscillations are natural, for example induced by astronomical oscillations, they would determine how climate change should be interpreted [Keenlyside \\emph{et al.}, 2008]. In fact, during its cooling phase a natural multi-decadal oscillation can hide a global warming caused by human GHG emissions or, alternatively, during its warming phase a natural oscillation can accentuate the warming. If the natural oscillations of the climate are not properly recognized and taken into account, important climate patterns, for example the global warming observed from 1970 to 2000, can be erroneously interpreted. Indeed, part of the 1970-2000 warming could have been induced by a multi-decadal natural cycle during its warming phase that the climate models used by the IPCC have not reproduced. The IPCC [2007] claims that the climate oscillations are induced by some still poorly understood and modeled internal dynamics of the climate system, such as the ocean dynamics. However, the oscillations of the atmosphere and of the ocean, such as the Pacific Decadal Oscillation (PDO) and the Atlantic Multidecadal Oscillation (AMO), may be induced by complex extraterrestrial periodic forcings that are acting on the climate system in multiple ways. Indeed, the climate system is characterized by interesting cyclical patterns that remind astronomical cycles. \\begin{figure}[t!] \\includegraphics[angle=-90,width=21pc]{figure1.eps} \\caption{ Top: Global surface temperature anomaly (gray) [Brohan \\emph{et al.}, 2006] against the GISS ModelE average simulation (blak) [Hansen \\emph{et al.}, 2007]. The figure also shows the quadratic upward trend of the temperature. Bottom: an eight year moving average smooth of the temperature detrended of its upward quadratic trend. This smooth reveals a quasi-60 year modulation. } \\end{figure} For example, surface temperature records are characterized by decadal and bi-decadal oscillations which are usually found in good correlation with the (11-year) Schwabe and the (22 year) Hale solar cycles [Hoyt and Schatten, 1997; Scafetta and West, 2005; Scafetta, 2009]. However, longer cycles are of interest herein. Klyashtorin and Lyubushin [2007] and Klyashtorin et al. [2009] observed that several centuries of climate records (ice core sample, pine tree samples, sardine and anchovy sediment core samples, global surface temperature records, atmospheric circulation index, length of the day index, fish catching productivity records, etc.) are characterized by large 50-70 year and 30-year periodic cycles. The quasi-60 year periodicity has been also found in secular monsoon rainfall records from India, in proxies of monsoon rainfall from Arabian Sea sediments and in rainfall over east China [for example see the following works and their references: Agnihotri \\emph{et al.}, 2002; Sinha A. \\emph{et al.} (2005); Goswami \\emph{et al.}, 2006; Yadava and Ramesh 2007]. Thus, several records indicate that the climate is characterized by a large quasi-60 year periodicity, plus larger secular climatic cycles and smaller decadal cycles. All these cycles cannot be explained with anthropogenic emissions. Errors in the data, other superimposed patterns (for example, volcano effects and longer and shorter cycles) and some chaotic pattern in the dynamics of these signals may sometimes mask the 60-year cycle. A multi-secular climatic record that shows a clear quasi-60 year oscillation is depicted in Figure 2: the G. Bulloides abundance variation record found in the Cariaco Basis sediments in the Caribbean sea since 1650 [Black \\emph{et al.}, 1999]. This record is an indicator of the trade wind strength in the tropical Atlantic ocean and of the north Atlantic ocean atmosphere variability. This record shows five 60-year large cycles. These cycles correlate well with the 60-year modulation of the global temperature observed since 1850 (the correlation is negative). On longer time scales, periods of high G. Bulloides abundance correlate well with periods of reduced solar output (the well-known Maunder, Sp\\\"orer, and Wolf minima), suggesting a solar forcing origin of these cycles [Black \\emph{et al.}, 1999]. \\begin{figure}[t!] \\includegraphics[angle=-90,width=21pc]{figure2.eps} \\caption{ Record of G. Bulloides abundance variations (1-mm intervals) from 1650 to 1990 A.D. (black line) [Black \\emph{et al.}, 1999]. The gray vertical lines highlight 60-year intervals. Five quasi-60 year cycles are seen in this record, which is a proxy for the Atlantic variability since 1650.} \\end{figure} Patterson et al. (2004) found 60-62 year cycles in sediments and cosmogenic nuclide records in the NE Pacific. Komitov (2009) found similar cycles in the number of the middle latitude auroras from 1700 to 1900. A cycle of about 60 years has been detected in the number of historically recorded meteorite falls in China from AD 619 to 1943 and in the number of witnessed falls in the world from 1800 to 1974 [Yu \\emph{et al.}, 1983]. Ogurtsov \\emph{et al.} [2002] found a 60-64 year cycle in $^{10}Be$, $^{14}C$ and Wolf number over the past 1000 years. The existence of a 60-year signal has been found in the Earth's angular velocity and in the geomagnetic field [Roberts \\emph{et al.}, 2007]. These results clearly suggest an astronomical origin of the 60-year variability found in several climatic records. Interestingly, the traditional Chinese calendar, whose origins can be traced as far back as the 14$^{th}$ century BCE, is arranged in major 60-year cycles [Aslaksen, 1999]. Each year is assigned a name consisting of two components. The first component is one of the 10 \\emph{Heavenly Stems} (\\emph{Jia, Yi, Bing}, etc.), while the second component is one of the 12 \\emph{Earthly Branches} that features the names of 12 animals (\\emph{Zi, Chou, Yin}, etc.). Every 60 years the stem-branch cycle repeats. Perhaps, this sexagenary cyclical calendar was inspired by climatic and astronomical observations. Some studies [Jose, 1965; Landscheidt, 1988, 1999; Charv\\'atov\\'a, 1990, 2009; Charv\\'atov\\'a and St\\v{r}e\\v{s}t\\'ik, 2004; Mackey, 2007; Wilson \\emph{et al.}, 2008; Hung, 2007] suggested that solar variation may be driven by the planets through gravitational spin-orbit coupling mechanisms and tides. These authors have used the inertial motion of the Sun around the center of mass of the solar system (CMSS) as a proxy for describing this phenomenon. Then, a varying Sun would influence the climate by means of several and complicated mechanisms and feedbacks [Idso and Singer, 2009]. Indeed, tidal patterns on the Sun well correlate with large solar flare occurrences, and the alignment of Venus, Earth and Jupiter well synchronizes with the 11-year Schwabe solar cycle [Hung, 2007]. In addition, the Earth-Moon system and the Earth's orbital parameters can also be directly modulated by the planetary oscillating gravitational and magnetic fields, and synchronize with their frequencies [Scafetta, 2010]. The Moon can influence the Earth through gravitational tides and orbital oscillations [Keeling and Whorf, 1997, 2000; Munk and Wunsch, 1998; Munk and Bills, 2007]. It could be argued that planetary tidal forces are \\emph{weak} and unlikely have any physical outcome. It can also be argued that the tidal forces generated by the terrestrial planets are comparable or even larger than those induced by the massive jovian planets. However, this is not a valid physical rationale because still little is known about the solar dynamics and the terrestrial climate. Complex systems are usually characterized by feedback mechanisms that can amplify the effects of weak periodic forcings also by means of resonance and collective synchronization processes [Kuramoto, 1984; Strogatz, 2009]. Thus, unless the physics of a system is not clearly understood, good empirical correlations at multiple time scales cannot be dismissed just because the microscopic physical mechanisms may be still obscure and need to be investigated. The above theory implies the existence of direct and/or indirect links between the motion of the planets and the climate oscillations, essentially claiming that the climate is synchronized to the natural oscillations of the solar system, which are driven by the movements of the planets around the Sun. If this theory is correct, it can be efficiently used for interpreting climate changes and forecasting climate variability because the motion of the planets can be rigorously calculated. In this paper we investigate this theory by testing a synchronization hypothesis, that is, whether the planetary motion and the climate present a common set of frequencies. Further, we compare the statistical performance of a phenomenological planetary model for interpreting the climate oscillations with that of a typical major general circulation model adopted by the IPCC. Our findings show that a planetary-based climate model would largely outperform the traditional one in reconstructing the oscillations observed in the climate records. ", "conclusions": "On secular, millenarian and larger time scales astronomical oscillations and solar changes drive climate variations. Shaviv's theory [2003] can explain the large 145 Myr climate oscillations during the last 600 million years. Milankovic's theory [1941] can explain the multi-millennial climate oscillations observed during the last 1000 kyr. Climate oscillations with periods of 2500, 1500, and 1000 years during the last 10,000 year (the Holocene) are correlated to equivalent solar cycles that caused the Minoan, Roman, Medieval and Modern warm periods [Bond \\emph{et al.}, 2001; Kerr, 2001]. Finally, several other authors found that multisecular solar oscillations caused bi-secular little ice ages (for example: the Sp\\\"orer, Maunder, Dalton minima) during the last 1000 years [for example: Eddy, 1976; Eichler \\emph{et al.}, 2009; Scafetta and West, 2007; Scafetta, 2009, 2010]. Herein, we have found empirical evidences that the climate oscillations within the secular scale are very likely driven by astronomical cycles, too. Cycles with periods of 10-11, 12, 15, 20-22, 30 and 60 years are present in all major surface temperature records since 1850, and can be easily linked to the orbits of Jupiter and Saturn. The 11 and 22-year cycles are the well-known Schwabe and Hale solar cycles. Other faster cycles with periods between 5 and 10 years are in common between the temperature records and the astronomical cycles. Long-term lunar cycles induce a 9.1-year cycle in the temperature records and probably other cycles, including an 18.6-year cycle in some regions [McKinnell and Crawford, 2007]. A quasi-60 year cycle has been found in numerous multi-secular climatic records, and it is even present in the traditional Chinese, Tibetan and Tamil calendars, which are arranged in major 60-year cycles. The physical mechanisms that would explain this result are still unknown. Perhaps the four jovian planets modulate solar activity via gravitational and magnetic forces that cause tidal and angular momentum stresses on the Sun and its heliosphere. Then, a varying Sun modulates climate, which amplifies the effects of the solar input through several feedback mechanisms. This phenomenon is mostly regulated by Jupiter and Saturn, plus some important contribution from Neptune and Uranus, which modulate a bi-secular cycle with their 172 year synodic period. This interpretation is supported by the fact that the 11-year solar cycles and the solar flare occurrence appear synchronized to the tides generated on the Sun by Venus, Earth and Jupiter [Hung, 2007]. Moreover, a 60-year cycle and other planetary cycles have been found in millennial solar records [Ogurtsov \\emph{et al.}, 2002] and in the number of middle latitude auroras [Komitov, 2009]. Alternatively, the planets are directly influencing the Earth's climate by modulating the orbital parameters of the Earth-Moon system and of the Earth. Orbital parameters can modulate the Earth's angular momentum via gravitational tides and magnetic forces. Then, these orbital oscillations are amplified by the climate system through synchronization of its natural oscillators. This interpretation is supported by the fact that the temperature records contain a clear 9.1-year cycle, which is associated to some long-term lunar tidal cycles. However, the climatic influence of the Moon may be more subtle because several planetary cycles are also found in the Earth-Moon system. The astronomical forcings may be modulating the length of the day (LOD). LOD presents a 60-year cycle that anticipates the 60-year temperature cycle [Klyashtorin 2001; Klyashtorin and Lyubushin, 2007, 2009; Mazzarella, 2007, 2008; Sidorenkov and Wilson, 2009]. A LOD change can drive the ocean oscillations by exerting some pressure on the ocean floor and by modifying the Coriolis' forces. In particular, the large ocean oscillations such as the AMO and PDO oscillations are likely driven by astronomical oscillations. The results herein found show that the climate oscillations are driven by multiple astronomical mechanisms. Indeed, the planets with their movement cause the entire solar system to vibrate with a set of frequencies that are closely related to the orbital periods of the planets. The wobbling of the Sun around the center of mass of the solar system is just the clearest manifestation of these solar system vibrations and has been used herein just as a proxy for studying those vibrations. The Sun, the Earth-Moon system and the Earth feel these oscillations, and it is reasonable that the internal physical processes of the Earth and the Sun synchronize to them. It is evident that we can still infer, by means of a detailed data analysis, that the solar system likely induces the climate oscillations, although the actual mechanisms that explain the observed climate oscillations are still unknown. If the true climate mechanisms were already known and well understood, the general circulation climate models would properly reproduce the climate oscillations. However, we found that this is not the case. For example, we showed that the GISS ModelE fails to reproduce the climate oscillations at multiple time scales, including the large 60-year cycle. This failure is common to all climate models adopted by the IPCC [2007] as it is evident in their figures 9.5 and SPM.5 that show the multi-model global average simulation of surface warming. This failure indicates that the models on which the IPCC's claims are based are still incomplete and possibly flawed. The existence of a 60-year natural cycle in the climate system, which is clearly proven in multiple studies and herein in Figures 2, 6, 10 and 12, indicates that the AGWT promoted by the IPCC [2007], which claims that 100\\% of the global warming observed since 1970 is anthropogenic, is erroneous. In fact, since 1970 a global warming of about 0.5 $^oC$ has been observed. However, from 1970 to 2000 the 60-year natural cycle was in his warming phase and has contributed no less than 0.3 $^oC$ of the observed 0.5 $^oC$ warming, as Figure 10B shows. Thus, at least 60\\% of the observed warming since 1970 has been naturally induced. This leaves less than 40\\% of the observed warming to human emissions. Consequently, the current climate models, by failing to simulate the observed quasi-60 year temperature cycle, have significantly overestimated the climate sensitivity to anthropogenic GHG emissions by likely a factor of three. Moreover, the upward trend observed in the temperature data since 1900 may be partially due to land change use, uncorrected urban heat island effects [McKitrick and Michaels, 2007; McKitrick, 2010] and to the bi-secular and millennial solar cycles that reached their maxima during the last decades [Bond \\emph{et al.}, 2001; Kerr, 2001; Eichler \\emph{et al.}, 2009; Scafetta, 2010]. Solomon \\emph{et al.} [2010] recently acknowledged that stratospheric water vapor, not just anthropogenic GHGs, is a very important climate driver of the decadal global surface climate change. Solomon \\emph{et al.} estimated that stratospheric water vapor has largely contributed both to the warming observed from 1980-2000 (by 30\\%) and to the slight cooling observed after 2000 (by 25\\%). This study reinforces that climate change is more complex than just a response to added $CO_2$ and a few other anthropogenic GHGs. The causes of stratospheric water vapor variation are not understood yet. Perhaps, stratospheric water vapor is driven by UV solar irradiance variations through ozone modulation, and works as a climate feedback to solar variation [Stuber \\emph{ et al.}, 2001]. Thus, Solomon's finding would partially support the findings of this paper and those of Scafetta and West [2005, 2007] and Scafetta [2009]. The latter studies found a significant natural and solar contribution to the warming from 1970-2000 and to the cooling afterward. A detailed reconstruction of the climate oscillations suggests that a model based on celestial oscillations, as shown in Figure 12, would largely outperform current general circulation climate models, such as the GISS ModelE, in reconstructing the climate oscillations. The planetary model would also be more accurate in forecasting climate changes during the next few decades. Over this time, the global surface temperature will likely remain approximately steady, or actually cool. In conclusion, data analysis indicates that current general circulation climate models are missing fundamental mechanisms that have their physical origin and ultimate justification in astronomical phenomena, and in interplanetary and solar-planetary interaction physics." }, "1005/1005.2425_arXiv.txt": { "abstract": "{ We study how the universe reheats following an era of chaotic Dirac--Born--Infeld inflation, and compare the rate of particle production with that in models based on canonical kinetic terms. Particle production occurs through non-perturbative resonances whose structure is modified by the nonlinearities of the Dirac--Born--Infeld action. We investigate these modifications and show that the reheating process may be efficient. We estimate the initial temperature of the subsequent hot, radiation-dominated phase.} \\preprint{arXiv:1005.2425} ", "introduction": "\\label{sec:introduction} Measurements of the microwave background temperature made by the WMAP satellite over the last decade \\cite{Komatsu:2010fb} are consistent with the idea that structure in the universe originated as small fluctuations during a primordial era of accelerated expansion, or `inflation'. Inflation leaves the universe in a cold vacuum state, devoid of radiation and matter. If our own universe experienced an inflationary era in its past, we must conclude that some process repopulated its sterile vacuum with the abundant quanta we see today. Much less is known about this `reheating' process than is understood about the observationally accessible density fluctuation. Properties of inflationary fluctuations largely decouple from the details of whatever physics drives the accelerating epoch, leaving only weak traces.% \\footnote{These traces are presently the subject of intense theoretical and observational effort; see, e.g. Ref.~\\cite{Chen:2010xk}.} This freedom has encouraged many proposals for the underlying microphysics, nearly all of which are equally unconstrained by observation. In an especially interesting proposal, due to Silverstein \\& Tong \\cite{Silverstein:2003hf} and later elaborated by Alishahiha, Silverstein \\& Tong \\cite{Alishahiha:2004eh}, inflation occurs while a $\\D{3}$-brane moves in a warped background spacetime.% \\footnote{D-branes are extended objects in spacetime. For a review of their properties see Johnson \\cite{Johnson:2003gi}; inflation in such models has been reviewed by McAllister \\& Silverstein \\cite{McAllister:2007bg}. Many detailed properties were discussed by Chen~\\cite{Chen:2005ad}.} As we shall discuss in \\S\\ref{sec:review}, the motion of the brane is controlled by an action of Dirac--Born--Infeld (``DBI'') type and the resulting model is known as DBI inflation. In this paper we study the reheating era which must follow such an inflationary phase. Reheating in related brane-world models has been studied in Refs.~\\cite{Cline:2002it, Brodie:2003qv, Dvali:2003ar, Frey:2005jk, Chialva:2005zy}. In contrast to the primordial perturbations, reheating depends strongly on many model-specific details~\\cite{Kofman:1996mv,Kaiser:1995fb}. It is a subtle function of nonlinear physics, probing information complementary to the linear physics of the density fluctuation. Nonlinear properties vary widely between competing inflationary models, and there are important differences between DBI models and those with canonical kinetic terms. Many authors have contributed to the theory of particle production, which is now well-developed \\cite{Traschen:1990sw,Kofman:1994rk,Shtanov:1994ce,Kofman:1997yn}.% \\footnote{The theory of preheating has been reviewed by Allahverdi {\\etal} \\cite{Allahverdi:2010xz}, Kofman {\\etal} \\cite{Kofman:2008zz} and Bassett {\\etal} \\cite{Bassett:2005xm}. See also Ref.~\\cite{Kohri:2009ac}.} In \\emph{perturbative reheating}, the decay rate of inflaton particles into other species of matter is calculated directly from an S-matrix. These calculations assume any relics of previous decays to be sufficiently diffuse that particle production takes place into an effectively pristine vacuum. It was later understood that a Bose enhancement of the decay rate may occur for integer-spin species if the products of previous decays accumulate, leading to a resonant phase of out-of-equilibrium production known as \\emph{preheating}. A different effect, so-called \\emph{tachyonic preheating} \\cite{Felder:2000hj,Felder:2001kt, Dufaux:2006ee}, is associated with the end of inflation in certain models. Unlike conventional preheating this does not merely convert one species of particle into another, but (like any tachyonic instability) reflects a preference to convert one vacuum state to another by the rapid accumulation and condensation of particles. These phenomena may occur in any effective field theory of the post-inflationary universe, whatever its microphysical origin. The general theory has subsequently been applied to concrete models, including some versions of brane inflation, in which it is possible to give a more refined interpretation \\cite{Kachru:2003sx,Barnaby:2004gg}. In these models, inflation ends when the moving $\\D{3}$-brane becomes close to an antibrane (a $\\Dbar{3}$-brane), allowing open strings stretching between the $\\D{3}$--$\\Dbar{3}$ pair to become excited. The open string states include a tachyon which induces fragmentation of the $\\D{3}$ brane into $\\D{0}$ branes, described on the $\\D{3}$ worldvolume by a phase of tachyonic preheating. The $\\D{0}$ fragments decay into closed string states, which are enumerated by the Kaluza--Klein (``KK'') modes of supergravity fields in the warped background. These KK modes subsequently decay into the species of matter and radiation which populate our universe \\cite{Barnaby:2004gg}. In this paper we study a version of the DBI model in which the end of inflation is less dramatic. We work in a model where the $\\D{3}$ comes to rest near an extremity of its warped background, referred to as the ``tip of the throat.'' As it settles on its resting-place it executes coherent oscillations, inducing growing fluctuations which fold and wrinkle its surface. This scenario was suggested by Silverstein \\& Tong \\cite{Silverstein:2003hf}. Our interest is in the \\emph{qualitative} differences which may emerge in concrete models. Should we expect the reheating temperature to increase or decrease? Is parametric resonance more efficient, perhaps due to the extra non-linearities of the DBI action, or do these same non-linearities conspire to shut off the resonant channels? For this reason we focus on differences with the case of canonical kinetic terms and largely ignore similarities. For example, the use of perturbation theory in either case will break down at the onset of back-reaction, when the population of matter species grows to the point where it cannot be ignored. This affects the canonical and DBI cases equally, and we restrict our comparison to times before back-reaction becomes significant. This paper is organized as follows. In \\S\\ref{sec:review} we review the dynamics of inflation in DBI models and discuss a toy model in which inflation ends as the brane settles to its final position in the warped background. This model is comparable to chaotic inflation in a theory with canonical kinetic terms. In \\S\\ref{sec:res} we discuss parametric resonance, making a comparison with canonical chaotic inflation. In \\S\\ref{sec:pr} we study the stage of reheating and estimate the reheating temperature. We conclude in \\S\\ref{sec:conclude}. Throughout this paper, we work in units where $\\hbar = c = 1$ and set the reduced Planck mass to unity, $\\Mp \\equiv (8 \\pi G)^{-1/2} = 1$. The metric signature is taken to be $(-, +, +, +)$. ", "conclusions": "\\label{sec:conclude} We have studied mechanisms by which the universe can be repopulated with matter species following an era of D-brane inflation. We work in the limit where the motion of the brane is at most perturbatively relativistic, which may ultimately be required by observation if $\\fnl$ measured on equilateral configurations of the bispectrum is nonnegative. Our analysis is valid in a regime where the amplitude of oscillations is significantly smaller than whatever infrared scale caps the throat in which the D-brane moves. Indeed, because we take the D-brane to be at most perturbatively relativistic, inflation is of the slow-roll type. We assume inflation ends as the D-brane gently coasts to its resting position, dissipating its remaining energy by executing small amplitude oscillations. The amplitude and phase of these oscillations are mildly perturbated in comparison with the oscillations of canonically normalized scalar reheating the universe after a phase of four-dimensional inflation. We encounter some novel effects which are not present in a theory with canonical kinetic terms. In canonical inflation there is no resonant production of inflaton particles during the reheating era. By contrast, in the perturbatively-relativistic limit of DBI inflation, non-linearities arising from the Dirac--Born--Infeld action cause a weak resonance which reaches an $\\Or(1)$ effect only after $\\sim 1/\\epsilon$ oscillations. Excessive production of inflaton particles is typically problematic, but since preheating typically concludes after $\\lesssim 10$ oscillations this weak resonance is unlikely to be fatal for the DBI model. In the ultra-relativistic limit where $\\gamma \\gg 1$, the potential becomes unimportant and presumably resonance will be strongly suppressed. Another novel effect is a `DBI friction,' which appears in the equation of motion for the inflaton. This slightly damps the preheating effect. Beyond perturbation theory, it is possible this effect represents the decreasing significance of the potential at large $\\gamma$. For a given set of measured masses and couplings which characterize the properties of the inflaton and the matter field into which it decays, we find that final reheating temperatures (assuming they are achieved by perturbative decays) are typically smaller than in the equivalent theory with canonical kinetic terms. This may be beneficial in concrete models, where lower reheating temperatures allow problems associated with overclosure of the universe by moduli or gravitinos to be ameliorated. The most stringent constraint on DBI models comes from observations of the microwave background bispectrum, which presently require $\\gamma \\lesssim 18$. If our results are representative, and the reheating temperature falls as $\\gamma$ increases, a constraint on $\\gamma$ may also emerge from demanding that the nucleosynthesis era is uninterrupted, which requires $T_{\\mathrm{RH}} \\gtrsim \\mbox{MeV}$." }, "1005/1005.2563_arXiv.txt": { "abstract": "{Within the framework of the HERM33ES key project, we are studying the star forming interstellar medium in the nearby, metal-poor spiral galaxy M33, exploiting the high resolution and sensitivity of \\herschel. } {We use PACS and SPIRE maps at 100, 160, 250, 350, and 500\\,$\\mu$m wavelength, to study the variation of the spectral energy distributions (SEDs) with galacto-centric distance. } {Detailed SED modeling is performed using azimuthally averaged fluxes in elliptical rings of 2\\,kpc width, out to 8\\,kpc galacto-centric distance. Simple isothermal and two-component grey body models, with fixed dust emissivity index, are fitted to the SEDs between 24\\,$\\mu$m and 500$\\,\\mu$m using also MIPS/\\spitzer\\, data, to derive first estimates of the dust physical conditions. } {The far-infrared and submillimeter maps reveal the branched, knotted spiral structure of M33. An underlying diffuse disk is seen in all SPIRE maps (250--500\\,$\\mu$m). Two component fits to the SEDs agree better than isothermal models with the observed, total and radially averaged flux densities. The two component model, with $\\beta$ fixed at 1.5, best fits the global and the radial SEDs. The cold dust component clearly dominates; the relative mass of the warm component is less than 0.3\\% for all the fits. The temperature of the warm component is not well constrained and is found to be about 60\\,K$\\pm$10\\,K. The temperature of the cold component drops significantly from $\\sim24$\\,K in the inner 2\\,kpc radius to 13\\,K beyond 6\\,kpc radial distance, for the best fitting model. The gas-to-dust ratio for $\\beta=1.5$, averaged over the galaxy, is higher than the solar value by a factor of 1.5 and is roughly in agreement with the subsolar metallicity of M33. } {} ", "introduction": "% In the local universe, most of the observable matter is contained in stellar objects that shape the morphology and dynamics of their ``parent\" galaxy. In view of the dominance of stellar mass, a better understanding of star formation and its consequences is mandatory. There exists a large number of high spatial resolution studies related to individual star forming regions of the Milky Way, as well as of low linear resolution studies of external galaxies. For a comprehensive view onto the physical and chemical processes driving star formation and galactic evolution it is, however, essential to combine local conditions affecting individual star formation with properties only becoming apparent on global scales. At a distance of 840\\,kpc \\citep{freedman1991}, M33 is the only nearby, gas rich disk galaxy that allows a coherent survey at high spatial resolution. It does not suffer from any distance ambiguity, as studies of the Milky Way do, and it is not as inclined as the Andromeda galaxy. M33 is a regular, relatively unperturbed disk galaxy, as opposed to the nearer Magellanic Clouds, which are highly disturbed irregular dwarf galaxies. M33 is among the best studied galaxies; it has been observed extensively at radio, millimeter, far-infrared (FIR), optical, and X-ray wavelengths, ensuring a readily accessible multi-wavelength database. These data trace the various phases of the interstellar medium (ISM), the hot and diffuse, the warm and atomic, as well as the cold, dense, star forming phases, in addition to the stellar component. However, submillimeter and far-infrared data at high angular and spectral resolutions have been missing so far. In the framework of the open time key project ``\\herschel\\, M33 extended survey ({\\tt HERM33ES})'', we use all three instruments onboard the ESA \\herschel\\, Space Observatory \\citep{pilbratt2010} to study the dusty and gaseous ISM in M33. One focus of {\\tt HERM33ES} is on maps of the FIR continuum observed with PACS \\citep{poglitsch2010} and SPIRE \\citep{griffin2010}, covering the entire galaxy. A second focus lies on observing diagnostic FIR and submillimeter cooling lines \\CII, \\OI, \\NII, and H$_2$O, toward a $2'\\times40'$ strip along the major axis with PACS and HIFI \\citep{degraauw2010}. \\begin{figure}[t] \\centering \\includegraphics[width=7cm]{14613fg1.eps} \\caption{A composite 500\\,$\\mu$m (red) and 160$\\,\\mu$m (blue) map of M33. The most extended emission is traced by the longest wavelength map revealing the presence of the cold dust in the outskirts of the galaxy. The shorter wavelengths mostly trace the branched spiral structure as well as distinct warm \\HII\\ regions and star-forming complexes (such as the ones labeled).} \\label{fig-maps} \\end{figure} In this first {\\tt HERM33ES} paper, we use continuum maps covering the full extent of M33, at 100, 160, 250, 350, and $500\\,\\mu$m. These data are an improvement over previous data sets of M33, obtained with ISO and \\spitzer\\, \\citep{hippelein2003,hinz2004,tabatabaei2007-466}, in terms of wavelength coverage and angular resolution. The total bolometric luminosity of normal galaxies is only about a factor of 2 larger than the total IR continuum emission \\citep{hauser-dwek2001}, which in turn accounts for more than $\\sim98$\\% of the emission of the ISM (dust$+$gas) \\citep[e.g.][]{malhotra2001,dale2001}. Massive star formation heats the dust mainly via its far-ultraviolet (FUV) photons and the absorbed energy is then reradiated in the IR. FIR continuum fluxes are therefore often used as a measure of the interstellar radiation field (ISRF) \\citep[e.g.][]{kramer2005} and the star formation rate (SFR) \\citep[e.g.][]{schuster2007}. However, a number of authors have suggested that half of the FIR emission or more is due to dust heated by a diffuse ISRF, and not directly linked to massive star formation \\citep{israel1996, verley2009}. Another disputed topic is the evidence for a massive, cold dust component in galaxies. The SCUBA Local Universe Galaxy Survey \\citep{dunne2001} identified a cold dust component at an average temperature of 21\\,K. A number of studies of the millimeter continuum emission of galaxies found indications for even lower temperatures \\citep{misiriotis2006,weiss2008,liu2010}. In order to estimate the amount of dust at temperatures below about 20\\,K, and to improve our understanding of the physical conditions of the big grains, well calibrated observations longward of $\\sim150\\,\\mu$m wavelength are needed. ", "conclusions": "" }, "1005/1005.5343_arXiv.txt": { "abstract": "The multiplicity distributions produced by the variation of time-dependent gravitational fields in a conformally flat background geometry belong to the same class of infinitely divisible distributions found, for fixed centre of mass energies and symmetric (pseudo)rapidity intervals, in charged multiplicities produced in $pp$, $p\\overline{p}$ and in heavy ion collisions. Apparently unrelated multiplicity distributions are classified in terms of the (positive) discrete representations of the $SU(1,1)$ group. The gravitational analogy suggest a global high-energy asymptote for the distributions measured in $pp$ and $p\\overline{p}$ collisions. Second-order cross correlations between positively and negatively charged distributions represent a relevant diagnostic for a closer scrutiny of the multiparticle final state. ", "introduction": " ", "conclusions": "" }, "1005/1005.3279_arXiv.txt": { "abstract": "{We present far-infrared and submillimetre spectra of three carbon-rich evolved objects, AFGL 2688, AFGL 618 and NGC 7027. The spectra were obtained with the SPIRE Fourier transform spectrometer on board the {\\it Herschel} Space Observatory, and cover wavelengths from 195-670$\\mu$m, a region of the electromagnetic spectrum hitherto difficult to study in detail. The far infrared spectra of these objects are rich and complex, and we measure over 150 lines in each object. Lines due to 18 different species are detected. We determine physical conditions from observations of the rotational lines of several molecules, and present initial large velocity gradient models for AFGL 618. We detect water in AFGL~2688 for the first time, and confirm its presence in AFGL~618 in both ortho and para forms. In addition, we report the detection of the J=1-0 line of CH$^+$ in NGC~7027.} {}{}{}{} ", "introduction": "Low- to intermediate-mass stars ($<$8M$_{\\odot}$) shed much of their mass during the final stages of their evolution, in the form of a slow molecular wind. The mass loss causes the star to leave the Asymptotic Giant Branch (AGB), moving to the left in the Hertzsprung-Russell diagram. The star becomes hotter, warming and eventually ionising its previously ejected outer layers, which become visible as a planetary nebula (PN). The ejecta from such stars are a major contributor to galactic chemical evolution (e.g. Matsuura et al. 2009). The evolution from the AGB to the PN stage is rapid, and relatively few objects in this intermediate stage are known (326 are listed by Szczerba et al. (2007), compared to around 3000 known or suspected PNe (Frew \\& Parker 2010)). AFGL~618 and AFGL~2688 (the Egg Nebula) are two of the best-known objects in this transition stage. AFGL~2688 is illuminated by a central star with spectral type F5, and Hubble Space Telescope images reveal a large number of round arcs crossed by `searchlight beams' (Sahai et al. 1998). Proper motion measurements give a distance of 420pc and a dynamical age for the ejected material of 350 years (Ueta et al. 2006). AFGL~618 entered the protoplanetary nebula (PPN) phase 100-200 years ago (Kwok \\& Bignell 1984, Bujarrabal et al. 1988), and is more evolved than AFGL~2688. It contains a B0 central star surrounded by a compact H~{\\sc ii} region (Wynn-Williams 1977, Kwok \\& Bignell 1984). The variability of free-free emission from this object over the last 30 years implies rapid evolution (Kwok \\& Feldman 1981, S\\'anchez Contreras et al. 2002). NGC~7027 is a very young planetary nebula (Masson 1989). It has been extensively studied due to its high surface brightness at all wavelengths, and observations with the Infrared Space Observatory (ISO) revealed a far-infrared spectrum rich in atomic and molecular lines (Liu et al. 1996). Its ionised inner regions are surrounded by and partly obscured by a PDR and massive molecular envelope. Previous far infrared and sub-mm observations of these evolving post-AGB objects, although limited in their spatial resolution and spectral coverage, have revealed rich spectra containing numerous molecular and dust features (e.g. Cox et al., 1996, Liu et al. 1996, Herpin et al. 2002, Pardo et al. 2004, 2005, 2007a, 2007b). The {\\it Herschel} Space Observatory (Pilbratt et al. 2010) now significantly extends our observational capabilities. The SPIRE spectrometer covers wavelengths from 195-670$\\mu$m, a region of the electromagnetic spectrum hitherto largely unexplored. The SPIRE instrument, its in-orbit performance, and its scientific capabilities are described by Griffin et al. (2010), and the SPIRE astronomical calibration methods and accuracy are outlined by Swinyard et al. (2010). Here we present observations of AFGL~2688, AFGL~618 and NGC~7027 using the SPIRE spectrometer. ", "conclusions": "The unprecedented wavelength coverage and excellent sensitivity of the SPIRE spectrometer have allowed us to see the incredibly rich far-infrared and submm spectra of three carbon-rich evolved stars for the first time. We have detected 150-200 emission lines in the SPIRE FTS spectra of AFGL~2688, AFGL~618 and NGC~7027, and presented preliminary investigations of the physical conditions in these objects based on the new data. We find water in both ortho and para forms in all three objects, and observe strong ionised lines in the more evolved NGC 7027, including the J=1-0 line of CH$^{+}$. This line, never previously detected in an astronomical source, is also detected in {\\it Herschel} spectra of H~{\\sc ii} region PDRs (Naylor et al. 2010). LVG models of AFGL~618 are in excellent agreement with those based on earlier ISO data. These observations show that the SPIRE spectrometer on board {\\it Herschel} can provide significant new insights into the temperatures structure and chemical content of the outflows from evolved stars." }, "1005/1005.4459_arXiv.txt": { "abstract": "{Classical T Tauri stars (CTTS) differ in their X-ray signatures from older pre-main sequence stars, e.g. weak-lined TTS (WTTS). CTTS show a soft excess and deviations from the low-density coronal limit in the He-like triplets.}{We test whether these features correlate with accretion or the presence of a disk by observing IM~Lup, a disk-bearing object apparently in transition between CTTS and WTTS without obvious accretion.}{We analyse a \\emph{Chandra} grating spectrum and additional \\emph{XMM-Newton} data of IM~Lup and accompanying optical spectra, some of them taken simultaneously to the X-ray observations. We fit the X-ray emission lines and decompose the H$\\alpha$ emission line in different components.}{In X-rays IM~Lup has a bright and hot active corona, where elements of low first-ionisation potential are depleted. The He-like \\ion{Ne}{ix} triplet is in the low-density state, but due to the small number of counts a high-density scenario cannot be excluded on the 90\\% confidence level. In all X-ray properties IM~Lup resembles a main-sequence star, but it is also compatible with CTTS signatures on the 90\\% confidence level, thus we cannot decide if the soft excess and deviations from the low-density coronal limit in the He-like triplets in CTTS require accretion or only the presence of a disk. IM~Lup is chromospherically active, which explains most of the emission in H$\\alpha$. Despite its low equivalent width, the complexity of the H$\\alpha$ line profile is reminiscent of CTTS. We present an estimate for the mass accretion rate of $10^{-11}M_{\\sun}$~yr$^{-1}$. }{} ", "introduction": "\\label{introduction} Our view on the formation of stars and planetary systems and their emergence from molecular clouds has made significant progress over the last decades. Stars form when giant molecular clouds fragment and contract to form proto-stars. Mass accretion onto those stellar cores proceeds via an accretion disk, while the surrounding envelope eventually disperses. The low-mass pre-main sequence stars in this stage come in two types, the classical T~Tauri stars (CTTS) and the weak-lined T~Tauri stars (WTTS). Traditionally, they were distinguished only by their H$\\alpha$ emission line equivalent width, with those stars of EW $> 10$~\\AA{} defined to be CTTS. It turned out that the EW of H$\\alpha$ is a good tracer of the accretion flow from the circum-stellar disk the CTTS still possess. The H$\\alpha$ EW of WTTS is smaller and their line profiles are symmetric; in contrast CTTS exhibit broader emission lines, which are sometimes asymmetric \\citep{2009A&A...504..461F}. Usually the combination of strong H$\\alpha$ emission and line asymmetry is a reliable accretion indicator \\citep{1998ApJ...492..743M,2003ApJ...592..266M}. WTTS with their low H$\\alpha$ EW are generally expected not to show accretion. It is suggestive to interpret the WTTS as more evolved CTTS, where accretion has stopped already. This does not necessarily imply the absence of a disk, as transitional disks may exist without ongoing accretion for some time \\citep{2006ApJ...645.1283P}. For many TTS the H$\\alpha$ EW is known to be variable, possibly because the accretion switches on and off or --at least-- the accretion rate changes. Both types of TTS have been known for a long time to be copious X-ray emitters \\citep{1999ARA&A..37..363F}. As a class CTTS stand out from other X-ray sources by their strong soft X-ray excess \\citep{RULup,manuelnh}, but CTTS also exhibit normal coronal activity and stellar flares. The CTTS TW~Hya, observed with \\emph{Chandra}/HETGS and \\emph{XMM-Newton} \\citep{2002ApJ...567..434K,twhya}, was the first star where unusual line ratios in the He-like triplets was found indicating high densities in the formation region. This phenomenon repeats in most CTTS observed so far (\\object{BP Tau}: \\citet{bptau}; \\object{V4046 Sge}: \\citet{v4046}; \\object{RU Lup}: \\citet{RULup}; \\object{MP Mus}: \\citet{2007A&A...465L...5A}; \\object{Hen 3-600}: \\citet{2007ApJ...671..592H}). There are exceptions to this rule -- in the more massive, eponymous T~Tau itself, although known for its high accretion rate, the \\ion{O}{vii} triplet is consistent with the coronal limit \\citep{ttau}. The same has been found for HAeBe stars, which are in a similar evolutionary state as the CTTS, yet of higher mass (\\object{AB Aur}: \\citet{ABAur}; \\object{HD 163296}: \\citet{HD163296}). The soft excess and the He-like triplet line ratios in CTTS can be naturally linked to mass accretion from proto-planetary disks \\citep{lamzin,acc_model}. Systems containing typical WTTS such as \\object{TWA 4} \\citep{2004ApJ...605L..49K} and \\object{TWA 5} \\citep{twa5} show no sign of active accretion and no high densities in their X-ray spectra. Well studied examples of young main-sequence (MS) stars such as AU~Mic, Speedy~Mic and AB~Dor \\citep{2004A&A...427..667N} also show triplets indicating low densities similar to older MS objects. Still, 5 of 83 WTTS observed by \\citet{2006ApJ...645.1283P} do show an IR-excess caused by dust within a few AU from the star, detected by \\emph{Spitzer}, thus there is a matter reservoir for accretion. Observationally, we could not distinguish in the past if the X-ray signatures observed in CTTS in contrast to WTTS and young MS stars are caused by active accretion or the presence of a disk. To solve this question we observed a system with \\emph{Chandra}, that was previously classified as disk-bearing, non-accreting WTTS. Simultaneous to the X-ray observations, the accretion state is observed using optical spectroscopy. The properties of the our target IM~Lup are summarised in Sect.~\\ref{imlupprop}. We then present the X-ray observations of IM~Lup and the accompanying optical data in Sect.~\\ref{observations}. We show our new results and compare IM~Lup to CTTS, WTTS and MS stars in Sect.~\\ref{results}. In Sect.~\\ref{discussion} we discuss the implications and end with a short summary in Sect.~\\ref{summary}. ", "conclusions": "\\label{discussion} \\subsection{X-ray properties} IM~Lup shows many characteristics of an active star. It has a bright, hot corona and its abundances follow the IFIP pattern. These traits are common to both CTTS and WTTS. In the density diagnostic of the He-like triplets there is no deviation from the coronal limit, however, due to the low signal a high-density state is allowed at the 90\\% confidence level. The signal of the \\ion{O}{vii} triplet is too weak to check for any soft X-ray excess. Thus, we conclude that -- from the X-ray point-of-view -- it is more likely that IM~Lup shares its characteristics with WTTS, despite the presence of its disk. If confirmed, this suggests that the distinctive characteristics of CTTS are really due to accretion: An excess of soft X-ray emission originates in the high-density environment of an accretion spot as simulated by \\citet{lamzin} and \\citet{acc_model}. Our observation is incompatible with any hypothetical scenario where the high-density signatures of the He-like triplets are attributed to the accretion disk itself because they are not observed (on the 1$\\sigma$ confidence level) in IM Lup, which has a disk. The only peculiarity in comparison to the WTTS samples of COUP and XEST is the high bolometric luminosity of IM~Lup. It belongs to the brightest objects of its kind. This may be due to an overestimation of the distance. We use $190\\pm27$~pc from \\citet{1998MNRAS.301L..39W} based on the \\emph{HIPPARCOS} parallax, but other methods lead to lower values of $140\\pm20$~pc \\citep{1993AJ....105..571H}. Both values are compatible within the $2\\sigma$ errors, but if the lower value was the true one, X-ray emission measure and bolometric luminosity would be only half of the values given above. This does not change any of the other conclusions of our analysis, but IM~Lup would fit in better with the bulk of the COUP and XEST objects in this case. Still, IM~Lup is an active star with a bright and hot corona and $\\log (L_X/L_{bol})=-3.1$, similar to saturated MS stars and also typical for CTTS \\citep{2009A&ARv..17..309G}. % \\subsection{Mass accretion?} \\label{massacretion} IM~Lup has a high level of chromospheric activity, which causes the emission observed in H$\\alpha$. Other M dwarfs with active chromosphere also show H$\\alpha$ emission, typically with an equivalent width of a few~\\AA{} \\citep{1981ApJS...46..159W, 1990ApJS...74..891R}, very similar to IM~Lup. Their Ca~H and K lines often show a very narrow emission peak, too. The H$\\alpha$ line profiles vary, but in most active stars there is a single, narrow emission peak, sometimes with a reversal due to self-absorption in the centre. This is in line with the central peak in the IM~Lup data. In the samples of \\citet{1981ApJS...46..159W} and \\citet{1990ApJS...74..891R} there is not a single case of H$\\alpha$ in emission, where simultaneously broad absorption is present. The same is true for WTTS, which have narrow emission lines of symmetric shape \\citep{2009A&A...504..461F}. In contrast, red-shifted absorption dips and red- and blue-shifted emission components appear in accretion models in most viewing geometries \\citep{2001ApJ...550..944M}. Thus, in IM~Lup the H$\\alpha$ line profile must be caused by additional material which is not present in MS stars: Candidates are a circumstellar envelope, the disk or accretion funnels. We do not detect veiling of the photospheric emission lines, thus any mass accretion must be very low, if present at all. Still, the narrow emission components might originate in a weak accretion funnel. They are shifted by $\\pm100$~km~s$^{-1}$, because we see accretion towards both, the observers side of the star and the back side. The line shift is smaller than usually seen in emission lines from accretion funnels of CTTS, but given the uncertain stellar parameters of IM~Lup this could be due to a combination of low mass, i.e. smaller gravitational potential, and viewing geometry, where the accretion funnel is not parallel to the line-of-sight. % In some stars the H$\\alpha$ line profile is also influenced by strong flares but this cannot be the case in IM~Lup, because no flare shows up in the X-ray lightcurve. The study of \\citet{2003ApJ...582.1109W} correlated the width of the H$\\alpha$ line at 10~\\% of the maximum with accretion. They suggest a full width at 10~\\% of the maximum $>270$~km~s$^{-1}$ as borderline between accreting and non-accreting sources, a slightly lower limit of $>200$~km~s$^{-1}$ was presented by \\citet{2003ApJ...592..282J}. The line width of IM~Lup falls just between those two limits; it would be classified as non-accreting by the one and as accreting by the other boundary. \\citet{2004A&A...424..603N} showed that a quantitative relation between line width and mass accretion rate exists, which is valid from CTTS to sub-stellar objects. This relation yields an accretion rate of $10^{-11} M_{\\sun}\\;\\mathrm{yr}^{-1}$, but IM~Lup is chromospherically very active, which explains a large fraction of the H$\\alpha$ flux. Additionally, not all relations employed to convert the H$\\alpha$ flux to a mass accretion rate agree, if accretion is present at all, therefore we treat this number as a rough estimate only. In summary, the H$\\alpha$ EW classifies IM~Lup as a WTTS, but the H$\\alpha$ line profile is complex as in CTTS. This suggests that mass accretion occurs at a very small rate. The energy released from accretion is too low to influence the X-ray properties significantly. For a mass accretion rate of $10^{-11} M_{\\sun}\\;\\mathrm{yr}^{-1}$, at most 0.5\\% of the X-rays can be due to accretion, assuming that 10\\% of the accretion energy is released in X-rays \\citep{acc_model}. For the interpretation of the X-ray data we can thus safely treat IM~Lup as a non-accreting source. \\subsection{Varying absorption?} The HARPS H$\\alpha$ line profiles are presented in Fig.~\\ref{ha_harps}. While the fitting with Gaussians (table~\\ref{tabha}) yields an absorption component reaching over the entire line profile, for non-Gaussian components absorption is only required on the red side where the observed line profile falls below the continuum. The situation is different in the ANU data, which was obtained simultaneous to the \\emph{Chandra} observation. No absorption is visible here on the first day and wide absorbtion wings are present on both sides of the line on the second day. We postulate that matter from the disk on its way onto the star causes the H$\\alpha$ absorption. With IM~Lup's mass and radius (Sect.~\\ref{imlupprop}) we calculate the Keplerian rotation of the disk. At the co-rotation radius of $6\\;R_{*}$ it matches the stellar rotation period from Sect.~\\ref{opticalspectra}; in this region the inner disk should be truncated. The absorption component is strong at 100-400~km~s$^{-1}$, which corresponds to the free-fall velocity from the co-rotation radius. Thus the H$\\alpha$ absorbing gas may be found in the gap between star and disk; however, this scenario can only explain the red-shifted part of the absorption, but it strengthens the argument in Sect.~\\ref{massacretion} that the H$\\alpha$ line width at 10\\% maximum could be interpreted as accretion. We speculate, that the blue-shifted absorption in the ANU data is caused by an outflow. CTTS commonly drive winds or even jets, where the mass loss is of the order of 10\\% of the accretion rate \\citep{1990ApJ...354..687C,2008ApJ...689.1112C}. The blue-shifted absorption appears only on the second day of the ANU observation, when the absoprtion on the red side, which we interpret as accretion, is stronger. We divided the \\emph{Chandra} observation in two parts and fitted the $n_H$ indenpedently to test if the absorption in the X-ray data changes with time. We found no change of the fitted value. Also, the hardness ratio of the X-ray spectra does not change during the observations. This does not rule out variability in $n_H$, but is sets an upper limit. We estimate that an additional absorbing column of $>10^{21}$~cm$^{-2}$ would have been detected. Does the accretion scenario explain why strong H$\\alpha$ absorption coincides with weak X-ray emission? Given the weak limit on variability in $n_H$ and the fact that we have only two datapoints of simultaneous X-ray and optical spectra, the relation between H$\\alpha$ absorption and X-ray luminosity is --at best-- suggestive. Still it is intriguing to interpret the lower X-ray luminosity and the H$\\alpha$ line profiles as absoption by an accretion funnel. This should cause a hardening of the X-ray absorption. This is not observed, but the limit on extra absorption is weak due to the low count rates. If taken seriously, the relation between X-ray luminosity and H$\\alpha$ absorption fits to the fact that CTTS are on average less luminous in X-rays than WTTS \\citep[e.g.][]{2005ApJS..160..401P,2007A&A...468..443T}. \\citet{2009ApJ...699L..35D} suggested that it is not the accreting nature of the CTTS that causes their low luminosity; rather, it is the low luminosity that allows the disk to reach closer to the star, i.e., weaker emitters accrete and become CTTS. \\subsection{IM Lup in context} \\citet{2009A&A...501.1013S} observed variability in optical spectra of \\object{T Cha}. Like IM~Lup T~Cha is also a transition object from CTTS to WTTS without veiling of photospheric lines, i.e. with little or no mass accretion. \\citet{2009A&A...501.1013S} find the H$\\alpha$ profiles to change from emission lines to absorption components over time, simultaneously they observe changes in the reddening and develop a scenario based on inhomogeneous circumstellar extinction similar to our ideas for IM~Lup. They argue this to be caused by the inner parts of the disk, where grain growth takes place and possibly planets form. Only a small number of CTTS have been observed with X-ray grating spectroscopy, most of them show low $f/i$ ratios in the He-like triplets, but \\object{T Tau} does not. Does IM~Lup confirm a CTTS population with low densities like T~Tau or does it behave as WTTS? Although the complex H$\\alpha$ line profile points to mass accretion, IM~Lup differs from T~Tau in other characteristics. The main component of the multiple system T~Tau is much more massive, close to the boundary to the HerbigAe/Be stars and it accretes at much higher rates. Whereas IM~Lup's X-ray emission must be coronal, T~Tau could power a significant part of its luminosity from accretion. We therefore expect IM~Lup to resemble the WTTS and young MS stars in its X-ray properties. Again, only very few WTTS are observed with X-ray gratings, but they all look similar to MS stars. Despite the small number, it is credible that their emission is driven exclusively by a solar-type corona and we find no evidence that the transition object IM~Lup, which still posseses a disk, is different. The transition from a CTTS to a WTTS in X-rays seems to coincide with the decline of accretion as expected from accretion shock models for CTTS. However, the \\ion{Ne}{ix} triplet is the strongest discriminator and the original driver of the reported observation, but at the 90\\% confidence level the CTTS-like high-density scenario cannot be excluded for IM~Lup." }, "1005/1005.4729_arXiv.txt": { "abstract": "We analysed thermonuclear (type-I) X-ray bursts observed from the low-mass X-ray binary 4U~1728$-$34 by {\\it RXTE}, {\\it Chandra}\\/ and {\\it INTEGRAL}. We compared the variation in burst energy and recurrence times as a function of accretion rate with the predictions of a numerical ignition model including a treatment of the heating and cooling in the crust. We found that the measured burst ignition column depths are significantly below the theoretically predicted values, regardless of the assumed thermal structure of the neutron star interior. While it is possible that the accretion rate measured by {\\it Chandra}\\/ is underestimated, due to additional persistent spectral components outside the sensitivity band, the required correction factor is typically 3.6 and as high as 6, which is implausible. Furthermore, such underestimation is even more unlikely for {\\it RXTE}\\/ and {\\it INTEGRAL}, which have much broader bandpasses. Possible explanations for the observed discrepancy include shear-triggered mixing of the accreted helium to larger column depths, resulting in earlier ignition, or the fractional covering of the accreted fuel on the neutron star surface. ", "introduction": "\\label{intro} Thermonuclear (type I) bursts are triggered by unstable nuclear burning of the material accreted onto the neutron star (NS) surface in low-mass X-ray binary (LMXB) systems \\citep[e.g.,][]{1995xrbi.nasa..175L,2006csxs.book..113S}. The basic theory of type I bursts was outlined shortly after their detection \\citep[e.g.,][]{1976Natur.263..101W,1977Natur.270..310J,1977xbco.conf..127M,1978ApJ...220..291L}. According to these models, the accreted material, which usually consists of hydrogen and helium, accumulates as a thin layer on the NS surface (typically on time-scales of hours to days), and when the pressure and temperature at its base reach critical values, the fuel will ignite and burn unstably until exhausted. Further observations \\citep[see][for a review]{2008ApJS..179..360G} and subsequent modeling \\citep{1981ApJ...247..267F,1987ApJ...319..902F,1987ApJ...323L..55F,1998mfns.conf..419B,2003ApJ...599..419N,2004ApJS..151...75W,2006ApJ...652..584C} showed in detail how the burst properties depend on the accretion rate, composition of the accreted material (the H/He fraction and CNO metallicity), and internal properties of the neutron star. For example, for systems accreting mixed H/He, the heat generated from hydrogen burning is usually a dominant factor for ignition. On the other hand, in evolved systems in which little or no H is present but the fuel consist mainly of He, the heat required for the burst ignition must come entirely from the electron captures and pycnonuclear reactions in the NS crust. Comparisons of He-bursts with ignition models thus offer a powerful probe of the physical conditions in the neutron star crust, below the fuel layer, and the cooling processes in the core \\citep{1987ApJ...319..902F,2006ApJ...646..429C}. The best-known He-accretor is a low-mass binary system 4U~1820$-$30, in which the neutron star orbits its companion once every 11.4 minutes \\citep{1987ApJ...312L..17S}. Such a tiny orbit cannot accommodate a H-rich companion, and the mass donor is likely to be a He-rich white dwarf \\citep{1986Natur.323..105K}. 4U~1820$-$30 is in a bursting mode for around 40 days after switching to the low state \\citep{2001ApJ...563..934C}, while during the rest of its $\\approx$176-day accretion cycle \\citep{1984ApJ...284L..17P} the source does not exhibit bursts \\citep[][and references therein]{1984ApJ...282..713S}. In a 20-hour EXOSAT observation during the low state, \\citet{1987ApJ...314..266H} observed nearly regular bursting from 4U~1820$-$30, detecting seven bursts with a mean recurrence time of $3.21\\pm0.04$ hours and persistent luminosity of $L_X=2.8\\times 10^{37}$ ergs s$^{-1}$ between the bursts. The burst ignition conditions for this source were modelled by \\citet{2003ApJ...595.1077C}, who compared the predicted burst properties with the measurements of \\citet{1987ApJ...314..266H}. \\citet{2003ApJ...595.1077C} presented models for pure He fuel, but also estimated the effect of adding a small amount of hydrogen -- 5$-$35\\% by mass -- as predicted by some stellar evolutionary models \\citep[e.g.,][]{2002ApJ...565.1107P}. The amount of energy released in pycnonuclear and electron capture reactions that escapes from the surface ($Q_{\\rm crust}$) is a free parameter in this model, which also assumes the time-averaged rather than instantaneous accretion rate to set the crust temperature profile, because the thermal time in the crust is much longer than the $\\approx$176-day accretion cycle. \\citet{2003ApJ...595.1077C} found a good agreement between the model, which assumes a mixed fuel (10\\% of hydrogen), and the data, providing that $Q_{\\rm crust}=0.1$ MeV/nucleon \\citep{2000ApJ...531..988B} and the time-averaged accretion rate is $\\approx$2 times larger than the measured rate. However, for a fuel consisting entirely of He, the required $Q_{\\rm crust}$ was 0.4 MeV/nucleon and the required time-averaged accretion rate was 4--5 times larger. Self-consistent results were obtained with improved burst ignition models, in which the flux flowing outwards was calculated directly from the neutron star crust and core neutrino emissivity and the core thermal conductivity \\citep{2006ApJ...646..429C}. \\citet{2006ApJ...646..429C} concluded that, in order to produce the bursts separated by $\\approx$3 hours, the core neutrino emissivity must be very inefficient (e.g., suppresed modified Urca process) and the accretion rate must be $\\approx$2 times larger than that inferred from the X-ray luminosity. Although all previous studies of He-bursts have focused on 4U~1820$-$30, the intermittent occurrence of the bursts makes triggering of burst observations extremely difficult, and only a few recurrence times and corresponding accretion rate estimates are available. A much more suitable candidate for such studies is the source 4U~1728$-$34, which consistently exhibits frequent bursts characteristic of pure He fuel. There is a total of 106 bursts from this source in the {\\sl RXTE} burst catalogue \\citep{2008ApJS..179..360G}. The $\\alpha$ values ($\\approx$200), short rise times and decay time scales suggest a He-rich fuel. The persistent flux during {\\sl RXTE} observations was $1-7 \\times 10^{-9}$ erg cm$^{-2}$ s$^{-1}$ while the burst recurrence times were on average $\\approx$4 hours. Evidence for a short orbital period of 10.77~min has been detected recently in the analysis of {\\sl Chandra} observations (Galloway et al.~2010; in prep.), supporting the long-suspected identification of 4U~1728$-$34 as an ultracompact LMXB. The source is probably accreting pure He from its evolved companion and is practically a twin of 4U~1820$-$30, except for the much more frequent and reliable bursting. 4U~1728$-$34 has also been observed extensively by {\\sl INTEGRAL} \\citep{2006A&A...458...21F,2006AstL...32..456C}. \\citet{2006A&A...458...21F} detected 36 type I bursts during the transition from hard to soft state, where the source luminosity increased from 2$-$12\\% of the Eddington luminosity. In this paper we present analysis of new {\\sl Chandra} observations of 4U~1728$-$34. We measured the recurrence times and corresponding accretion rates of the 25 bursts detected during a 240-ks HETGS exposure. The detailed spectral analysis and the detection of the orbital period and radius expansion bursts are reported in the companion paper (Galloway et al.~2010), while we present the comparison of the observed burst properties with a new ignition model. In Section~2 we present the data and describe our analysis of the {\\sl Chandra} observations. The measured burst properties, which include the accretion rates (estimated from the persistent flux), burst fluences, $\\alpha$ values and burst recurrence times are presented in Section~3. The new burst ignition model is described in Section~4, while Section~5 shows the comparison of the model and data. Possible explanations of the discrepancy between the data and our ignition model are discussed in Section~6. Finally, our conclusions are summarized in Section~7. ", "conclusions": "\\label{conclusions} We compared the observed properties of 38 bursts detected in {\\sl Chandra}, {\\sl RXTE} and {\\sl INTEGRAL} observations of the helium-rich accretor 4U~1728$-$34 with new ignition models. We find that the observationaly-inferred ignition depths, assuming complete fuel spreading on the neutron star surface, are significantly smaller than the theoretically-derived minimum possible ignition depth of $1-2\\times 10^8$ g cm$^2$. One way to reconcile the observed and predicted burst recurrence times would be to assume that the observed X-ray luminosities underestimate the accretion rates (for example, due to a non-detection of the extremely soft or/and hard spectral components). However, for 4U\\,1728$-$34 this scenario is not plausable because these additional spectral components would have to be 2$-$6 times larger than the mesured bolometric luminosity. In addition, it would imply significantly larger $\\alpha$ values than theoretically predicted. Alternatively, the ignition column could be increased, without increasing the global accretion rate, if we assume that the accreted material is confined to a resticted area on the neutron star surface. To match the observations with our ignition model, we find that the spot radii would have to be in the range 4.5$-$6 km, and that they seem to be correlated with the global accretion rate as predicted by the \\citet{2009ApJ...706..417L} model. However, detailed ignition models that include all possible effects of the magnetic confinement are needed to confirm this result. An additional confirmation could come from the comparison of the X-ray emission from the accretion disk with the nuclear oscillation amplitudes detected simultaneously from the source. To explain the weak, frequent bursts observed from 4U~1728$-$34, we also consider shear-triggered mixing of the accreted helium to larger column depths \\citep{1987ApJ...319..902F,2007ApJ...663.1252P}. However, while this mechanism could push the fuel buffer from the outer envelope closer to the burning zone and explain the observed large $\\alpha$ deviations, it might not be sufficient to explain the observed burst properties alone at the relatively low observed accretion rates. {\\it We would like to thank Ed Brown, Maurizio Falanga and Stratos Boutloukos for helpful discussions. We would also like to thank the anonymous referee for the useful and constructive comments. }" }, "1005/1005.3938_arXiv.txt": { "abstract": "{We show that the extremely low-mass white dwarf NLTT~11748 ($0.17\\,M_\\odot$) is in a close binary with a fainter companion. We obtained a series of radial velocity measurements of the low-mass white dwarf using the H$\\alpha$ core and determined an orbital period of $5.64$ hours. The velocity semi-amplitude ($K=274.8$ km s$^{-1}$) and orbital period imply that it is a degenerate star, and that the minimum mass for the companion is $0.75\\ M_\\odot$ (assuming a mass of $0.167\\,M_\\odot$ for the primary). Our analysis of Balmer line profiles shows that a $0.75\\ M_\\odot$ white dwarf companion does not contribute more than 2\\% or 5\\% of the flux (V-band) for helium- or hydrogen-rich surfaces, respectively. The kinematics of the system suggest that it belongs to the Galactic halo. } ", "introduction": "The high proper-motion star NLTT~11748 was identified as an extremely low mass (ELM) white dwarf by \\citet{kaw2009}. The first ELM white dwarfs were discovered to be companions to neutron stars \\citep[e.g.,][]{van1996,bas2003}, but recent follow-up observations of ELM white dwarfs identified in colourimetric or proper motion surveys (e.g., SDSS, NLTT) often find them to be in a close binary with a white dwarf companion \\citep[e.g., SDSS~J125733.63$+$542850.5,][]{bad2009,kul2010,mar2010}. The companions of several other ELM white dwarfs found to be in close binary systems are yet to be formally identified, but they are likely to be more massive white dwarfs \\citep[see,][]{agu2009a,agu2009b}. The stellar and kinematical properties of NLTT~11748 \\citep{kaw2009} and of two similar objects, LP400-22 \\citep{kil2009,ven2009} and SDSS~J1053+5200 \\citep{kil2010}, also suggest that they are old halo stars. The constraints placed on their total ages help retrace the prior evolution of these systems \\citep[see][]{tau1999,nel2001, nel2004}. The formation of ELM white dwarfs requires that the systems go through at least two phases of mass transfer, where the second mass transfer phase strips the less massive companion of its outer envelope before the helium ignition. This generic scenario appears valid whether the more massive component is a neutron star or a white dwarf \\citep[see for example][]{nel2001,nel2004}. An example of a possible progenitor to an ELM white dwarf is the bright subluminous B (sdB) star, HD~188112 \\citep{heb2003}. The subdwarf star has a lower mass than average, $0.24$ versus $\\sim0.5\\,M_\\odot$ \\citep{zha2009}, and orbits a degenerate companion with a mass $\\ga 0.73\\,M_\\odot$ every 14.6 hours. The subdwarf is a likely progenitor of a helium white dwarf similar to NLTT~11748. On the other hand, a possible progenitor to an ELM white dwarf with, this time, a neutron star companion is SDSSJ102347.6+003841 \\citep{wan2009}. This is a close binary comprising a G-type star and a 1.69ms pulsar \\citep{arc2009} with an orbital period of 4.75 hours. Although the spectral type of the visible component implies a mass of $\\approx1\\,M_\\odot$, the mass ratio of the system dictates a much lower mass of $\\sim 0.2\\ M_\\odot$ suggesting that the star is significantly evolved with a helium-enriched core. This system probably represents a first link between low-mass X-ray binaries and ELM plus neutron star binaries \\citep{arc2009}. We report new observations that help clarify the nature of the ELM white dwarf NLTT~11748. We obtained two sets of intermediate resolution spectra that show that NLTT~11748 is in a close binary system (Sect. 2). Our analysis of the spectroscopic data and a first determination of the binary parameters are presented in Sect. 3. We discuss our results in Sect. 4 and summarise in Sect. 5. ", "conclusions": "We find that the ELM white dwarf NLTT~11748 is part of a short-period binary system ($P=5.64$ hrs). Based on our orbital analysis alone, we find that the minimum mass of the companion is $0.75\\,M_\\odot$. However, the companion has independently been identified as a faint white dwarf \\citep{ste2010}, and assuming a mass of $0.75 M_\\odot$, we constrained the companion temperature to $\\le 9600$ K (DA) or $\\le 7200$ K (DC). The kinematics of NLTT~11748 also place the system in the Galactic halo. \\vspace{-0.2cm}" }, "1005/1005.1536_arXiv.txt": { "abstract": "The multifractal nature of solar photospheric magnetic structures are studied using the 2D wavelet transform modulus maxima (\\WTMM{}) method. This relies on computing partition functions from the wavelet transform skeleton defined by the \\WTMM{} method. This skeleton provides an adaptive space-scale partition of the fractal distribution under study, from which one can extract the multifractal singularity spectrum. We describe the implementation of a multiscale image processing segmentation procedure based on the partitioning of the \\WT{} skeleton which allows the disentangling of the information concerning the multifractal properties of active regions from the surrounding quiet-Sun field. The quiet Sun exhibits a average H\\\"older exponent $\\sim -0.75$, with observed multifractal properties due to the supergranular structure. On the other hand, active region multifractal spectra exhibit an average H\\\"older exponent $\\sim 0.38$ similar to those found when studying experimental data from turbulent flows. ", "introduction": "\\label{Introduction} Since the late 70's and the propagation of fractal ideas throughout the scientific community~\\citep{bMan82}, there have been numerous applications of the concepts of scale invariance, self-similarity, long-range dependence in many areas of physics, chemistry, biology, geology, meteorology, economy, social and material sciences~\\citep{b89b,bWes90,b94,bFractal94,b95,b95bis,bFri95,bArn95,bDis02}. Various methods were developed to quantify scale-invariance properties through the computation of the fractal dimension $D_F$ for self-similar objects or the roughness exponent $H$ for self-affine fractals~\\citep{bMan82,b87,bFed88,aArg90,aLea93,b95,aTaq95}. Unfortunately $D_F$ and $H$ are global quantities that do not account for the possibility of point-to-point fluctuations of the scaling properties of a fractal object. The multifractal formalism was introduced in the mid-eighties to provide a statistical description of the fluctuations of regularity of singular measures that are found in chaotic dynamical systems~\\citep{aHas86,aCol87,aRan89} or in modelling of the energy cascading process in turbulent flows~\\citep{aMan74b,aPal87,bMan89,aMen91}. Box-counting and correlation algorithms were successfully adapted to resolve multifractal scaling for isotropic self-similar fractals by computation of the generalized fractal dimensions $D_q$~\\citep{aGras83a,aGras83b,aGra88}. As to self-affine fractals, Parisi and Frisch~\\citep{aPar85} proposed, for the analysis of fully-developed turbulence velocity data, an alternative multifractal description based on the investigation of the scaling behavior of the so-called structure functions~\\citep{bFri95,bMon75}: $S_p(l)=<(\\delta v_l)^p>\\sim l^{\\zeta_p}$ ($p$ integer $> 0$), where $\\delta v_l(x)=v(x+l)-v(x) \\sim l^{h(x)}$ is an increment of the recorded signal over a distance $l$. Then, after reinterpreting the roughness exponent as a local quantity~\\citep{aPar85,aMuz91,aMuz94,aArn95}: $\\delta v_l(x) \\sim l^{h(x)}$ (power-law behavior), the $D(h)$ {\\em singularity spectrum} is defined as the Hausdorff dimension of the set of points $x$ where the local roughness (or H\\\"older) exponent $h(x)$ of $v$ is $h$. In principle, $D(h)$ can be attained by Legendre transforming the structure function scaling exponents $\\zeta_p$~\\citep{aPar85,aMuz91,aMuz94,aArn95}. Unfortunately, as noticed by \\citep{aMuz91,aMuz93,aMuz94}, both the box-counting and structure function methodology have intrinsic limitations and fail to fully characterize the corresponding singularity spectrum since only the strongest singularities are a priori amenable to these techniques. As such, both methods are limited in their application to real data sets \\citep{aMuz94,Georgoulis:2005,Conlon:2008}. In previous work, Arneodo and collaborators~\\citep{aMuz91,aMuz93,aMuz94,aArn95} have shown that there exists a natural way of performing a unified multifractal analysis of both singular measures and multi-affine functions, which consists in using the \\textit{continuous wavelet transform}~\\citep{aGou84,aGros84,bMey90,bDau92,bMal98}. By using wavelets instead of boxes, one can take advantage of the freedom of the choice of these ``generalized oscillating boxes'' to get rid of possible smooth behavior that might either mask singularities or perturb the estimation of their strength $h$. The other fundamental advantage of using wavelets is that the skeleton defined by the \\textit{wavelet transform modulus maxima} (\\WTMM{})~\\citep{aMalZho92,aMalHwa92}, provides an adaptative space-scale partitioning from which one can extract the $D(h)$ singularity spectrum \\textit{via} the scaling exponent $\\tau(q)$ of some partition functions defined from the \\WT{} skeleton. The so-called \\WTMM{} method~\\citep{aMuz91,aMuz93,aMuz94,aArn95} therefore gives access to the entire $D(h)$ spectrum \\textit{via} the usual Legendre transform $D(h)=min_q(qh-\\tau(q))$. We refer the reader to \\citet{aBac93} and \\citet{aJaf97b} for rigorous mathematical results and to \\citet{aHen94} for the theoretical treatment of random multifractal functions. Applications of the \\WTMM{} method to \\UD{} signals have already provided insight into a wide variety of problems~\\citep{refaarn02}, e.g. the validation of the log-normal cascade phenomenology of fully developed turbulence~\\citep{aArn98,aArn99,aDelArn01} and of high-resolution temporal rainfall~\\citep{aVenug06,aVenug06b,aRoux09}, the characterization and the understanding of long-range correlations in DNA sequences~\\citep{aArn95b,aArn96b,aAud01,aAud02}, the demonstration of the existence of a causal cascade of information from large to small scales in financial time series~\\citep{aArn98e,aMuz00}, the use of the multifractal formalism to discriminate between healthy and sick heartbeat dynamics~\\citep{aIva96,aIva99}, the discovery of a Fibonacci structural ordering in \\UD{} cuts of diffusion limited aggregates~\\citep{aArn92,aArn92d,aArn92c,aKuh94} and in hard X-ray emission from solar flares~\\citep{McAteer2007-apj}. The \\WTMM{} method has been generalized from \\UD{} to \\DD{} with the specific goal to achieve multifractal analysis of rough surfaces~\\footnote{The fractal dimension of a rough surface associated to the graph $z=S(x,y)$, where $S(x,y)$ represents the height of the surface at location $(x,y)$, is a quantity $D_F$ between 2 and 3.} with fractal dimension $D_F$ anywhere between 2 and 3~\\citep{aArr97,aArnDec00a,aArnDec00b}. The \\DD{} \\WTMM{} method has been successfully applied to characterize the intermittent nature of cloud structure from satellite images~\\citep{aArn99c,aArnDec00c} and to assist in the diagnosis of breast tissue lesions in digitized mammograms~\\citep{akes01}. In astrophysics, this method was adapted and used to characterize the anisotropic structure of atomic hydrogen gas (HI) in the Galatic disk~\\citep{kha06}. From the analysis of very large mosaics taken from the Canadian Galatic Plane Survey~\\citep{tay03}, directional roughness exponents were introduced to show that the HI in the Galactic spiral arms has a scale-dependent anisotropic signature while the HI in the inter-spiral arm regions exhibits scale-independent anisotropy. Along that line, the \\DD{} \\WTMM{} method was further applied to characterize the space-scale nature of anisotropic structures~\\citep{aSnow08a,aSnow08b,akha09} and to perform objective segmentation of image features of interest from noisy backgrounds~\\citep{kha07,acad07,athi09}. We refer the reader to \\citet{bArn02} for a review of the \\DD{} \\WTMM{} methodology, from the theoretical concepts to experimental applications. Recently, the \\WTMM{} method has been further extended to \\TD{} scalar~\\citep{aKes03_prl} as well as \\TD{} vector~\\citep{akes04prl,aSerra07} fields analysis and applied to \\TD{} (velocity, vorticity, dissipation, enstrophy) numerical data issued from direct numerical simulations (DNS) of incompressible Navier-Stokes equations. Because it combines singular-value decomposition and multifractal description, the so-called tensorial wavelet transform modulus maxima method for vector fields~\\citep{akes04prl,aSerra07} looks very promising for future simultaneous multifractal and structural (vorticity sheets, vorticity filaments) analysis of turbulent flows. Our aim here is to exploit the ability of the \\WTMM{} method to study compound systems that display some non-analyticity in their multifractal spectra as the signature of some phase transition between two underlying scale invariant components with different multifractal properties~\\citep{bBoh88,aMuz94,aArn95}. These two components can both have some physical significance as previously experienced when using the \\WTMM{} method to detect vorticity filaments in swirling turbulent flows~\\citep{aRouMuz99} or microcalcifications from breast tissue background in digitized mammograms~\\citep{akes01,bArn02}. One of these components can be noise that may cause drastic distortions in the returned multifractal spectra. In this work we will follow a wavelet-based strategy inspired from the one previously used in \\UD{} to detect replication origins and promoters as jumps (discontinuities) in \\UD{} noisy skew profiles in mammalian genomes~\\citep{aBro05,touchon05,nicolay07} and in \\DD{} to perform an objective and automatic segmentation of chromosome territories in fluorescence microscopy imaging of mouse cell nuclei~\\citep{kha07,acad07} and of gold formation on vapodeposited thin gold films~\\citep{athi09}. The purpose of the manuscript is to demonstrate the suitability and reliability of the \\WTMM{} method to propose a wavelet-based segmentation procedure adapted to solar magnetogram data. In section~\\ref{segmentation}, the basics of the \\DD{} \\WTMM{} method are presented. Its ability to disentangle the underlying scale invariant components of a compound system displaying a phase transition in its singularity spectra is discussed and a strategy of segmentation is implemented. Section~\\ref{section_synthetic_data} is devoted to a test application of the proposed segmentation procedure on a theoretical data set with known multifractal properties. In section~\\ref{observations} we report the results obtained when using this wavelet-based segmentation method to separate active regions from quiet-Sun features in solar line-of-sight magnetogram data. Our conclusion and future directions are then given in Section~\\ref{conclusions}. ", "conclusions": "\\label{conclusions} Many complex physical systems analyzed using fractal and multifractal techniques are surrounded or embedded in a noisy background sometimes originating from instrumental noise. As such these systems are a statistical combination of two distinct self-similar structures. This work addressed the need for an accurate calculation of the multifractal parameters of such complex systems. The presence of compound scale-invariant structures can result in an inaccurate or skewed calculation of the fractal and multifractal parameters when studied as a whole. Using a wavelet-based multi-scale segmentation method, we show that it is possible to disentangle to some extent these two processes and accurately (up to finite-size effects) recover the multifractal characteristics of the system of interest. A theoretical test example for this method was provided in section \\ref{segmentation}. The removal of information relating to the background noise was highlighted in Figure~\\ref{fig_skeleton}. The quantitative results reported in Figure~\\ref{fig_pf} attest of the ability of this segmentation method to recover the multifractal parameters in question. Let us emphasize here that the application of this method to experimental data for which we do not have a priori knowledge of the possible underlying multifractal processes is a difficult task that requires much attention to perform the most objective segmentation which can not be infered or guided by some physical rule or information. The multifractal analysis is a statistical tool that has direct connections with signal and image processing, but not necessarily with the physics of the system per se. As noticed in section \\ref{sec:sol_mag_active}, the use of a clustering algorithm should greatly help in adjusting the multifractal parameters of the differents components as well as in providing an automated procedure for processing large data sets. This will be reported in a future publication. The application of this wavelet-based methodology to quiet-Sun data has revealed the multifractal nature of this intermittent noisy component ($=-0.75$) as mainly resulting from the super-granular magnetic structures. The quiet-Sun study was also necessary to get expertise for further analyzing more complex images that involve a segmentation before being able to clearly identify the underlying multifractal properties. We have checked that the partition functions computations for the segmented quiet-Sun phase provide (i) convincing scaling properties and (ii) multifractal spectra $\\tau(q)$ and $D(h)$ estimates in good numerical agreement with the one measured in the previous calibration step. The assumption of two non-overlapping $D(h)$ is not inconsistent with the data. Let us notice that a phase transition can be observed in the partition function log-log plots when inappropriate segmentation parameters are chosen. In the case of overlapping $D(h)$, there would be a large set of maxima lines in the WT skeleton that could not be genuinely sorted, which would prevent from building accurate multifractal $D(h)$ measurement. From the analyzed data, we were not able to distinguish more than two phases. Finally, this gives us good confidence in the segmentation proposed for solar magnetogram containing an active region. However, further study is needed to precisely quantify scaling properties associated to specific active region features (\\textit{e.g.} emerging magnetic flux along the main polarity inversion line, sunspots build-up, delta-configuration...) and how the \\WTMM{} method can be sensitive to these elements. More precisely, when analyzing higher resolution images than MDI, we expect that this segmentation tool will be all the more necessary as quiet-Sun and active region features are more entangled. The main outcome of the present study is the demonstration that the proposed multi-scale segmentation procedure provides an objective way of studying the complexity in active regions separately from the surrounding quiet Sun. As such our results are significantly more stable and robust when compared to previous fractal and multifractal analysis~\\citep{Pontieri:2003,McAteer2005-apj}. In a forthcoming paper, we will report on the application of this segmentation method to characterize the evolution of active regions keeping track of the multifractal parameters for possible correlations with extreme solar events~\\citep{Gallagher:2007}. Other applications of this method are in progress as the analysis of the intrinsic multifractal properties of entangled hot and cold interstellar atomic gas from \\TD{} numerical simulations~\\citep{kes_audit:2009}." }, "1005/1005.4914_arXiv.txt": { "abstract": "We present simulations of scattering phenomena which are important in pulsar observations, but which are analytically intractable. The simulation code, which has also been used for solar wind and atmospheric scattering problems, is available from the authors. These simulations reveal an unexpectedly important role of dispersion in combination with refraction. We demonstrate the effect of analyzing observations which are shorter than the refractive scale. We examine time-of-arrival fluctuations in detail: showing their correlation with intensity and dispersion measure; providing a heuristic model from which one can estimate their contribution to pulsar timing observations; and showing that much of the effect can be corrected making use of measured intensity and dispersion. Finally, we analyze observations of the millisecond pulsar J0437$-$4715, made with the Parkes radio telescope, that show timing fluctuations which are correlated with intensity. We demonstrate that these timing fluctuations can be corrected, but we find that they are much larger than would be expected from scattering in a homogeneous turbulent plasma with isotropic density fluctuations. We do not have an explanation for these timing fluctuations. ", "introduction": "\\label{sec:intro} Observations of radio pulsars have shown the effects of the interstellar plasma since pulsars were discovered. The first and most obvious effect is dispersion due to the column density of free electrons between the pulsar and the observer. However the effects of scattering due to fluctuations in the electron density were soon recognized as they are very strong in pulsars (Rickett, 1969; Rankin et al., 1970; Cordes, 1986). It is now known that scattering by fluctuations in electron density due to Kolmogorov turbulence are diffractive in nature but the diffractive process is strongly modulated by large scale refraction. As a result there are two spatial scales of the intensity fluctuations, a small scale now called the ``diffractive scale'' $s_\\text{dif}$ and a larger one called the ``refractive scale'' $s_\\text{ref}$ (Prokhorov et al., 1975). The early observations showed only the diffractive scale. The refractive scale was discovered by Sieber (1982) and explained by Rickett et al. (1984). Furthermore the column density changes slowly and this is also observable (Rawley et al., 1988; Ramachandran et al., 2006; You et al., 2007). These fluctuations in ``dispersion measure'' (the pulsar observer's term for column density) have their origin in the same interstellar turbulence and they merge with the diffractive and refractive effects. Observations and theory of interstellar scattering have been reviewed by Rickett (1990) and Narayan (1992). Scattering is an inherently spatial effect, but it is observed as time variation in, for example, the pulse intensity. The temporal variation is simply caused by spatial variation convected past the observer by the relative motion of the pulsar, the Earth, and the ionized interstellar medium (IISM). Thus we relate an observed time scale $\\tau$ to the corresponding spatial scale S using an effective velocity $S = V_\\text{eff} \\tau$ (Cordes and Rickett, 1998). As there are two spatial scales $s_\\text{dif}$ and $s_\\text{ref}$, there are two temporal scales $\\tau_\\text{dif}$ and $\\tau_\\text{ref}$. In practice $\\tau_\\text{dif}$ is usually minutes to hours so the diffractive component is well sampled in a typical observation. However $\\tau_\\text{ref}$ is usually days to months, so refractive variations are seldom seen in a single observation period. This is why they were not discovered for nearly two decades. Consequently observations are usually analyzed neglecting the refractive variation, but the derived parameters of the diffractive effects are seen to vary with observing epoch (Gupta et al., 1994; Gupta et al., 1999). The question then arises, ``Is this variation part of a continuous variation due to homogeneous turbulence in the IISM, or must we invoke an inhomogeneity in the electron density as was used by Fiedler et al. (1987) to explain extreme scattering events (ESE)?'' ", "conclusions": "Simulations have been shown to be a very useful tool in understanding observations of scattering in the interstellar plasma. The simulation engine has been regularized to the point where it can be used by others and it is available from the authors. Here we show a number of examples where simulations have allowed us to understand long standing peculiarities in the observations. We also show that the simulated timing noise due to homogeneous isotropic turbulence is considerably smaller than two cases of observed timing noise which is correlated with intensity fluctuations. This opens a number of interesting questions about the effects of the assumptions of homogeneous isotropic turbulence, and about the observations themselves." }, "1005/1005.3643_arXiv.txt": { "abstract": "Solar flares presumably have an impact on the deepest layers of the solar atmosphere and yet the observational evidence for such an impact is scarce. Using ten years of measurements of the Na D$_{1}$ and Na D$_2$ Fraunhofer lines, measured by GOLF onboard SOHO, we show that this photospheric line is indeed affected by flares. The effect of individual flares is hidden by solar oscillations, but a statistical analysis based on conditional averaging reveals a clear signature. Although GOLF can only probe one single wavelength at a time, we show that both wings of the Na line can nevertheless be compared. The varying line asymmetry can be interpreted as an upward plasma motion from the lower solar atmosphere during the peak of the flare, followed by a downward motion. ", "introduction": "\\label{sec:intro} Solar flares are the most powerful events occurring in the solar system. The mechanism invoked for their energy release is the conversion of magnetic energy into radiation, thermal and kinetic energy through the reconnection of magnetic field lines. More precisely, the coronal magnetic field suddenly changes its configuration and free energy when the stress imposed at its footpoints by photospheric plasma motion exceeds a certain threshold \\citep{2005ApJ...635.1299A}. The details of why and how the energy is released are still largely debated (\\textit{e.g} \\citep{2009AdSpR..43..739S}). Once converted, the kinetic energy accelerates particles; part of these are directed downwards and deposit their energy in the solar atmosphere while another part can eventually escape into interplanetary space. The interaction of these accelerated particles with the surrounding plasma ultimately results in electromagnetic radiation at various wavelengths, from X-ray to radio domains \\citep{2008LRSP....5....1B}. One of the best observed manifestations of flares is the appearance, after the impulsive phase, of loops that emit strongly in the extreme ultraviolet and soft X-rays (SXR). Downward-accelerated particles lose their energy by collisions when they encounter the dense chromospheric plasma. The local material is heated, expands, and then rises to fill the newly configured magnetic-field loop where plasma cools down through radiative and conductive losses. This is the model of chromospheric evaporation \\citep{1984ApJ...287..917A,1987ApJ...317..502F}: the hot plasma located just above the energy-deposition layer evaporates towards the corona. The cool plasma just underneath is pushed down; this is called chromospheric condensation. In this model, the energy deposition layer is located in the middle chromosphere or in the transition region. The rise of chromospheric plasma is manisfested by the observation of Doppler-shifted spectral lines that form at these temperatures \\citep{1987ApJ...317..956S,2005A&A...430..679B} and by the downflow of plasma as observed from the H$\\alpha$ line \\citep{1987SoPh..113..307F}.\\\\ White-light (WL) continuum emission has been also observed in some flares -- that are then called white light flares (WLF) -- but their origin is unclear \\citep{2006SoPh..234...79H}. It is generally agreed that the WL emission takes place in the lower atmosphere, near the minimum temperature region or below \\citep{1986A&A...156...73A,2003A&A...403.1151D}. How these layers are heated, however, remains unknown (see the review by \\citep{2007ASPC..368..417D}). Among the various proposed mechanisms, direct heating by non-thermal electron beams, chromospheric radiative back-warming, or H$^{-}$ continuum emission \\citep{2006ApJ...641.1210X,2000A&A...354..691G} are often cited. When taken individually, none of these mechanisms can provide the amount of energy required by the observations. Direct heating in or near the photosphere has also been suggested \\citep{1994ApJ...429..890D,2009AdSpR..43..739S}. The latter mechanisms suggest that the low solar atmosphere (the chromosphere of course, but also the temperature minimun region and the upper photosphere below) is heated during these flares. As technology evolves, more WL emission is observed during flares \\citep{2008ApJ...688L.119J,Kretzschmar_sub}. We present here observations in visible light made by the \\textit{Global Oscillations at Low Frequency} (GOLF:~Gabriel \\textit{et al}, 1995) instrument onboard SOHO (space mission of international cooperation between ESA and NASA) since 1996 \\citep{1997SoPh..175..207G}. These observations give information on the low solar atmosphere during solar flares.\\\\ One question that we address in this paper is the recurrent problem of retrieving a global picture of the flare since, in addition to the intrinsic variability of the events, the conditions (position, amplitude, ...) vary and the observations are often made by different instruments. We follow here the same approach as in Kretzschmar \\textit{et al}, (2010) and search for a statistical signature of flares by analysing ten years of SOHO/GOLF data consisting in the intensity of the Fraunhofer absorption Na D lines integrated over the full solar disk. These lines are formed in the solar photosphere below the temperature minimum region \\citep{1980A&A....92...63C,1992A&A...265..237B}. We refer here to the photosphere as a few hundred kilometer thick region between the altitude where the optical depth at 500 nm is unity ($ \\tau_{500} = 1 $) and the temperature minimum region. More precisely, the wings of the Na D line that we study in this paper are mostly generated between 300 and 500 km, just below the transition region and the chromosphere \\citep{2001A&A...371.1128E}. Simulations and observations for a quiet Sun atmosphere have been recently made for the formation of the Na D lines \\citep{2010ApJ...709.1362L}: this study strongly contributes to the idea that these lines are originating from the photosphere. However during a flare, the formation heights of the Na D sodium may differ from quiet conditions. Nevertheless changes in the Na D lines during flares thus indicate that flares affect the lower solar atmosphere including both photosphere and chromosphere. Our analysis therefore provides an opportunity to study the influence of solar flares on the lower solar atmosphere. This is particularly interesting since the best observed flare features usually occur in the upper solar atmosphere, where the contrast between the flare signal and the background emission is largest.\\\\ The intensity of the Sodium line has been measured every five seconds by GOLF. The good cadence, duration and duty cycle of the observations, together with the high quality associated with space-based observations, partly allow us to compensate for the lack of spatial resolution. As we shall see below, large and even medium amplitude flares do have a significant impact on the Na line, which we interpret as the signature of photospheric heating. Furthermore, this work also points to the possible influence of flares on helioseismic studies. \\\\ In Section 2 we introduce the GOLF instrument, the data, and the analysis technique. In Section 3, the results are presented and discussed. Conclusions follow in Section 4. ", "conclusions": "The aim of this study was to investigate whether solar flares impact the Na D$_1$ and Na D$_2$ Fraunhofer lines, as measured by GOLF. A statistical analysis based on conditional averaging reveals intensity increases during X and M-class flares, which demonstrates that the Na D lines are affected. Although GOLF observes only one wing at a time, the average evolution of the red and blue wings can be compared in a statistical sense. The observed intensity differences suggest the existence of plasma motions. The blueshift that coincides with the peak of the flare can been interpreted as a upward plasma motion, which is most likely associated with photospheric Na rising towards the chromosphere. The peak of the flare is followed by a redshift, which corresponds to a downward flow. Since the Na D line is mostly generated in the upper photosphere, our study thereby provides strong support for the influence of flares on the lower solar atmosphere. This interpretation is \\textit{a priori} in contradiction with the classical picture of chromospheric evaporation. It can however be explained if we assume that white-light flares, for which the lower solar atmosphere needs to be heated, have a preponderant effect in our statistics. White-light flares could indeed be more common than expected, and our study supports this idea. However, further quantitative studies are necessary before we can conclude that solar flares do really affect the lower solar atmosphere. Plasma motions could be more accurately determined with more than two measurements on the profile of Na D lines. The GOLF-NG has been proposed for that purpose. Another important conclusion is that future helioseismic instruments should take into account the impact of solar flares in order to improve the measurement of solar oscillations.\\\\ \\textbf" }, "1005/1005.4549_arXiv.txt": { "abstract": "{\\small The new release of data from Wilkinson Microwave Anisotropy Probe improves the observational status of relic gravitational waves. The 7-year results enhance the indications of relic gravitational waves in the existing data and change to the better the prospects of confident detection of relic gravitational waves by the currently operating Planck satellite. We apply to WMAP7 data the same methods of analysis that we used earlier [W. Zhao, D. Baskaran, and L.P. Grishchuk, Phys. Rev. D {\\bf 80}, 083005 (2009)] with WMAP5 data. We also revised by the same methods our previous analysis of WMAP3 data. It follows from the examination of consecutive WMAP data releases that the maximum likelihood value of the quadrupole ratio $R$, which characterizes the amount of relic gravitational waves, increases up to $R=0.264$, and the interval separating this value from the point $R=0$ (the hypothesis of no gravitational waves) increases up to a $2\\sigma$ level. The primordial spectra of density perturbations and gravitational waves remain blue in the relevant interval of wavelengths, but the spectral indices increase up to $n_s =1.111$ and $n_t=0.111$. Assuming that the maximum likelihood estimates of the perturbation parameters that we found from WMAP7 data are the true values of the parameters, we find that the signal-to-noise ratio $S/N$ for the detection of relic gravitational waves by the Planck experiment increases up to $S/N=4.04$, even under pessimistic assumptions with regard to residual foreground contamination and instrumental noises. We comment on theoretical frameworks that, in the case of success, will be accepted or decisively rejected by the Planck observations.} ", "introduction": "} The Wilkinson Microwave Anisotropy Probe (WMAP) Collaboration has released the results of 7-year (WMAP7) observations \\cite{wmap7, wmap7-larson}. In this paper, we apply to WMAP7 data the same methods of analysis that we have used before \\cite{stable} in the analysis of WMAP5 data. This is important for updating the present observational status of relic gravitational waves and for making more accurate forecasts for the currently operating Planck mission \\cite{planck}. In Sec.~\\ref{section2.0} we briefly summarize our basic theoretical foundations and working tools. Part of this material was present in the previous publication \\cite{stable}, but in order to make the paper self-contained we briefly repeat it here. Section \\ref{section2} exposes full details of our maximum likelihood analysis of WMAP7 data. In the focus of attention is the interval of multipoles $2 \\leq \\ell \\leq 100$, where gravitational waves compete with density perturbations. We compare all the results that we derived by exactly the same method from WMAP7, WMAP5 and WMAP3 datasets. This comparison demonstrates the stability of data and data analysis. On the grounds of this comparison, one can say that the perturbation parameters found from the consecutive WMAP data releases have the tendency of saturating near some particular values. The WMAP7 maximum likelihood (ML) value of the quadrupole ratio $R$ is close to previous evaluations of $R$, but increases up to $R=0.264$. The interval separating this ML value from the point $R=0$ (the hypothesis of no gravitational waves) increases up to a $2\\sigma$ level. The primordial spectra remain blue, but the spectral indices in the relevant interval of wavelengths increase up to $n_s = 1.111$ and $n_t=0.111$. In Sec.~\\ref{section3} we analyze why, to what extent, and in what sense our conclusions with respect to relic gravitational waves differ from those reached by the WMAP Collaboration. The WMAP team has found ``no evidence for tensor modes.\" A particularly important issue, which we discuss in some detail, is the presumed constancy (or simple running) of spectral indices. We derive an exact formula for the spectral index $n_t$ as a function of wavenumbers and discuss in this context the formula for running that was used in WMAP analysis. Another contributing factor to the difference of conclusions is the difference in our treatments of the inflationary ``consistency relations\" based on the inflationary ``classic result.\" We do not use the inflationary theory. A comprehensive forecast for Planck findings in the area of relic gravitational waves is presented in Sec.~\\ref{section4}. We discuss the efficiency of various information channels, i.e. various correlation functions and their combinations. We perform multipole decomposition of the calculated $S/N$ and discuss physical implications of the detection in various intervals of multipole moments. We stress again that the $B$-mode detection provides the most of $S/N$ only in the conditions of very deep cleaning of foregrounds and relatively small values of $R$. The improvements arising from a 28-month, instead of a nominal 14-month, Planck survey are also discussed. In the center of our attention is the model with the WMAP7 maximum likelihood set of parameters. For this model, the signal-to-noise ratio $S/N$ in the detection of relic gravitational waves by Planck experiment increases up to $S/N=4.04$, even under pessimistic assumptions with regard to residual foreground contamination and instrumental noises. Section \\ref{bayescomp} gives Bayesian comparison of different theoretical frameworks and identifies predictions of $R$ that may be decisively rejected by the Planck observations. ", "conclusions": "} Having analyzed the 7-year data release, the WMAP collaboration concludes that a minimal cosmological model without gravitational waves and with a constant spectral index $n_s$ across the entire interval of considered wavelengths remains a ``remarkably good fit\" to ever improving CMB data and other datasets. The WMAP team emphasizes: ``We do not detect gravitational waves from inflation with 7-year WMAP data, however the upper limits are $16\\%$ lower...\" \\cite{wmap7-larson}, p.11; ``The 7-year WMAP data combined with BAO and ${\\rm H_0}$ excludes the scale-invariant spectrum by more than $3\\sigma$, if we ignore tensor modes (gravitational waves)\" \\cite{wmap7}, p.15; ``We find no evidence for tensor modes...\" \\cite{wmap7}, p.30; ``We find no convincing deviations from the minimal model\" \\cite{wmap7}, p.1, etc. In contrast, our analysis of WMAP3, WMAP5 and WMAP7 data leads us in the opposite direction: the improving data make the gw indications stronger. The major points of tension between the two approaches seem to be the constancy of spectral indices and the continuing use by the WMAP team of the inflationary theory in data analysis and interpretation. We shall start from the discussion of spectral indices. The constancy of spectral indices is a reasonable assumption, but not a rule. If the power-law dependence (\\ref{inscf}) is not a good approximation to the gravitational pump field during some interval of time, the constancy of $n_t, ~n_s$ is not a good approximation to the generated primordial spectra (\\ref{primsp}) in the corresponding interval of wavelengths. In fact, the future measurements of frequency dependence of the spectrum of relic gravitational waves will probably be the best way to infer the ``early history of the Hubble parameter\" \\cite{grsol}. The frequency-dependence of a gw spectrum is fully determined by the time-dependence of the function $\\gamma(t) \\equiv - {\\dot H}/{H^2}$. (In more recent papers of other authors this function is often denoted $\\epsilon(t)$.) The function $\\gamma(t)$ describes the rate of change of the time-dependent Hubble radius $l_{H}(t) \\equiv c/H(t)$: \\[ \\gamma(t) = \\frac{d}{dt}\\left(\\frac{1}{H(t)}\\right) = \\frac{1}{c} \\frac{dl_{H}(t)}{dt}. \\] The function $\\gamma(t)$ is a constant for power-law scale factors (\\ref{inscf}): $\\gamma = (2+\\beta)/(1+\\beta)$, and $\\gamma=0$ for a period of de Sitter expansion. The interval of time $dt$ during the early era when gravitational waves were entering the amplifying regime and their today's frequency spread $d\\nu$ are related by (see Eq.(21) in \\cite{grsol}): \\[ \\frac{d}{dt} =[1-\\gamma(\\nu)]H(\\nu)\\frac{d}{d \\ln \\nu}. \\] The today's spectral energy density of gravitational waves $\\epsilon(\\nu)$ is related to the early universe parameter $\\gamma(t)$ by (see Eq.(22) in \\cite{grsol}): \\begin{eqnarray} \\label{22} \\gamma(\\nu) = -\\frac{[d~\\ln\\epsilon(\\nu)/d~\\ln\\nu]} {2-[d~\\ln\\epsilon(\\nu)/d~\\ln\\nu]}. \\end{eqnarray} The spectral index $n_g$ of a pure power-law energy density $\\epsilon(\\nu) \\propto \\nu^{n_g}$ is defined as $n_g = [d~\\ln\\epsilon(\\nu)/d~\\ln\\nu]$. It is reasonable to retain this definition for more complicated spectra. Then, Eq.(\\ref{22}) can be rewritten as \\begin{eqnarray} \\label{ngnu} n_{g}(\\nu) = -\\frac{2 \\gamma(\\nu)}{1- \\gamma(\\nu)}. \\end{eqnarray} Obviously, in the case of pure power-laws (\\ref{inscf}) we return to the constant spectral index $n_g = -2\\gamma/(1-\\gamma) = 2(2+\\beta)$. Equation (\\ref{ngnu}) was derived for the energy density of relatively high-frequency gravitational waves, $\\nu > 10^{-16} {\\rm Hz}$, which started the adiabatic regime during the radiation-dominated era. In our CMB study we deal with significantly lower frequencies. It is more appropriate to speak about wavenumbers $n$ rather than frequencies $\\nu$, and about power spectra of metric perturbations $h^2(n)$ rather than energy density $\\epsilon(\\nu)$. The $k$-dependent spectral index $n_{t}(k)$ entering primordial spectrum (\\ref{PsPt}) is defined as $n_{t}(k) = [d~\\ln P_{t}(k)/d~\\ln k]$. Then formula for $n_{t}(k)$ retains exactly the same appearance as Eq.(\\ref{ngnu}): \\begin{eqnarray} \\label{ntk} n_{t}(k) = - \\frac{2 \\gamma(k)}{1- \\gamma(k)}. \\end{eqnarray} The spectral index $n_{t}(k)$ reduces to a constant $n_t=2(2+\\beta)$ in the case of power-law functions (\\ref{inscf}). The spectral index $n_{s}(k) -1$ for density perturbations is defined by $n_{s}(k)-1 = [d~\\ln P_{s}(k)/d~\\ln k]$. The formula for $n_s(k) - 1$ is more complicated than Eq.(\\ref{ntk}) as it contains also $d\\gamma(t)/dt$ as a function of $k$. However, it is important to stress that adjacent intervals of power-law evolution (\\ref{inscf}) with slightly different constants $\\beta$, will result in slightly different pairs of constant indices $n_t,~n_s-1$ in the corresponding adjacent intervals of wavelengths. Of course, the spectrum itself is continues at the wavelength marking the transition between the two regions. Extending the minimal model, the WMAP Collaboration works with the power spectrum \\cite{wmap7} \\begin{eqnarray} \\label{runPs} P_{s}(k) = P_{s}(k_0) \\left(\\frac{k}{k_0}\\right)^{n_{s}(k_0) -1 + \\frac{1}{2}\\alpha_s \\ln(k/k_0)}, \\end{eqnarray} which means that the $k$-dependent (running) spectral index $n_{s}(k)$ is assumed to be a constant plus a logarithmic correction: \\begin{eqnarray} \\label{runns} n_{s}(k) = n_{s}(k_0) +\\alpha_s \\ln(k/k_0). \\end{eqnarray} The aim of WMAP data analysis is to find $\\alpha_s$, unless it is postulated from the very beginning, as is done in the central (minimal) model, that $\\alpha_s \\equiv 0$. We note that although logarithmic corrections do arise in simple situations and can even be termed ``natural\" \\cite{grsol}, they are not unique or compulsory, as we illustrated by exact formula (\\ref{ntk}). Nevertheless, we do not debate this point. We accept WMAP's definitions, and we want to illustrate their results graphically, together with our evaluations of $n_s$ in this paper. The main result of WMAP7 determination is $n_s=0.963 \\pm 0.012$ (68\\% C.L.) derived under the assumption of no gravitational waves and constant $n_s$ throughout all wavelengths included in the considered datasets \\cite{wmap7,wmap7-larson}. When the presence of gravitational waves is allowed, but $n_s$ is still assumed constant, the $n_s$ rises to $n_s= 0.982^{+0.020}_{-0.019}$ from WMAP7 data alone. Finally, from WMAP7 data alone, the WMAP team finds $n_s(k_0)=1.027,~\\alpha_s= -0.034$ in the case of no gw but with running of $n_s$, and $n_s(k_0) =1.076,~\\alpha_s = -0.048$ in the case of running and allowed gravitational waves (we quote only central values without error bars, see Table 7 in \\cite{wmap7}). All the resulting values of $n_s(k)$ derived by WMAP team are shown in Fig.~\\ref{sec2-fig2}. For comparison, we also plot by red lines our evaluations of $n_s$, see Sec.~\\ref{results}. The lines in Fig.~\\ref{sec2-fig2} show clearly that our finding of a blue shape of the spectrum, i.e. $n_s =1.111$, at longest accessible wavelengths is pretty much in the territory of WMAP findings, if one allows running, even as simple as Eq.(\\ref{runns}), and especially when running is combined with gravitational waves. On the other hand, as was already explained in \\cite{stable}, the attempt of constraining relic gravitational waves by using the data from a huge interval of wavelengths and assuming a constant $n_s$ (or its simple running) across all wavelengths is unwarranted. The high-$\\ell$ CMB data, as well as other datasets at relatively short wavelengths, have nothing to do with relic gravitational waves, and their use is dangerous. As we argued in Sec.~\\ref{results}, the spectral index $n_s$ appears to be sufficiently different even at the span of two adjacent intervals of wavenumbers. The restriction to a relatively small number of multipoles $2 \\leq \\ell \\leq 100$ is accompanied by relatively large uncertainties in $R$, but there is nothing we can do about it to improve the situation, this is in the nature of efforts aimed at measuring $R$. The difference in the treatment of $n_s$ is probably the main reason why we do see indications of relic gravitational waves in the data, whereas the WMAP team does not. Another contributing factor to the difference of conclusions is the continuing use by WMAP Collaboration of the inflationary theory and its (incorrect) relation $n_t = -r/8$, which automatically sends $r$ to zero when $n_t$ approaches zero. This formula is a part of the inflationary `consistency relations' \\[ r= 16 \\epsilon = - 8 n_t. \\] Only one equality in this formula, $16 \\epsilon = -8n_t$, is correct being an approximate version (for small $\\gamma$, $\\epsilon \\equiv \\gamma$) of our exact formula, Eq. (\\ref{ntk}). The `consistency relation' $r= 16 \\epsilon$ is incorrect. It is an immediate consequence of the ``classic result\" of inflationary theory, namely, the prediction of arbitrarily large amplitudes of density perturbations generated in the limit of de Sitter inflation ($\\epsilon =0,~n_s=1$), regardless of the strength of the generating gravitational field (curvature of space-time) regulated by the Hubble parameter $H$ \\footnote{It is difficult to give adequate references to the origin of the ``classic result\". Judging from publications, conference talks and various interviews, there exists harsh competition among inflationists for its authorship. One popular inflationary activity of present days is calculation of small loop corrections to the theory which is wrong by many orders of magnitude in its lowest tree approximation. For a more detailed criticism of inflationary theory, see \\cite{disca}.}. Certainly, it would be inconsistent, even by the standards of inflationary theory, not to use the relation $r= - 8n_t$ in data analysis, if the inflationary ``classics\" is used in derivation of power spectra and interpretation of results (see, for example, Fig.~19 in \\cite{wmap7}). Obviously, in our analysis we do not use the inflationary theory and its relations. \\begin{figure} {\\includegraphics[height=10cm]{a8.eps}} \\caption{The spectral index $n_s$ as a function of the wavenumber $k$. The ML results of this work are shown by red lines. Other lines are our plots of WMAP7 findings \\cite{wmap7}, Table 7.} \\label{sec2-fig2} \\end{figure} The analysis of the WMAP7 data release amplifies observational indications in favor of relic gravitational waves in the Universe. The WMAP3, WMAP5, and WMAP7 temperature and polarization data in the interval of multipoles $2 \\leq \\ell \\leq 100$ persistently point out to one and the same area in the space of perturbation parameters. It includes a considerable amount of gravitational waves expressed in terms of the parameter $R=0.264$, and somewhat blue primordial spectra with indices $n_s = 1.111$ and $n_t =0.111$. If the maximum likelihood set of parameters that we derived from this analysis is a fair representation of the reality, the relic gravitational waves will be detected more confidently by Planck observations. Even under pessimistic assumption about hindering factors, the expected signal-to-noise ratio should be at the level $S/N = 4.04$. ~ {\\bf Acknowledgements} We acknowledge the use of the LAMBDA and CAMB Websites. W. Z. is partially supported by Chinese NSF Grants No. 10703005, and No. 10775119 and the Foundation for University Excellent Young Teacher by the Ministry of Zhejiang Education. \\baselineskip=12truept" }, "1005/1005.4063_arXiv.txt": { "abstract": "We study the peculiar velocities of density peaks in the presence of primordial non-Gaussianity. Rare, high density peaks in the initial density field can be identified with tracers such as galaxies and clusters in the evolved matter distribution. The distribution of relative velocities of peaks is derived in the large-scale limit using two different approaches based on a local biasing scheme. Both approaches agree, and show that halos still stream with the dark matter locally as well as statistically, i.e. they do not acquire a velocity bias. Nonetheless, even a moderate degree of (not necessarily local) non-Gaussianity induces a significant skewness ($\\sim 0.1-0.2$) in the relative velocity distribution, making it a potentially interesting probe of non-Gaussianity on intermediate to large scales. We also study two-point correlations in redshift-space. The well-known Kaiser formula is still a good approximation on large scales, if the Gaussian halo bias is replaced with its (scale-dependent) non-Gaussian generalization. However, there are additional terms not encompassed by this simple formula which become relevant on smaller scales ($k \\gtrsim 0.01\\iMpch$). Depending on the allowed level of non-Gaussianity, these could be of relevance for future large spectroscopic surveys. ", "introduction": "\\label{sec:intro} Observations of the large scale structure in the Universe that use galaxies, clusters or other tracers of the density field are done in redshift space: the distance is generally inferred using the redshift $z$, which receives a contribution from the line-of-sight velocity of the object. These velocities are due to the gravitational field which is correlated with the density field itself. On large scales where linear perturbation theory in the density field applies, the leading contribution is the squashing (or stretching, in case of an underdensity) of a volume element in redshift space relative to real space. In this limit, there is a simple relation between the real- and redshift-space power spectra, $P_g$ and $P_{g,s}$, respectively, of a tracer `$g$' \\cite{Kaiser,Fisher95}: \\be P_{g,s}(k, \\mu) = \\left ( 1 + \\frac{f}{b_1} \\mu^2\\right)^2 P_g(k), \\label{eq:Kaiser} \\ee where $\\mu$ is the cosine of the $\\vk$ vector with the line of sight, $f=d\\ln D/d\\ln a$ is the logarithmic derivative of the linear growth factor, and $b_1$ is the linear bias of the tracer population. Apart from the large-scale, small-correlation limit, the relation \\refeq{Kaiser} makes two assumptions: first, that the tracer population is characterized by a deterministic local bias. In particular, if the tracer density $\\d_g(\\vx)$ is a local function $F(\\d(\\vx))$ of the matter density perturbation $\\d$, we can expand in $\\d(\\vx)$ to obtain \\cite{FryGaztanaga} \\be \\d_g(\\vx) = b_1\\:\\d(\\vx) + \\frac{b_2}{2} \\d^2(\\vx) +\\dots, \\label{eq:local} \\ee where the bias parameters are either to be seen as free empirical parameters, or can be determined using various theoretical approaches. Local biasing amounts to the assumption that dark matter halos form in high-density regions (peaks) in the initial density field. This assumption holds well in the high-peak / massive halo regime, which we assume throughout. Hence, in the following we will somewhat loosely use ``peaks'' and ``halos'' interchangeably. The second assumption used for \\refeq{Kaiser} is that the cosmological density field is Gaussian on large scales. While we will retain the local biasing scheme, we are interested in relaxing the second assumption of Gaussianity. Recently, there has been renewed interest in probing the Gaussianity of the primordial seed perturbations via large scale structure (see \\cite{DesjacquesSeljak10} for a review). The simplest way of obtaining a non-Gaussian field is by adding a \\emph{local} non-linearity: \\be \\Phi(\\vx) = \\Phi_{\\rm G}(\\vx) + \\fNL (\\Phi_{\\rm G}^2(\\vx) - \\<\\Phi_{\\rm G}^2\\>), \\label{eq:philocal} \\ee where $\\Phi_{\\rm G}$ is a Gaussian random field, and $\\Phi$ is the resulting non-Gaussian field. Of course, one can add higher powers to the series \\refeq{philocal}, though the quadratic term usually has the largest impact. Following standard convention, we let $\\Phi$ stand for the primordial potential, related to the density field through the transfer function and Poisson equation (see \\refapp{bispectra}). As shown by \\cite{DalalEtal08} and confirmed by \\cite{MV08,SlosarEtal}, in the presence of non-Gaussian initial conditions of the local type, halos acquire a scale-dependent correction to their bias which becomes important on large scales: \\be b_1 \\rightarrow b_1 + 2\\fNL (b_1-1)\\d_c \\M(k), \\ee where $\\M(k) \\propto k^{-2}$ is the relation between density and potential in Fourier space (see \\refapp{bispectra}). Besides the local model, several other possible bispectrum configurations have been proposed in the literature, such as the \\emph{equilateral} (e.g., \\cite{CreminelliEtal06}) and \\emph{folded} types (e.g., \\cite{MeerburgEtal09}). We summarize these types of non-Gaussianity and the relation between the potential and matter perturbations in \\refapp{bispectra}. Given the significant impact of (local) non-Gaussianity on the power spectrum of biased tracers, it is then natural to ask what happens to the power spectrum in redshift space, and whether \\refeq{Kaiser} still holds. Furthermore, the distribution of relative velocities between tracers at $\\vx_1$ and $\\vx_2$, \\be \\d \\vu = \\vu(\\vx_2)-\\vu(\\vx_1) \\ee is itself an interesting probe of non-Gaussianity \\cite{CatelanScherrer}. In this paper, we assume sub-horizon scales throughout, and adopt the Newtonian gauge; further, we work in Lagrangian coordinates. At first order, the transition to Eulerian coordinates (which all observations as well as simulation measurements are made in) simply amounts to replacing the Lagrangian bias parameters with their Eulerian counterparts, in case of the linear bias simply $b_{E,1} = b_{L,1}+1$. While we will focus on the local type of primordial non-Gaussianity for the most part, the expressions obtained can easily be evaluated for any given primordial bispectrum, and we will show selected results for other bispectrum shapes. The paper is structured as follows: \\refsec{not} introduces density and velocity correlations, and some notation. \\refsec{deriv} contains two different derivations of the moments of the relative peak velocity distribution. We discuss the distribution of peak velocities in \\refsec{Pdu}. Finally, \\refsec{Pks} presents the power spectrum of peaks (halos) in redshift space. We conclude in \\refsec{concl}. ", "conclusions": "\\label{sec:concl} In this paper, we have studied the large-scale motions of peaks in the density field in the presence of primordial non-Gaussianity. In the high-peak regime we are interested in, peaks can be identified with dark matter halos and visible structures such as galaxies and clusters. We have derived the statistics of the relative velocity of halo pairs in two different approaches based on a local biasing scheme. Note that the bias is local in the physical, \\emph{non-Gaussian} density field. The first, simpler and more accessible approach assumes that on large scales, halo velocities follow those of the total matter. Given linear and quadratic bias parameters, it is then straightforward to calculate the moments of the relative velocity distribution. In the second approach (see \\refapp{deriv}), no assumptions are made apart from local biasing and the large-scale, small-correlation limit. Both approaches agree, showing that the assumption of statistically unbiased halo velocities is consistent on large scales even in the presence of non-Gaussianity. Interestingly, the presence of non-zero three-point correlations on large scales leads to significant changes in the velocity distribution of halos, and matter itself. For a positive $\\fNL$, the mean streaming velocity is significantly enhanced on large scales, and a non-zero third moment leads to a positive skewness of the velocity difference distribution. In the local model, the skewness $\\<\\d u^3\\>\\:/\\:\\<\\d u^2\\>^{3/2}$ is of order $(0.1-0.2)\\times(\\fNL/100)$ on a wide range of scales. Even in the equilateral model, which has little effect on the halo power spectrum, the skewness still reaches 0.06 for the same $\\fNL$. This suggests that the velocity difference distribution can serve as an interesting probe of non-Gaussianity (as already pointed out by \\cite{CatelanScherrer}). Furthermore, these findings could be of relevance to recent observational reports of significantly higher velocities than expected in the $\\L$CDM (or more generally, General Relativity + Dark Energy) scenario \\cite{FeldmanEtal,KashlinskyEtal,LeeKomatsu}. In order to evaluate this quantitatively however, one has to take into account non-linear corrections, via perturbation theory and/or simulations. Our second result is an expression for the redshift-space power spectrum of halos (or matter) in the presence of non-Gaussianity. The well-known Kaiser formula relating real- and redshift-space power spectra, extended by the scale-dependent halo bias (see e.g. \\cite{DesjacquesSeljak10}), receives non-Gaussian corrections which become relevant on small scales. Note that through these corrections, the redshift-space power spectrum measured at a given scale as function of line-of-sight angle in principle allows for a direct measurement of non-Gaussianity from a given tracer population, without any reference to the underlying matter power spectrum (in the Kaiser formula at a fixed scale, the non-Gaussian effects are perfectly degenerate with the galaxy bias). Furthermore, the redshift-space power spectrum of matter itself receives corrections of order $\\fNL$ from non-Gaussianity, which are not present in real space. The non-Gaussian corrections to the Kaiser formula lead to the interesting question of whether there are degeneracies of the effects of non-Gaussianity with other parameters measured from redshift-space distortions, such as dark energy parameters or consistency tests of General Relativity \\cite{ZhangEtal,SongPercival09}. The severity of the degeneracies will depend on the level of allowed non-Gaussianity. While the local model is likely to be constrained tightly in the near future, other models such as the equilateral model are much less constrained, but can still be lead to noticeable effects in redshift space (\\reffig{terms-eql}). Finally, we expect that non-Gaussian effects on halo velocities will have an even larger impact on the \\emph{bispectrum} of halos in redshift space. Again, these questions deserve more study via higher order perturbation theory as well as N-body simulations. \\vspace*{-0.5cm}" }, "1005/1005.2190_arXiv.txt": { "abstract": "We present spectral and kinematic evidence that 2MASS~J06085283$-$2753583 (M8.5$\\gamma$) is a member of the $\\beta$ Pictoris Moving Group (BPMG, age~$\\sim$12~Myr), making it the latest-type known member of this young, nearby association. We confirm low-gravity spectral morphology at both medium and high resolutions in the near-infrared. We present new radial velocity and proper motion measurements and use these to calculate galactic location and space motion consistent with other high-probability members of the BPMG. The predicted mass range consistent with the object's effective temperature, surface gravity, spectral type, and age is 15--35~$M_{Jup}$, placing 2MASS~0608$-$27 well within the brown dwarf mass regime. 2MASS~J06085283$-$2753583 is thus confidently added to the short list of very low mass, intermediate age benchmark objects that inform ongoing searches for the lowest-mass members of nearby young associations. ", "introduction": "Low mass members of nearby clusters, moving groups, and loose associations are important objects of study for stellar and planetary astrophysics, formation, and evolution. These groups are older than regions of ongoing star formation (e.g., in Taurus and Orion, ages $\\sim$1--5~Myr) but younger than the disk population (ages $\\sim$1--10~Gyr). Such associations have only recently been discovered because they are low-density groups and their members are spread out on the sky. Currently identified groups include the TW~Hydrae Association (TWA, age $\\sim$10~Myr), the $\\beta$ Pictoris Moving Group (BPMG, $\\sim$12~Myr), and the Tucana/Horologium Association (Tuc-Hor, $\\sim$30~Myr), along with several other clusters/associations with ages younger than $\\sim$500~Myr \\citep[e.g.,][]{Zuckerman04,Torres08}. Young low mass objects are hotter and have larger radii than older objects of the same mass \\citep[e.g.,][]{Chabrier00}, thus they are brighter and can be studied more efficiently than their older counterparts. Young associations and moving groups are closer to the Sun than the closest star-forming regions ($d$$\\sim$120--140~pc), allowing for more detailed study of their lowest-mass members. Furthermore, low mass members of slightly older associations ($>$20~Myr) have typically ceased accretion and dissipated their primordial disks, making it easier to disentangle atmospheric and circumstellar properties. On the other hand, they still may show signs of youth, such as low gravity spectral features and activity indicators. Eventually a complete census of the low mass members of nearby associations will address outstanding questions in star formation, including the low mass end of the initial mass function and its dependence on environment as well as the mass and age dependence of various processes in star and planet formation. This Letter establishes 2MASS~J06085283$-$2753583 (hereafter 2MASS~0608$-$27) as the lowest-mass confirmed member of the BPMG. Section~2 describes previous observations of 2MASS~0608$-$27, and Section~3 presents the observations analyzed herein. The kinematic and spatial properties of 2MASS~0608$-$27 are determined in Section~4, and its membership in the BPMG is discussed in Section~5. We present the conclusions of this Letter in Section~6. ", "conclusions": "Spectral features, kinematics, and the space location of 2MASS~0608$-$27 clearly confirm it as a member of the BPMG. 2MASS~0608$-$27 is the latest-type non-companion member by a substantial margin. Similarly late-type objects in the BPMG are the M7.5 companion HR~7329B, possibly the M6.5/M9 binary 2MASS~J0041353$-$562112, and the candidate planetary-mass companion to $\\beta$~Pictoris. Ongoing proper motion and RV measurements for apparently young M and L dwarfs in the field will likely confirm larger numbers of very low mass members of the BPMG and other young, nearby loose associations (\\citealt{Schlieder10}, K.~L.~Cruz et al. 2010, in preparation). The identification of the M8.5 object 2MASS~0608$-$27 as a member of the BPMG provides an important late-type benchmark object for further studies." }, "1005/1005.0195_arXiv.txt": { "abstract": "A kinetic equation for the collisional evolution of stable, bound, self gravitating and slowly relaxing systems is established, which is valid when the number of constituents is very large. It accounts for the detailed dynamics and self consistent dressing by collective gravitational interaction of the colliding particles, for the system's inhomogeneity and for different constituent's masses. It describes the coupled evolution of collisionally interacting populations, such as stars in a thick disk and the molecular clouds off which they scatter.\\\\ The kinetic equation derives from the BBGKY hierarchy in the limit of weak, but non-vanishing, binary correlations, an approximation which is well justified for large stellar systems. The evolution of the one-body distribution function is described in action angle space. The collective response is calculated using a biorthogonal basis of pairs of density-potential functions.\\\\ The collision operators are expressed in terms of the collective response function allowed by the existing distribution functions at any given time and involve particles in resonant motion. These equations are shown to satisfy an H-theorem. Because of the inhomogeneous character of the system, the relaxation causes the potential as well as the orbits of the particles to secularly evolve. The changing orbits also cause the angle Fourier coefficients of the basis potentials to change with time. We derive the set of equations which describes this coupled evolution of distribution functions, potential and basis Fourier coefficients for spherically symmetric systems. In the homogeneous limit, which sacrifices the description of the evolution of the spatial structure of the system but retains the effect of collective gravitational dressing, the kinetic equation reduces to a form similar to the Balescu-Lenard equation of plasma physics. ", "introduction": "\\label{introduction} The description of collisional relaxation in a self-gravitating system usually rests on a Fokker-Planck equation in which the diffusion and braking coefficients are calculated in the local approximation, taking the finite dimension of the system into account by limiting the impact parameter of the collisions to a length of order of the system's size \\citep{Chandra, Chandra43, BinneyTremaine, SpitzerAmas}. Although characteristic relaxation times may be somewhat overestimated by this approximation due to the neglect of collective self-gravitational effects \\citep{Weinberg93}, such a kinetic equation may provide in practice a reasonable description of the collisional relaxation of gravitationally bound systems. It nevertheless rests on assumptions which, from a principle point of view, are unsatisfactory because the motion of particles during the collision is regarded as rectilinear and uniform and the system's inhomogeneity, which is basically the reason why collisions with an infinite impact parameter do not occur, is treated by way of an ill-defined cutoff. Moreover, the collective response of the system is not taken into account, since the Fokker-Planck collision term only considers binary collisions between naked particles. A self-gravitating medium, unlike an electrical plasma, does not respond to the presence in it of a particle by screening its interaction potential with other particles. As a result, even distant particles effectively interact, while in electrical, globally neutral, plasmas, the effective interaction distance is limited to the Debye length. In a self-gravitating system, the distance between interacting particles is only limited by the system's inhomogeneity. The spatial structure of the system matters as well as the details of the particle orbits. \\bigskip The consistent inclusion of collective screening effects in a kinetic equation for electrically interacting weakly coupled particles has been one of the major theoretical achievements in plasma physics when \\citet{Balescu} and \\citet{Lenard} could derive an equation surpassing in consistency the simple Fokker-Planck equation \\citep{petitSpitzer}. It is the aim of this paper to derive a similar equation for self-gravitating systems. The task is slightly more difficult because the screening of the electrical interaction at the, usually small, Debye length allows, in electrically interacting systems, to take the homogeneous and uniform motion limits. These limits cannot be taken in a self-gravitating system. We overcome this difficulty by expressing the kinetic equation in action angle space rather than in position momentum space. This is possible when the Hamiltonian corresponding to the average potential $U({\\mathbf{r}})$ of the system is integrable. It is nevertheless uneasy in general to toggle from one to the other space, although this is certainly possible for spherically symmetric potentials, for flat systems (which may however be unstable) and for special thick disk potentials. Numerical methods could be used to achieve the necessary transformation \\citep{PichonCannon, McMillanBinney}. As an illustrative example, we shall give special attention to spherically symmetric potentials, expanding their kinetic equation into a system which almost entirely avoids any calculation in the position-momentum space. The system's inhomogeneity requires that solutions to the Poisson equation are easily found for any inhomogeneous mass distributions. This is achieved by projecting on a biorthogonal basis of pairs of density-potential functions. \\bigskip Many astrophysical systems which have evolved to a quasi-stationary collisionless equilibrium still keep evolving on time scales longer than the dynamical time as a result of gravitational noise induced by their own constituents or by external ones. We disregard external perturbators, which we define as unbound to the system, although, as did \\citet{Weinberg2001II}, these could be treated, if numerous and frequent enough, as a given, non-evolving, population providing a source of gravitational noise for other populations. Loosely bound satellites or remote star populations are regarded as internal to the system. This is possible because our set of kinetic equations allows to simultaneously follow different mass populations. Dwarf satellite galaxies could be regarded for example as one such mass population. Globular clusters, dwarf galaxies, disk galaxies and their haloes are examples of bound systems still evolving as a result of internal noise caused by particle discreteness. Such systems are the object of our study. As in any weakly coupled system, the particles suffering collisions are dressed by the polarization clouds caused by their own influence on other particles. Collisions between dressed particles have quantitatively different outcomes than collisions between naked ones \\citep{Weinberg98}. This may reflect in significant differences in calculated effective relaxation times and braking or diffusion coefficients, especially when the system, though stable, is not too far from instability \\citep{Weinberg93}. It is therefore useful to account for collective dressing when calculating such processes as secular thick disk evolution, mass segregation in galaxies or in star clusters, or the damping by dynamical friction of galactic populations on high energy orbits. For simplicity, the kinetic equations to be derived below assume that the system is stationary on a dynamical time scale. They thus cannot address questions in which the distribution in angle variable matters, such as the dissolution of freshly accreted satellites, although a simple extension of the theory could. Since however our equations describe the coupled evolution of all populations present in the system, they are well suited to study, for example, the simultaneous evolution by dynamical friction and diffusion of a stellar population and the population of molecular clouds off which these stars scatter. \\bigskip The collective response of a self gravitating system to the presence of a perturbing body has been considered by a number of authors, analytically \\citep{Weinberg89, Weinberg95, MuraliTremaine, SahaJog} or numerically \\citep{ThielheimWolff, GnedinOstriker}. Sometimes, the reaction of this perturbation on the perturbing body itself is calculated, as did \\citet{Kalnajs72}, who computed the drag on a large body moving in an homogeneous medium, taking the collective response of this medium into account, and \\citet{WeinbergTremaine}, who considered the global, self-consistent, perturbation caused by a satellite or a barred structure in a spherically symmetric system and its reaction on the perturbator object by the effect of dynamical friction. The secular evolution of the system in response to such perturbations has been considered by \\citet{Weinberg2001}, who considered general types of perturbations on a galaxy, and by \\citet{PichonAubert} who considered perturbations caused by the cosmological environment on dark matter haloes. This evolution is of course in principle observable in N-body simulations, which however have their own difficulties in calculating the long term evolution of such systems \\citep{Binney2004}. A number of authors \\citep{Murali, Weinberg2001, PichonAubert} have studied the collective perturbations caused in a massive spherical galactic halo by its environment. They could calculate the response of this system by resorting to a representation of the particle's motion in action and angle variables, a method first used by \\citet{Kalnajs77}. We follow them on this road. They also made good use of a basis of biorthogonal pairs of density-potential functions. \\citet{Weinberg93} first derived a kinetic equation for the collisional relaxation of a self gravitating system along these lines. His equation accounts for the self-consistent gravitational dressing of the particles, but is otherwise simplified, the geometry supposedly being that of an homogeneously filled periodic cube. The inhomogeneous nature of the system should be described more accurately, still accounting for collective gravitational dressing effects. This is specifically the aim of this paper. \\citet{Chavanis} presented a similar approach to ours for one-dimensional systems, the constituents of which interact by a general long range force. In this paper we further elaborate in section \\ref{secsystemcouple} on the structural evolution of the inhomogeneous system and on the secular evolution of the orbits. ", "conclusions": "\\label{secconclu} Kinetic equations for the collisional evolution of the constituents of self-gravitating inhomogeneous systems have been derived. These equations (\\ref{LAequation}) surpass in consistency the usual Fokker-Planck equations (\\ref{formeFP})~--~(\\ref{Asanseffetscoll})~--~(\\ref{Bsanseffetscoll}). The latter are unsatisfactory from a principle point of view, being local and non-collective. \\bigskip By contrast, the proposed equations fully account for the system's inhomogeneity and for the collective gravitational dressing of the colliding particles. \\bigskip Equations (\\ref{LAequation}) describe the evolution of distribution functions in action and angle space, which is possible when the hamiltonian associated with the average potential is integrable. \\bigskip Physically, these equations describe the evolution of the distribution functions in action space as a result of the weak gravitational noise caused by the discreteness of the particles, dressed with the polarization clouds that their own gravity induces in the system. This gravitational polarization is accounted for in equation (\\ref{LAequation}) in a manner that is fully consistent with the distribution functions, as they are at the moment. \\subsection{Properties of the kinetic equation} \\label{subsecproperties} Equation (\\ref{LAequation}) is the sum of a second order derivative term with respect to actions and of a first order one. It therefore basically is of the Fokker-Planck type, although it is definitely simpler in the form of expression (\\ref{LAequation}). The diffusion coefficient involved depends on the 1-body distributions themselves, in particular through the factor $\\mid \\!{\\cal{D}}\\!\\mid^{-2}$ which represents the effect of the dressing of the colliding particles by the gravitational polarization induced around them by their own influence. \\bigskip Unlike in electrical plasmas, the polarization dressing in self-gravitational systems does not cause any screening of the interaction, which remains effective even between distant particles. The mutual distance of such particles is limited only by the finite size of the system. Were the gravitational influence of particles on their surrounding to be neglected, the response matrix $\\varepsilon$ (equation (\\ref{epsilonalphabeta})) would reduce to unity and the coefficients of the corresponding Fokker-Planck kinetic equation would simply be averages by the distribution functions of functions of velocity, as in equations (\\ref{Asanseffetscoll})~--~(\\ref{Bsanseffetscoll}). \\bigskip It is apparent from the developments of appendix \\ref{grossesmagouilles}, which lead to equation (\\ref{LAequation}), that the ${\\mathbf{k}}$ component in angle Fourier space of the gravitational polarization response given to a particle has frequency $\\omega = {\\mathbf{k}}\\! \\cdot \\! {\\mathbf{\\Omega}}$. This means that the polarization cloud which accompanies a particle forms a structure in angle space which vary as ${\\mathbf{w}} - {\\mathbf{\\Omega}}t$: it corotates in angle with that particle. \\bigskip The presence of the Dirac function $\\delta({\\mathbf{k}}_1\\!\\cdot {\\mathbf{\\Omega}}_1 - {\\mathbf{k}}_2\\!\\cdot {\\mathbf{\\Omega}}_2)$ in equation (\\ref{LAequation}) indicates that particles interact resonantly. This certainly is an important physical property of remote interactions, for which the components of the angle wave vectors ${\\mathbf{k}}_1$ and ${\\mathbf{k}}_2$ must be small. For closer encounters, the modulus of these wave vectors is larger and the resonance condition ${\\mathbf{k}}_1\\cdot {\\mathbf{\\Omega}}_1 = {\\mathbf{k}}_2\\cdot {\\mathbf{\\Omega}}_2$ becomes less selective, being more easily satisfied. \\bigskip The correlation function has been calculated on the basis of a linearized theory, which is justified by the weakness of the average interactions in this many-body system. This means that the trajectories of the particles during the collision are regarded as being the unperturbed trajectories. Similarly, the gravitational polarization cloud around any one of the colliding particles is calculated as if the partner in the collision were not present: equation (\\ref{LAequation}) is still a weak collision approximation. A cutoff at small impact parameters is therefore needed to account for the rare strong collisions. \\bigskip Equation (\\ref{LAequation}) takes full account of the inhomogeneity of the system, which is embodied in the dependence of the distribution functions on the actions $\\mathbf{J}$'s. It requires no artificial cutoff at large impact parameters. The details of the trajectories followed by the particles in the present gravitational potential are also fully accounted for, being implicit in the relations which link the angle and action variables to the position and momentum ones. These relations depend on the actual global gravitational potential of the system, which slowly evolves in time together with the distribution functions. \\bigskip The density-potential basis functions $\\psi^{\\alpha}({\\mathbf{r}})$ are choosen at the beginning of the calculation once and for all, but their angle Fourier transforms $\\psi^{\\alpha}_{\\mathbf{k}}({\\mathbf{J}})$, which depend on the actual trajectories of the particles, change with time because the trajectory of a particle of given actions slowly evolves with the general potential of the system as the relaxation proceeds. As long as it suffers no collision, a given particle keeps its vector ${\\mathbf{J}}$ fixed because the actions are adiabatic invariants. Collisions, however, cause a secular evolution of the functions $f^a({\\mathbf{J}})$, which is exactly what equation (\\ref{LAequation}) describes. \\bigskip \\noindent The description of particle motions is made simple by the use of action and angle variables. Their complexity is embodied in the supposedly known relation between position and momentum variables and action and angle variables. The usefulness of equation (\\ref{LAequation}) is therefore limited to systems for which this relation can be established, either analytically or, possibly, numerically \\citep{PichonCannon, McMillanBinney}. \\bigskip \\noindent While the relaxation proceeds, the gravitational potential and the orbits of the particles evolve. As a result, the kinetic equation must be completed by evolution equations for the potential and for other relevant quantities. Section \\ref{secsystemcouple} establishes, for spherical systems, this set of coupled equations. \\subsection*{Acknowledgements} \\label{acknowledgements} I am indebted to C. Pichon and D. Aubert for awaking my interest for this problem. Their work \\citep{PichonAubert} introduced me to the tools which make it possible to extend the methods of plasma physics to these systems. I thank the referee, C. Pichon, for many suggestions which helped to improve the first version of this paper. I also thank the Observatoire Astronomique and the Universit\\'e de Strasbourg for accepting me as professor emeritus." }, "1005/1005.4705_arXiv.txt": { "abstract": "{% Preliminary results on the discovery and follow-up observations of a new $\\delta$ Scuti pulsator in the Cygnus field are presented. The variability of the star HD 207331 was detected while testing a Str\\\"omgren spectrophotometer attached to the H.~L. Johnson 1.5-m telescope at the San Pedro M\\'artir observatory, M\\'exico. CCD photometric data acquired soon after confirmed its variability. A few hours of $uvby$ differential photoelectric photometry during three nights revealed at least two beating periods. A two-site observational campaign carried out during one week in 2009 confirms the multi-periodic nature of this new $\\delta$ Scuti pulsator.} ", "introduction": "The $\\delta$ Scuti-type pulsators are stars with masses between 1.5 and 2.5 $M_{\\odot}$ located at the intersection of the classical Cepheid instability strip with the main sequence. They have spectral types A and F, a period range between 0.5 h to 6 h, and generally pulsate with a large number of radial and nonradial modes excited by the $\\kappa$ mechanism. This makes them interesting targets for seismic studies. Therefore, any new detection of a $\\delta$ Scuti star can be a valuable contribution to asteroseismology. Since most of the $\\delta$ Scuti stars are short period variables with typical photometric amplitude of 20 mmag, their oscillations can be easily detected from the ground. In fact, several $\\delta$ Scuti stars have been discovered accidentally when taken as reference stars of observations of well known $\\delta$ Scuti stars (e.g. Fox Machado et al. 2002 and 2007; Li et al. 2002). Others have been catalogued either in surveys devoted to the characterization of new variables or, as in this case, when considered as constant stars while testing observatory equipment. In particular, this paper presents a summary of the observations which yielded the discovery and characterization of the new $\\delta$ Scuti star, HD 207331. \\begin{figure*} \\includegraphics[width=15cm,height=17cm]{fig1.ps} \\caption{CCD differential light curve: HD 207331 - Comparison. The fit of the four-frequency solution to the data is shown by solid line.} \\label{fig:curves} \\end{figure*} ", "conclusions": "The determination of the evolutionary stage of a field star requires precise estimates of its global parameters. In the case of HD 207331, there is little information in the literature about its physical parameters. In particular, the Hipparcos catalogue (Perryman et al. 1997) provides a parallax of 3.31 $\\pm$ 0.88 mas, from which a distance of 302 pc can be estimated. The large relative error of the measured distance ($\\sigma (\\pi))/ \\pi \\sim 0.27$) implies $\\sigma (M_V)\\sim 0.6$ mag, making the Hipparcos absolute magnitude very imprecise for HD 207331. Therefore, with the information available it is not possible to make a reliable seismic modeling for HD 207331. However, the complicated oscillation spectrum of HD~207\\-331, with two beating modes, points to it being a fast rotating $\\delta$ Scuti star. This would imply that the four frequencies listed in Table~\\ref{tab:frec} are due to nonradial oscillations. In fact, as is well known, most of the $\\delta$ Scuti stars are rapid rotators with low amplitudes of pulsations, and pulsating with nonradial modes (e.g. Fox Machado et al. 2006). Nonetheless, more precise information about its evolutionary stage is needed for a more conclusive study. A summary of the observations which led to the discovery and characterization of the new $\\delta$ Scuti star HD 207331 has been presented. The star shows complicated pulsations as do most of the $\\delta$ Scuti stars. To date our observations represent the most extensive work on HD 207331. \\subsection{Future Work} For the future, we will make spectroscopic observations of HD 207331 to obtain an accurate MKK spectral classification to fix more exactly the basic physical parameters of this star in order to be able to make a reliable seismic modeling. Also, spectroscopic observations will be carried out to test whether this $\\delta$ Scuti star rotates as rapidly as we might expect. In addition, more differential photometric observations, better distributed in time, are needed to better understand this interesting object." }, "1005/1005.1316_arXiv.txt": { "abstract": "% We investigate the use of mid-infrared PAH bands, continuum and emission lines as probes of star-formation and AGN activity in a sample of 100 `normal' and local ($z \\sim 0.1$) galaxies. The MIR spectra were obtained with the {\\it Spitzer} IRS as part of the Spitzer-SDSS-GALEX Spectroscopic Survey (SSGSS) which includes multi-wavelength photometry from the UV to the FIR and optical spectroscopy. The spectra were decomposed using PAHFIT (Smith et al. 2007), which we find to yield PAH equivalent widths (EW) up to $\\sim 30$ times larger than the commonly used spline methods. Based on correlations between PAH, continuum and emission line properties and optically derived physical properties (gas phase metallicity, radiation field hardness), we revisit the diagnostic diagram relating PAH EWs and \\neii/\\oiv\\ and find it more efficient as distinguishing weak AGNs from star-forming galaxies than when spline decompositions are used. The luminosity of individual MIR component (PAH, continuum, Ne and H$_2$ lines) are found to be tightly correlated to the total IR luminosity and can be used to estimate dust attenuation in the UV and in \\ha\\ lines based on energy balance arguments. ", "introduction": "We aim at determining the main source of ionizing radiation and star-formation rate of normal galaxies from MIR spectroscopy. ", "conclusions": "$\\bullet$ We find systematic trends between MIR spectral properties and optically derived physical properties, in particular between short wavelength PAH EWs and \\n2ha\\ (gas phase metallicity), and between \\neii/\\oiv\\ versus \\o3hb\\ (radiation field hardness) (Fig. 2, left panel); \\\\ $\\bullet$ The Genzel et al. (1998) diagram has better resolution using PAHFIT than spline decompositions. It is very similar to the optical ``BPT'' diagram (Fig. 2, right panel). The mixed SF/composite region may be revealing obscured AGNs in a large fraction of optically defined ``pure'' SF galaxies.\\\\ $\\bullet$ The PAH, continuum, Ne and H$_2$ luminosities are tightly and nearly linearly correlated to the total IR luminosity, less so to the dust corrected \\ha\\ luminosity (SFR) (Fig. 3, left panel);\\\\ $\\bullet$ Following \\cite{Kennicutt_etal2009}, the MIR components can be used to estimate dust attenuation in \\ha\\ and UV based on energy balance arguments (Fig. 3, right panel)." }, "1005/1005.4019_arXiv.txt": { "abstract": "s{ Current analyses of the Lyman-alpha forest assume that the primordial power spectrum of density perturbations obeys a simple power law, a strong theoretical assumption which should be tested. Employing a large suite of numerical simulations which drop this assumption, we reconstruct the shape of the primordial power spectrum using Lyman-alpha data from the Sloan Digital Sky Survey (SDSS). Our method combines a minimally parametric framework with cross-validation, a technique used to avoid over-fitting the data. Future work will involve predictions for the upcoming Baryon Oscillation Sky Survey (BOSS), which will provide new Lyman-alpha data with vastly decreased statistical errors. } ", "introduction": "The Lyman-$\\alpha$ forest is the name given to a series of absorption lines in quasar spectra, caused by the scattering of photons via interaction with neutral hydrogen at redshifts $2-4$. At these redshifts, a large proportion of the baryon density of the universe is contained within hydrogen clouds. Most of the hydrogen is ionized, but a small fraction remains neutral, and absorbs photons via the Lyman-$\\alpha$ transition. Hence, the Lyman-$\\alpha$ forest is sensitive to the matter power spectrum on scales from a few up to tens of Mpc, making it the only currently available probe of fluctuations at these weakly non-linear scales. A number of authors have examined the constraints obtainable from the Lyman-$\\alpha$ forest in the past, including Croft et al \\cite{croft}, Gnedin \\& Hamilton \\cite{gnedin}, Viel, Haehnelt \\& Springel \\cite{viel} . Previous analyses of constraints from the Lyman-$\\alpha$ forest have assumed that the primordial power spectrum is described by a nearly scale-invariant power law. This deserves further attention for a number of reasons. First, it is a strong assumption; if the data are inconsistent with it, derived constraints could be biased to some extent. Second, it is a generic prediction of inflationary models; hence, any test of a power law primordial power spectrum which cannot be attributed to data systematics is a test of inflation. Third, of all current datasets, the Lyman-$\\alpha$ constrains the smallest cosmological scales; thus, it provides the best opportunity presently available to understand the overall shape of the power spectrum. To do this, we shall reconstruct the primordial power spectrum in a minimally parametric way, using a technique called cross-validation to robustly recover the signal. If the data are in agreement with theoretical expectations, the recovered power spectrum will be nearly scale-invariant. In these Proceedings, we discuss a minimally parametric framework for constraining the primordial matter power spectrum, the cross-validation technique, and the methodology for obtaining constraints from observations. Finally, some preliminary results are presented. ", "conclusions": "" }, "1005/1005.3852_arXiv.txt": { "abstract": "We present a study of the average properties of luminous infrared galaxies detected directly at 24~$\\mu$m in the COSMOS field using a median stacking analysis at 70~$\\mu$m and 160~$\\mu$m. Over 35000 sources spanning 0$\\leq$$z$$\\leq$3 and $0.06$ mJy$\\leq$$S_{24}$$\\leq$3.0 mJy are stacked, divided into bins of both photometric redshift and 24~$\\mu$m flux. We find no correlation of $S_{70}/S_{24}$ flux density ratio with $S_{24}$, but find that galaxies with higher $S_{24}$ have a lower $S_{160}/S_{24}$ flux density ratio. These observed ratios suggest that 24~$\\mu$m selected galaxies have warmer spectral energy distributions (SEDs) at higher mid-IR fluxes, and therefore have a possible higher fraction of active galactic nuclei. Comparisons of the average $S_{70}/S_{24}$ and $S_{160}/S_{24}$ colors with various empirical templates and theoretical models show that the galaxies detected at 24~$\\mu$m are consistent with ``normal'' star-forming galaxies and warm mid-IR galaxies such as Mrk 231, but inconsistent with heavily obscured galaxies such as Arp 220. We perform a $\\chi^{2}$ analysis to determine best fit galactic model SEDs and total IR luminosities for each of our bins. We compare our results to previous methods of estimating $L_{\\rm{IR}}$ and find that previous methods show considerable agreement over the full redshift range, except for the brightest $S_{24}$ sources, where previous methods overpredict the bolometric IR luminosity at high redshift, most likely due to their warmer dust SED. We present a table that can be used as a more accurate and robust method for estimating bolometric infrared luminosity from 24 $\\mu$m flux densities. ", "introduction": "Although rare in the present-day universe, luminous infrared galaxies (LIRGs) were much more numerous in the past, and they may have played a significant role in the evolution of a large fraction of $L>L^{*}$ galaxies \\citep{Sanders:1996p245,Blain:2002p2091,Lagache:2004p2113,LeFloch:2005p2157}. However, their exact contribution is still poorly understood due to two limitations that have plagued deep surveys performed so far: (1)~the difficulty to identify the most obscured and distant of these objects, as well as measure their redshifts \\citep[they are often faint at optical wavelengths because of dust extinction,][]{Houck:2005p2186}, and (2) the difficulty to accurately characterize their nature (bolometric luminosity, mass, physical processes powering their energy output). Furthermore, because of limited sensitivity of current space- (\\emph{Spitzer} MIPS) and ground-based (SCUBA, BOLOCAM, MAMBO, AzTEC,...) observations in the far-IR/submillimeter, only a small number of the most luminous of these sources has been studied in detail. In addition, at high redshifts there are significant limitations due to confusion, which results from the very large instrument beam characterizing current far-IR/submillimeter observations. Many previous studies of LIRGs have been based on data obtained with the \\emph{Spitzer Space Telescope}, in particular with the Multiband Imaging Photometer \\citep[MIPS,][]{Rieke:2004p2277} at 24~$\\mu$m, the detector's most sensitive band. Using extrapolations based on libraries of galactic infrared (IR) spectral energy distributions (SEDs), the observed 24~$\\mu$m flux is converted to a bolometric IR luminosity, $L_{\\rm{IR}} \\equiv L(8$--$1000 \\mu$m), which is then used to calculate properties such as instantaneous star formation rate (SFR). However, at higher redshifts the 24~$\\mu$m band probes shorter rest frame wavelengths, probing rest frame 12 $\\mu$m at $z \\sim 1$, and rest frame 8 $\\mu$m at $z \\sim 2$. The typical peak of the IR SED of star-forming galaxies and galaxies containing active galactic nuclei (AGN) falls around 50--200 $\\mu$m; at higher redshifts, the 24~$\\mu$m band probes wavelengths farther away from the peak of the IR SED and begins to be heavily affected by broad mid-infrared PAH emission and silicate absorption features. Observations at longer wavelengths, such as in the \\emph{Spitzer} MIPS 70~$\\mu$m and 160~$\\mu$m bands (which probe rest frame 24~$\\mu$m and 54 $\\mu$m at $z \\sim 2$, respectively), are needed to more accurately characterize the bolometric luminosity, especially at higher redshifts. However, the MIPS 70~$\\mu$m and 160~$\\mu$m bands are significantly less sensitive and have worse angular resolution than the 24~$\\mu$m band. This leads to a drastic decrease in the number of sources directly detected at 70~$\\mu$m and 160~$\\mu$m, and the galaxies that are detected are biased toward the most luminous sources. Therefore, we use a stacking analysis \\citep[as in][]{Dole:2006p1785,Papovich:2007p38} to study the average 70~$\\mu$m and 160~$\\mu$m flux densities of galaxies detected at 24~$\\mu$m. In using a stacking analysis we lose the ability to study individual galaxies, but find average properties of galaxies that would otherwise be undetectable. In this work we explore the average mid- to far-IR flux densities of galaxies detected at 24~$\\mu$m and derive a more accurate method to estimate bolometric IR luminosity. To accomplish this, we measure stacked 70~$\\mu$m and 160~$\\mu$m flux densities of galaxies detected at 24~$\\mu$m in the Cosmic Evolution Survey (COSMOS) field, binned in both redshift and 24~$\\mu$m flux. We use these stacked fluxes to examine the evolution of mid- to far-IR colors of galaxies as a function of luminosity and redshift. Our stacked fluxes are fit to libraries of galactic IR SED templates, from which we derive an estimate of the average bolometric IR luminosity. \\citet{Papovich:2007p38} carried out a similar study employing stacking at 70~$\\mu$m and 160~$\\mu$m, but their analysis was limited by area, with a significantly smaller number of sources. With an area almost 10 times larger, we obtain more reliable statistics and the ability to bin our sources in narrower bins of redshift and flux. Our stacked fluxes will eventually be merged with Herschel PACS (100 \\& 160~$\\mu$m), Herschel SPIRE (200--500~$\\mu$m), and SCUBA2 data to get the best sampled SEDs of the high-$z$ literature. Throughout this work we denote flux density, $f_{\\nu}$ in MIPS 24~$\\mu$m, 70~$\\mu$m, and 160~$\\mu$m bands as $S_{24}$, $S_{70}$, and $S_{160}$, respectively. When calculating rest-frame quantities, we use a cosmology with $\\Omega_{m} = 0.3$, $\\Lambda = 0.7$, and $H_{0} = 70$ km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "\\subsection{Comparison to Models}\\label{sec:comparez} In this section, we compare the results of our stacking analysis with the expected fluxes and colors from theoretical models and empirical templates. We first compare our stacked $S_{70}/S_{24}$ and $S_{160}/S_{24}$ flux ratios with empirical models of ``normal'' star forming galaxies by \\citet[][hereafter DALE]{Dale:2002p1836} and models of Arp 220 and Mrk 231 from the SWIRE template library \\citep{Polletta:2007p1905}. We then constrain the average IR luminosity for each bin by fitting to many libraries of theoretical models and empirical templates. Figure \\ref{fig:comparez} plots $S_{70}/S_{24}$ and $S_{160}/S_{24}$ color as a function of redshift, along with the expected values from the DALE models of star forming galaxies and the SWIRE models of Arp 220 and Mrk 231. The DALE models are a one parameter family of models and we show models that cover a range of $1 < \\alpha < 2.5$, which describe normal star-forming galaxies with $8.3 < log(L_{\\rm{IR}}) < 14.3$. Arp 220 is a well-studied galaxy representative of heavily obscured ULIRGs, while Mrk 231 is representative of galaxies with warm mid-IR colors which are known to host AGN. We use the code \\emph{Le Phare}\\footnote{http://www.cfht.hawaii.edu/~arnouts/lephare.html} developed by S. Arnouts and O. Ilbert to determine the flux ratios of the DALE and SWIRE models at varying redshifts. \\emph{Le Phare} is a data analysis package used primarily to compute photometric redshifts, but a preliminary phase of the code also computes theoretical magnitudes, given SED libraries and filter bands. We can see immediately that the average colors determined from our stacking analysis fall within the region spanned by normal star-forming galaxies and Mrk 231, but our colors do not match those of Arp 220 at any redshift. This suggests that our 24 $\\mu$m selection is biased against heavily obscured objects like Arp 220. \\subsection{Best-fit Model SEDs and Average Total IR Luminosity}\\label{sec:SEDfit} We use \\emph{Le Phare} to perform a $\\chi^{2}$ analysis to find best-fit galactic model SEDs for the stacked fluxes calculated in each bin. We use the empirical templates of \\citet{Dale:2002p1836}, \\citet{Lagache:2003p1867}, \\citet{Chary:2001p1425}, and the theoretical radiation pressure models of \\cite{Siebenmorgen:2007p1876}, finding the SED from each library that best fit our stacked data at the correct redshift. The average $S_{24}$, $S_{70}$, and $S_{160}$ in each bin are plotted along with the best fit models in Figures \\ref{fig:matchz0}--5h, arranged by redshift (Figures 5b--5h are only available in the online version of the paper). Fluxes from the ``non-detcted'' bins are shown as upper limits. The parameter space spanned by all four of the models is shaded to give an idea of the spread of possible SEDs that fit the data. For our model fits, we use four different libraries of ``normal'' star-forming galaxies. We do have a small fraction of \\emph{XMM-Newton}-detected sources that contain AGN, especially in the highest flux bins, but this does not imply that the mid-IR and/or far-IR fluxes of these galaxies is dominated by the AGN itself (many X-ray sources are PAH dominated in the infrared). Because of the small number of sources, we are unable to perform a separate stacking analysis of only these X-ray sources. Although we cannot account for the true contribution of AGN contamination, the tight fits we see suggest that most of our sources are indeed star formation dominated. From the maximum likelihood function of the $\\chi^{2}$ analysis, we estimate a median $L_{\\rm{IR}}$ and 1$\\sigma$ uncertainties in each bin of our stacking analysis. We repeat this measurement for each library separately, and then take the mean of the four $L_{\\rm{IR}}$ values to estimate the true luminosity. The dispersion of the luminosities derived from the different libraries is $\\sim6\\% (3\\sigma)$ in all our bins, which suggests that the four libraries are fairly consistent in their estimates of the best-fit $L_{\\rm{IR}}$. We add the 1$\\sigma$ uncertainties from each library in quadrature to estimate the error in $L_{\\rm{IR}}$, and find typical $3\\sigma$ errors around 3\\%. The average IR luminosities and errors measured for each of our bins are listed in Table \\ref{tab:LIR}. Bins classified as non-detections at both 70~$\\mu$m and 160~$\\mu$m are marked as upper limits. As expected, we see that total IR luminosity increases with $S_{24}$ and with redshift. These galaxies span a large range of IR luminosity, covering ``normal'' galaxies ($L_{\\rm{IR}} \\leq 10^{11} L_{\\sun}$), LIRGs ($10^{11} L_{\\sun} \\leq L_{\\rm{IR}} \\leq 10^{12} L_{\\sun}$), and ultra luminous infrared galaxies (ULIRGs; $L_{\\rm{IR}} \\geq 10^{12} L_{\\sun}$). To test our uncertainties, we also found best fit SEDs using 70~$\\mu$m and 160~$\\mu$m flux densities that were offset by the errors given in Tables \\ref{tab:70} and \\ref{tab:160} and then measured $L_{\\rm{IR}}$ from these fits. We find discrepancies much less than 6\\% from this trial, suggesting that our derived infrared luminosities are robust within the errors in our stacking analysis. Table \\ref{tab:LIR} gives a relation between \\emph{observed} $S_{24}$ and total IR luminosity, with no need for a $k$-correction. This is an extremely valuable tool given the poor sensitivity of longer wavelength instruments, and provides an effective way to estimate $L_{\\rm{IR}}$ when only having 24~$\\mu$m data. Some caution must be taken when using Table \\ref{tab:LIR} to estimate $L_{\\rm{IR}}$. Although the agreement between the best-fit models from many libraries is fairly robust, it is not a complete description of the errors. For most of these bins, we do not have data on the Rayleigh Jean side of our IR SEDs, which means we cannot estimate the cold dust component and its contribution to $L_{\\rm{IR}}$. This means that Table \\ref{tab:LIR} is valid under assumption that the libraries used are representative of the diversity of the SEDs beyond $160\\mu$m$/(1+z)$. These results will be tested in the near future with Herschel and SCUBA2. Figure \\ref{fig:matchLIR} shows a comparison of $L_{\\rm{IR}}$ calculated from our stacking analysis and $L_{\\rm{IR}}$ calculated from extrapolating mid-IR fluxes from only $S_{24}$ based on the Dale SED libraries, as is common practice. In general, the two methods are in good agreement at low redshifts, but the $L_{\\rm{IR}}$ calculated using only 24~$\\mu$m data is an overestimate of the true $L_{\\rm{IR}}$ at high redshifts, especially for the brighter $S_{24}$ sources. This is most likely due to the warmer dust SEDs that we find in the bright $S_{24}$ sources (Section \\ref{sec:color}). Our results are consistent with the findings of \\citet{Calzetti:2007p1434}, who find that rest-frame 24 $\\mu$m flux is a much better indicator of bolometric infrared luminosity than 8 $\\mu$m flux, and with the findings of \\citet{Papovich:2007p38}, who find that the $L_{\\rm{IR}}$ estimated without taking into account stacked 70~$\\mu$m and 160~$\\mu$m fluxes overestimates the true $L_{\\rm{IR}}$. To summarize, extrapolating a bolometric infrared luminosity from a 24 $\\mu$m flux density without taking into account 70~$\\mu$m and 160~$\\mu$m flux will result in an overestimate of $L_{\\rm{IR}}$ at high redshifts. Table \\ref{tab:LIR} will give a more accurate and robust estimate of $L_{\\rm{IR}}$ at these redshifts." }, "1005/1005.3549_arXiv.txt": { "abstract": "The photospheres of low-mass red giants show CNO isotopic abundances that are not satisfactorily accounted for by canonical stellar models. The same is true for the measurements of these isotopes and of the $^{26}$Al/$^{27}$Al ratio in presolar grains of circumstellar origin. Non-convective mixing, occurring during both Red Giant Branch (RGB) and Asymptotic Giant Branch (AGB) stages is the explanation commonly invoked to account for the above evidence. Recently, the need for such mixing phenomena on the AGB was questioned, and chemical anomalies usually attributed to them were suggested to be formed in earlier phases. We have therefore re-calculated extra-mixing effects in low mass stars for both the RGB and AGB stages, in order to verify the above claims. Our results contradict them; we actually confirm that slow transport below the convective envelope occurs also on the AGB. This is required primarily by the oxygen isotopic mix and the $^{26}$Al content of presolar oxide grains. Other pieces of evidence exist, in particular from the isotopic ratios of carbon stars of type N, or C(N), in the Galaxy and in the LMC, as well as of SiC grains of AGB origin. We further show that, when extra-mixing occurs in the RGB phases of population I stars above about 1.2 $M_{\\odot}$, this consumes $^3$He in the envelope, probably preventing the occurrence of thermohaline diffusion on the AGB. Therefore, we argue that other extra-mixing mechanisms should be active in those final evolutionary phases. ", "introduction": "Evolved low-mass stars show photospheric CNO isotopic ratios that in many cases are not reproduced by stellar evolutionary codes. In the past years it was recognized by many authors \\citep{boot,wbs,char,sb99} that these chemical anomalies derive from transport mechanisms linking the envelope to zones where partial H-burning occurs. We call these phenomena ``extra-mixing'' or ``deep mixing'' throughout this report. \\citet[][hereafter NBW03]{nol} presented a parametric study of such mixing episodes, suitable to account for the CNO abundances measured in presolar grains of AGB origin. The adopted formalism was based on two parameters, namely the rate of mass transport ($\\dot{M}$) and the temperature ($T_P$) of the deepest zones reached by the circulation. It was also demonstrated that important composition changes can occur without introducing feedbacks on the stellar luminosity, provided $T_P$ is kept low enough (typically, $\\Delta \\log~ T = \\log~T_{\\rm H} - \\log~T_P \\gtrsim$ 0.08$-$0.1, where $T_{\\rm H}$ is the temperature at which the maximum energy of the H-burning shell is released). Subsequently, physical models for extra-mixing have been explored, which avoid the difficulties previously found with rotationally-induced mechanisms \\citep{pal}. In particular, hydrodynamical models of diffusive processes induced by variations of the mean molecular weight $\\mu$ \\citep{ss69}, called {\\it thermohaline diffusion} \\citep{cz07}, were presented by \\citet{egg1,egg2}. They showed that $^3$He burning into $^4$He and two protons successfully induces the required $\\mu$ inversion ($\\Delta \\mu/\\mu \\simeq -$ 10$^{-4}$), thus driving mixing episodes. Complementarily, \\citet{bwnc}, \\citet{nord} and \\citet{den} suggested that extra-mixing might be driven by magnetic buoyancy, in a dynamo process operating below the envelope. The model by NBW03 referred explicitly to AGB stars, which are known to be the parents of most presolar grains \\citep{zin1}. Extra-mixing in these objects was considered as necessary by various authors \\citep{hoppe,zinner,heck,nit2}. However, a recent paper \\citep[][hereafter KCS10]{kar} sheds doubts on this requirement, suggesting that slow mass circulation on the RGB might be sufficient. In order to solve the above dilemma, we have to look at the observational data, asking whether there is any clear requirement for extra-mixing on the AGB. Our answer will be that this requirement exists and comes primarily from oxide presolar grains \\citep{nit1,choi,clay,nit2}. Then one must look for a similarly clear constraint from stars, and we find it in the carbon isotope ratios of C(N) giants. Other pieces of evidence (from Ba stars, C-enhanced metal poor stars, etc.), although relevant, will not be addressed here for reasons of space. In order to demonstrate all this we generalize the calculations by NBW03, extending them to cover also RGB stages and including two more values of the initial mass. In Section 2 we compare our results for RGB phases with constraints coming from presolar oxide grains of group 2. In Section 3 we integrate this by considering extra-mixing also in AGB stages; critical tests come again from presolar oxide grains (of group 2, but also of group 1). The evidence from O-rich AGB stars is also discussed. Finally, in Section 4 we comment on the information coming from C(N) stars and we derive preliminary conclusions, also addressing the physical mechanisms necessary to drive the transport. ", "conclusions": "Our results indicate that extended mixing on the AGB is required, when considering oxygen isotopes and $^{26}$Al/$^{27}$Al ratios in presolar grains. \\cite[In particular, the oxygen isotopes maintain their crucial role in constraining stellar physics, as early noticed by][]{dearb}. Stellar evidence also points to the same conclusion. This is so in particular for carbon isotopic ratios in C(N) stars. Indeed, while extra-mixing on the RGB is sufficient to explain such ratios (averaging around 60) in the C(N) star sample by \\citet{lamb}, this is no longer true for the stars of \\citet{ab02}. These authors agree rather well (on average) with \\citet{lamb} for the sources in common; however, they also identified a number of new C(N) stars (25\\%) with $^{12}$C/$^{13}$C ratios $\\simeq 10 - 40$. Mainstream SiC grains provide the same evidence (for 23\\% of them, in the St Louis database, the carbon isotope ratio is below 40, while the average value is around 60, as for C(N) stars). \\citet{ab02} found that even assuming, at the beginning of the AGB phase, $^{12}$C/$^{13}$C $\\simeq$ 12, as due to extra-mixing on the RGB, final C-isotope ratios in excess of 43 were always obtained at C/O=1, unless new extra-mixing episodes on the AGB were included. Hence, C(N) giants and SiC grains confirm that extra-mixing occurs on the AGB; this is required also by their oxygen isotopes \\citep{har1,kaha}. The requirements from O-rich stars and grains suggest however a more extended processing than for C(N) stars. As these last have, on average, a larger mass than MS-S giants \\citep{gb08}, the extra-mixing efficiency in population I seems to decrease with increasing stellar mass. For lower metallicities, extra-mixing effects are known to be enhanced \\citep{ch98,grat00}. In the KCS10 approach, C(N) stars and SiC grains with $^{12}$C/$^{13}$C ratios below about 33 (see also their Table 1) cannot be explained, so that the constraints of SiC grains and carbon stars with low carbon isotope ratios are not reproduced. Note that the difference between the values found by \\citet{kar} and by \\citet{ab02} for the $^{12}$C/$^{13}$C ratios in C-rich stars (without extra-mixing) is due to different choices for this same ratio at the end of the RGB phase, and for C/O. KCS10 also claim that they ``cover most of the range of observational data points''. They refer to C stars in the Galaxy and in the LMC clusters NGC1846 and NGC1978. Although their Figure 2 might give this impression, this is not the case for the clusters. In fact, O-rich and C-rich stars in each cluster represent an evolutionary sequence, whose C/O and $^{12}$C/$^{13}$C ratios must be explained together. In contrast with this requirement, KCS10 need, for explaining C(N) stars, initial C/O and $^{12}$C/$^{13}$C values of 10 and 0.32; for explaining O-rich AGB stars they instead need 22 and 0.23, respectively (see their Figure 2). KCS10 actually admit their inability at explaining the two constraints together; this failure shows that their approach is not correct, at least for NGC 1846. Conversely, a scenario accounting for both O-rich and C-rich stars was presented by \\citet{leb08}, introducing a moderate extra-mixing on the AGB. These authors also suggested an increased efficiency of extra-mixing for increasing values of the envelope C/O ratio, because the bottom of the convective envelope becomes progressively closer to the H-burning shell while the star climbs the AGB. For the same reason, the extra-mixing efficiency is expected to increase for AGB stars of low metallicity, as the convective envelope becomes hotter for them \\citep{cris}. The situation of NGC1978 is definitely more puzzling \\citep{lederer}: in that case, a fit to C-rich stars requires the concomitant absence of extra-mixing processes during both the RGB and AGB phases. Thus, it is hard to find a theoretical recipe suitable to reproduce the isotopic ratios of AGB stars in NCG1978 (both O-rich and C-rich) without invoking an ad-hoc solution for this peculiar cluster \\citep[see discussion in][]{lederer}. \\begin{figure}[t!!] \\centering{\\includegraphics[height=6.0cm]{mixf3.eps}} \\caption{Relative variation of the molecular weight ($\\Delta \\mu /\\mu$) for a 1.5 M$_{\\odot}$ star with a metallicity $Z = Z_{\\odot}/2$, in the layers where $^3$He burns. The inversion, present on the RGB, is reduced or erased on the AGB when $^3$He has been previously consumed (see labels for consumption factors).} \\label{three} \\end{figure} Finally, we would like to comment on the physical origin of extra-mixing on the AGB. A popular mechanism is today thermohaline diffusion \\citep{egg1,egg2}. However, it is unlikely that it occurs on the AGB of population I stars, when any deep mixing has been previously active on the RGB. Indeed, the envelope abundance of $^3$He, whose burning drives the mixing through a $\\mu$ inversion, would be considerably reduced. This is shown in Figure \\ref{three}, where we plot the mean molecular weight across the radiative region in RGB and AGB phases, for our 1.5 M$_{\\odot}$ star. While a $\\mu$ inversion of $\\Delta \\mu /\\mu \\simeq -3 \\times 10^{-4}$ (continuous line) is in fact driven by $^3$He burning on the RGB, the AGB cases (dashed lines) are more critical. They refer either to no destruction of $^3$He on the RGB ($^3$He$_{\\rm FDU}$), or to a consumption by factors 3 or 10 (this last case corresponds to the findings by KCS10). It is evident that, for the masses and metallicities considered here, the $\\mu$ inversion is either strongly reduced or erased if $^3$He has been previously consumed. As extra-mixing (hence $^3$He consumption) on the RGB of galactic disk stars is required by observations \\citep{cos,shetr}, the conditions for thermohaline diffusion might be suppressed in their AGB stages \\citep[see also][]{can}. Other mechanisms should therefore be looked for, including magnetic buoyancy \\citep{bwnc}. Note that thermohaline mixing was found to occur on both the RGB and the AGB (first interpulses) in a low-metallicity 1-$M_{\\odot}$ star \\citep{stan}. This is due to the known higher inventory of $^{3}$He in very low-mass stars \\citep{dearb1}. {\\bf Acknowledgements}. We acknowledge useful comments from G.J. Wasserburg and from two very constructive referee reports. We are indebted to E. Zinner and coworkers for maintaining the on-line repository of presolar grain abundances from which we took the measured data (http://presolar.wustl.edu/~pgd/). C. Abia acknowledges partial support by the Spanish grant AYA2008-04211-C02-02" }, "1005/1005.4475_arXiv.txt": { "abstract": "We selected a sample of 24 XMM-Newton light curves (LCs) of four high energy peaked blazars, PKS 0548$-$322, ON 231, 1ES 1426$+$428 and PKS 2155$-$304. These data comprise continuous light curves of 7.67h to 18.97h in length. We searched for possible quasi-periodic oscillations (QPO) and intra-day variability (IDV) timescales in the LCs of these blazars. We found a likely QPO in one LC of PKS 2155$-$304 which was reported elsewhere (Lachowicz et al.\\ 2009). In the remaining 23 LCs we found hints of possible weak QPOs in one LC of each of ON 231 and PKS 2155$-$304, but neither is statistically significant. We found IDV timescales that ranged from 15.7 ks to 46.8 ks in 8 LCs. In 13 LCs any variability timescales were longer than the length of the data. Assuming the possible weak QPO periods in the blazars PKS 2155$-$304 and ON 231 are real and are associated with the innermost portions of their accretion disk, we can estimate that their central black hole masses exceed 1.2 $\\times$ 10$^{7}$ M$_{\\odot}$. Emission models for radio-loud active galactic nuclei (AGN) that could explain our results are briefly discussed. ", "introduction": "Blazars, including BL Lacertae objects (BL Lacs) and flat spectrum radio quasars (FSRQs), are extragalactic radio sources with relativistic jets aligned nearly ($\\lesssim$ 10$^{\\circ}$) with the line of sight (e.g., Urry \\& Padovani 1995). Blazar emission extends across the entire electromagnetic (EM) spectrum, is predominantly nonthermal, and shows significant polarization in the radio and visible bands where it can be measured. Blazars show detectable flux variations on diverse timescales ranging from a few minutes through days and months to decades through all EM bands. Blazar variability timescales have often been somewhat arbitrarily divided into three classes: timescales from minutes to less than a day are called intra-day variability (IDV); those from several days to a few months are known as short timescale variability (STV); while long term variability (LTV) covers changes from several months to many years (e.g., Gupta et al.\\ 2004). Short and intra-day X-ray variability has been investigated in various classes of AGNs (e.g., Edelson \\& Nandra 1999; Uttley et al.\\ 2002; Markowitz et al.\\ 2003; Vaughan et al.\\ 2003b; McHardy et al.\\ 2004, 2005; Espaillat et al.\\ 2008; Lachowicz et al.\\ 2009; and references therein). The great majority of these studies have been done for the brighter sources, which are typically nearby Seyfert galaxies, and were reviewed recently by Uttley (2007). There were claims of the detection of QPOs in a few AGNs on the short and IDV timescales in early X-ray observations (Fiore et al.\\ 1989; Papadakis \\& Lawrence 1993; Iwasawa et al.\\ 1998). But all of these claimed QPO detections were later found to be not statistically significant (Tagliaferri et al.\\ 1996; Benlloch et al.\\ 2001; Vaughan 2005; Vaughan \\& Uttley 2006). There have, however, been few stronger claims of QPO detection recently in X-ray data of various classes of AGNs. The first significant detection of a short X-ray QPO of $\\sim$ 1 hour timescale has been reported for RE J1034$+$396, which is a narrow line Seyfert 1 galaxy (Gierli{\\'n}ski et al.\\ 2008). Espaillat et al.\\ (2008) have reported an X-ray QPO on the timescale of 3.3 ks in 3C 273, a FSRQ. Very recently, Lachowicz et al.\\ (2009) have reported a detection of a X-ray QPO in the BL Lac PKS 2155$-$304 on a timescale of $\\sim$ 4.6 hours. All three of these QPO detections on IDV timescales were based on observations made with XMM-Newton. By using All Sky Monitor data from the Rossi X-ray Timing Explorer, Rani et al.\\ (2009) have reported a probable QPO from the BL Lac AO 0235$+$164 on a STV timescale of $\\sim$ 18 days. Such QPOs may shed new light on the physical processes at the source and associated X-ray emission and so the search for their presence in the light curves (LCs) of AGN is important. The full spectral energy distributions (SEDs) of blazars have double humped structures with each large spectral feature having comparable total powers. In these double humped SEDs, the first (lower energy) component of the SED is quite clearly dominated by synchrotron radiation from the relativistic jet and the second component is probably due to Inverse-Compton (IC) radiation. On the basis of their SEDs, blazars can be classified into two subclasses known as LBLs (Low-Energy Peaked Blazars) and HBLs (High Energy Peaked Blazars), though those with intermediate peaks are certainly found (Nieppola et al.\\ 2006). In LBLs, the synchrotron component peaks at near-IR, optical or near-UV frequencies and the high energy component usually peaks at GeV energies. In HBLs, the synchrotron component peaks somewhere in the X-ray band and the high energy component is seen to peak (or is extrapolated to do so) at TeV energies. Here, we study four HBLs which are designated as TeV blazars because of their significant detections in that extreme energy band. Just six years ago only six TeV blazars were known (see Krawczynski et al.\\ 2004 for a summary of their properties). Thanks to the development of new TeV facilities such as HESS (High Energy Stereoscopic System; Hofmann et al.\\ 2003; Funk et al.\\ 2004. Aharonian et al.\\ 2006), MAGIC (Major Atmospheric Gamma-ray Imaging Cerenkov; (Baixeras et al.\\ 2004; Cortina et al.\\ 2005; Albert et al.\\ 2006a) and VERITAS (Very Energetic Radiation Imaging Telescope Array System; Holder et al.\\ 2006; Maier \\ 2007, Maier et al.\\ 2008) there has been a complete revolution in TeV gamma-ray astronomy. These groups have detected about a dozen new TeV emitting blazars (HBLs) (Albert et al.\\ 2006b, 2007, 2008, 2009; Acciari et al.\\ 2008, 2009a, 2009b; Aharonian et al.\\ 2007, 2008a, 2008b, 2008c, 2008d, 2009a, 2009b, 2009c, and references therein). The enlarged sample of TeV blazars will be very useful for our understanding of the emission mechanism of these extreme blazars through the study of their variability properties across the range of EM bands. Here, we report our search for X-ray variability on IDV timescales in four HBLs from the XMM-Newton satellite public archive. We plan to perform extensive variability studies of HBLs in other EM bands using ground based observations. The motivation of this work is to examine the nature of IDV in a sample of four high energy peaked blazars, namely, PKS 0548$-$322, ON 231, 1ES 1426$+$428 and PKS 2155$-$304, in which synchrotron emission from the jets dominates up through the X-ray component. We will use IDV data to search for QPOs (quasi-periodic oscillations) and variability timescales in the light curves (LCs) of these HBLs, recalling that the shortest timescales can give upper limits to the sizes of the emitting regions. When a blazar is in a low flux state it is possible that an accretion disk related component of the X-ray emission is dominant and then any IDV timescale could give an estimate of the mass of the black hole presumed to reside at the center of the galaxy (e.g., Gupta et al.\\ 2008, 2009). This possibility arises from the fact that some blazars show evidence of the ``big blue bump'' that is almost certainly indicative of accretion disk emission (e.g., Raiteri et al.\\ 2007) while in their low states, when the jets are weak or even absent. Under these circumstances the X-ray corona expected above such disks and observed in radio-quiet AGN should be visible. A break in the power spectral density (PSD) plot from intra-day LCs can also yield the black hole mass of the blazar (e.g., Uttley 2007). XMM-Newton IDV LCs also may possibly be used to make independent estimations of the Doppler factors of the blazars (Fan, Xie \\& Bacon 1999). The paper is structured as follows. In Section 2, we give brief descriptions of the data selection criterion and the data reduction method. In Section 3, we discuss the techniques we used to search for variability properties and provide the results in Section 4. Our discussion is in Section 5 and the conclusions are given in Section 6. ", "conclusions": "" }, "1005/1005.0851_arXiv.txt": { "abstract": "Like the majority of spiral galaxies, NGC 6155 exhibits an exponential surface brightness profile that steepens significantly at large radii. Using the VIRUS-P IFU spectrograph, we have gathered spatially resolved spectra of the system. Modifying the GANDALF spectral fitting routine for use on the complex stellar populations found in spirals, we find that the average stellar ages increase significantly beyond the profile break radius. This result is in good agreement with recent simulations that predict the outskirts of disk galaxies are populated through stellar migration. With the ability to bin multiple fibers, we are able to measure stellar population ages down to $\\mu_V\\sim24$ mag/sq arcsec. ", "introduction": "The nature of how and where stellar disks end can shed light on the limits of star formation and the importance of secular evolution. Stellar disks tend to be embedded in much larger HI disks \\citep{Bosma81,Broeils97,Begum05}, showing that while there are baryons in the outskirts of galaxies, they are not being converted into stars efficiently. It is often assumed that there is a surface density threshold for star formation \\citep{Kennicutt89} which would provide a natural mechanism for truncating the stellar component of galaxies. The surface brightness profiles of spiral galaxies are commonly fit with exponentials \\citep{Freeman70}. While the exponential function is a convenient description, few astronomers would suggest that stellar disks extend to infinite radii. Observations of edge-on systems originally suggested that stellar disks have well defined truncation radii \\citep{vdk81a,Kregel04}. Using large samples of nearly face-on galaxies, \\citet{Pohlen06} found a variety of behaviors at large galactic radii. \\citet{Pohlen06} find that only 10\\% of disks are well fit by a single exponential light profile, while in 60\\% of their sample the surface brightness profile is well fit by a downbending broken exponental, with a break radius 1.5-4.5 times the inner scale length. The final 30\\% of their sample are described with a broken upbending exponential profile. Unlike the edge-on studies, they do not find evidence for sharp truncations. \\citet{Roskar08} present a possible formation mechanism for surface brightness profile breaks. They find that a star formation threshold seeds the onset of a truncation, but radial migration populates the disk beyond the break producing a down-bending broken profile. Since the stellar migration is a random walk, only ancient stars have had enough time to get to very large radii beyond where star formation shuts off. These simulations thus predict that the break radius should correspond with an increase in the average stellar ages. The formation and appearance of outer disks has since been investigated in cosmological simulations by several groups who also find an upturn in age at the break radius \\citep{Sanchez09b,Martinez09}. However, the relative importance of radial migration and in-situ star formation in a fully-cosmological setting remains an open question. While simulations find that profile breaks can be caused by a combination of a star formation threshold and stellar migration, observations suggest that star formation can proceed even in the outskirts of disks. Deep \\ha\\ imaging by \\citet{Ferguson98} show signs of extended star formation. \\citet{Thilker08} also find that $\\sim30$\\% of spiral galaxies have UV emission extending beyond their traditional star formation threshold radius. Similarly, \\citet{Herbertfort09} find clustering of point sources around NGC 3184, implying star cluster formation in the outer disk. Recent observations of resolved stars have found a variety of results for stellar truncation regions. \\citet{Williams09} find that in low-mass M33 the the age gradient changes from negative to positive at the truncation radius. Meanwhile, \\citet{deJong07} find that the truncation region in NGC 4244 is the same for stars of all ages. \\citet{deJong07} interpret this as a sign the truncation is the result of a dynamical interaction, but these results were subsequently shown to be consistent with the radial migration model \\citep{Roskar08}. See \\citet{Vlajic10} for a review of the recent advances in studying the outskirts of spiral disks. The observational results point to an interesting conundrum that while many studies find active star formation in the outskirts of disk galaxies, the stars in the region are often older than the inner disk. To date, truncation regions have been studied with broadband colors and resolved stars. There are a limited number of systems we can resolve, and faint broadband surface photometry is dominated by systematic errors associated with flat-fielding and sky subtraction. While there have been searches targeting emission lines \\citep{Ferg98,Chirst10}, in this letter we present the first spectroscopic study of stellar continuum spectral features across a surface brightness profile break. ", "conclusions": "\\citet{Pohlen06} describe NGC 6155 as lopsided and classify the galaxy as a Type II-AB (a downbending profile in an asymmetric disk). While our fiber-photometry shows slight lopsidedness, the velocity field is quite regular and there are no nearby neighbors. All the galaxies within a 1 Mpc projected radius and similar redshift of NGC 6155 are at least 3 magnitudes fainter. It therefore seems unlikely that the profile break in NGC 6155 is the result of a dynamical interaction with a neighboring galaxy. Our result that average stellar ages increase beyond the break radius is consistent with the results of \\citet{Bakos08} who stacked multiple SDSS images and found the $g-r$ color became redder beyond the profile break. This is also similar to the age gradient observed in M33, with the age decreasing out to the profile break where the gradient reverses and the stellar population becomes older beyond the break \\citep{Williams09}. \\citet{Bakos08} claim that while their stacked surface brightness profiles shows a break, the stellar mass profile is described by a single exponential. Our model mass-to-light values are not well constrained beyond the profile break, however, the increase in age we detect does suggest the M/L is larger in this region. Therefore we expect the stellar mass profile to be smoother than the light profile. The increase in stellar age we measure is very similar to the simulations of \\citet{Roskar08} where the average stellar age jumps from 4 to 6 Gyr. The best fits beyond the break radius have $\\tau\\sim0-9.5$ Gyr, consistent with significant ongoing star formation in the region. We therefore take our results as evidence that while stellar migration can populate outer disks, our fits are consistent with some low-level \\emph{in situ} star formation as well. Along with the change in stellar age, the surface brightness break also corresponds with a lack of line emission (Figure~\\ref{sb}, lower left). This could be a simple coincidence where the break radius is also where the emission becomes faint enough that we no longer detect it, but it is also consistent with the theory that the break radius represents the location of a star formation threshold. It should be pointed out that our observations are also consistent with a radial change in the stellar IMF. If the outer disk only hosts lower-mass molecular clouds, we could expect star formation to be unable to produce the highest mass stars \\citep{Kroupa03,Koeppen06}. Because young massive stars contribute heavily to the overall luminosity, their absence would result in a young stellar population that spectroscopically appears older. A change in the IMF can also explain the lack of emission lines beyond the break \\citep{Pflamm09}. In a forthcoming paper, we will present a larger sample of similar observations. Our larger sample includes galaxies with no surface brightness profile break as well as up-bending profiles. We also plan to further refine our spectral synthesis template fitting method to improve the metallicity measurements of the stellar populations." }, "1005/1005.0256_arXiv.txt": { "abstract": "We measure the topology of the main galaxy distribution using the Seventh Data Release of the Sloan Digital Sky Survey, examining the dependence of galaxy clustering topology on galaxy properties. The observational results are used to test galaxy formation models. A volume-limited sample defined by $M_r<-20.19$ enables us to measure the genus curve with amplitude of $G=378$ at $6h^{-1}$Mpc smoothing scale, with 4.8\\% uncertainty including all systematics and cosmic variance. The clustering topology over the smoothing length interval from 6 to $10 h^{-1}$Mpc reveals a mild scale-dependence for the shift ($\\Delta\\nu$) and void abundance ($A_V$) parameters of the genus curve. We find substantial bias in the topology of galaxy clustering with respect to the predicted topology of the matter distribution, which varies with luminosity, morphology, color, and the smoothing scale of the density field. The distribution of relatively brighter galaxies shows a greater prevalence of isolated clusters and more percolated voids. Even though early (late)-type galaxies show topology similar to that of red (blue) galaxies, the morphology dependence of topology is not identical to the color dependence. In particular, the void abundance parameter $A_V$ depends on morphology more strongly than on color. We test five galaxy assignment schemes applied to cosmological N-body simulations of a $\\Lambda$CDM universe to generate mock galaxies: the Halo-Galaxy one-to-one Correspondence model, the Halo Occupation Distribution model, and three implementations of Semi-Analytic Models (SAMs). None of the models reproduces all aspects of the observed clustering topology; the deviations vary from one model to another but include statistically significant discrepancies in the abundance of isolated voids or isolated clusters and the amplitude and overall shift of the genus curve. SAM predictions of the topology color-dependence are usually correct in sign but incorrect in magnitude. Our topology tests indicate that, in these models, voids should be emptier and more connected, and the threshold for galaxy formation should be at lower densities. ", "introduction": "Galaxy clustering has long been used to constrain cosmological models and to understand formation of galaxies. The most extensively studied clustering statistics are the autocorrelation function and the power spectrum. They are Fourier transforms of each other, and measure the clustering strength as a function of scale. The ``Standard Cold Dark Matter'' model (with the density parameter $\\Omega_m=1$, scale-invariant primordial fluctuations, and standard relativistic particle background), popular in the 1980s, was ruled out by showing that it was inconsistent with the observed galaxy correlation function (Maddox et al. 1990) and power spectrum (Vogeley et al. 1992; Park et al. 1992, 1994). These statistics are also used to measure the biasing in the galaxy clustering amplitude with respect to matter, and are key constraints on galaxy formation models connecting dark matter halos and luminous galaxies such as semi-analytic models (SAM; e.g., White \\& Frenk 1991; Kauffmann et al.\\ 1993; Cole et al. 1994, 2000; Benson et al. 2002; Bower et al. 2006; Cattaneo et al. 2006; Croton et al. 2006; De Lucia \\& Blaizot 2007; Monaco et al. 2007; Somerville et al. 2008) and halo occupation distribution models (e.g., Seljak 2000; Peacock \\& Smith 2000; Scoccimarro et al. 2001; Berlind \\& Weinberg 2002; Kang et al. 2002; Berlind et al. 2003; Zheng, Coil, \\& Zehavi 2007). Topology analysis was introduced by Gott et al. (1986, 1987) to test the Gaussianity of the primordial density fluctuations, which is one of the key characteristics of simple inflationary models (Bardeen et al. 1986). At large scales density fluctuations are still in the linear regime and maintain their initial topology, and it is possible to check whether or not the primordial fluctuations were a Gaussian field. At smaller scales, non-linear gravitational evolution and biased galaxy formation make the topology of the observed galaxy distribution deviate from the Gaussian form even if the initial conditions were Gaussian distributed as shown by perturbation theories and large N-body simulations (Park \\& Gott 1991; Weinberg \\& Cole 1992; Matsubara 1994; Matsubara \\& Suto 1996); using fractional volume rather than density threshold as the independent variable in topology analysis mitigates but does not eliminate these non-linear and biasing effects (Weinberg et al.\\ 1987; Melott et al.\\ 1988). Through studies of many observational samples, the topological properties of the large-scale distribution of galaxies have been examined (Gott et al. 1989; Park, Gott \\& da Costa 1992; Moore et al. 1992; Vogeley et al. 1994; Rhoads, Gott, \\& Postman 1994; Protogeros \\& Weinberg 1997; Canavezes et al. 1998; Park, Gott \\& Choi 2001; Hoyle, Vogeley \\& Gott 2002; Hikage et al. 2002, 2003; Park et al. 2005; James, Lewis \\& Colless 2007; Gott, Choi \\& Park 2009; James et al. 2009). On non-linear or quasi-linear scales, topology analysis is useful in constraining both cosmological parameters and galaxy formation mechanisms (Park, Kim \\& Gott 2005; Gott et al. 2008). In particular, differences in clustering topology for different types of galaxies reflect their different history of formation and evolution. Therefore, looking at the topology of large-scale structure traced by different types of galaxies can put strong constraints on galaxy formation mechanisms. In this paper we use the Seventh Data Release (DR7; Abazajian et al. 2009) of the Sloan Digital Sky Survey (York et al.\\ 2000) to measure the topology of the galaxy distribution and its dependence on galaxy luminosity, color, and morphology. DR7 constitutes the final release of the SDSS Legacy Survey, and thus of the main SDSS galaxy redshift survey. We supplement the SDSS data with missing redshifts to increase the completeness of the redshift catalog. We then generate a set of volume-limited samples of the SDSS galaxies divided according to their luminosity, morphology, and color to study the relation between the topology and properties of galaxies tracing the large-scale structure. ", "conclusions": "We use the SDSS DR7 main galaxy catalog supplemented with missing redshifts and with increased spectroscopic completeness to measure the galaxy clustering topology over a range of smoothing scales. The distribution of galaxies observed by the SDSS reveals extremely diverse structures. A volume-limited sample, BEST defined by $M_r<-20.19$, enables us to measure the genus curve with amplitude of $G=378\\pm 18$ at a smoothing scale of $6h^{-1}$Mpc, with the quoted uncertainty including all systematics and cosmic variance. The amplitude is 5.4 times larger than our previous measurement using the SDSS DR3 sample (Park et al. 2005) and the uncertainty decreases from 10.5\\% to 4.8\\% at the same smoothing length of $R_G=6h^{-1}$Mpc. We calculate the galaxy clustering topology over the interval from $R_G=6h^{-1}$Mpc to $10h^{-1}$Mpc, and find mild scale-dependence for the shift ($\\Delta\\nu$) and void abundance ($A_V$) parameters. The measured genus curve is qualitatively similar to the form predicted for Gaussian primordial fluctuations (Hamilton et al. 1986), but the differences are statistically significant at these scales: a shift of the peak towards negative $\\nu$, and fewer isolated voids and isolated clusters than the Gaussian prediction ($A_C$ and $A_V < 1$). The bias in topology of galaxy clustering with respect to that of matter is measured by assuming that the matter density field is given by our $\\Lambda$CDM N-body simulation. We detect strong topology bias in galaxy clustering, which is also scale-dependent. We confirm the luminosity dependence of galaxy clustering topology discovered by Park et al. (2005). The distribution of brighter galaxies is more shifted towards ``meatball' topology (lower $\\Delta\\nu$) and shows greater percolation of voids (lower $A_V$). We find galaxy clustering topology depends also on morphology and color. Even though early (late)-type galaxies show topology similar to that of red (blue) galaxies, morphology-dependence of topology is not identical with color-dependence. In particular, the void abundance parameter $A_V$ depends on morphology more strongly than color. We tested five galaxy formation models, which are used to assign galaxies to the outputs of N-body simulations. Three of these are semi-analytic models, one is an HOD model that assigns galaxies to halos with parameters tuned to match other clustering statistics, and one is a scheme that assigns galaxies to halos and subhalos. None of them reproduces all aspects of the observed topology, though the differences from one model to another are comparable to the discrepancies with the observations. For this reason, and because the initially Gaussian $\\Lambda$CDM model successfully reproduce the observed topology of LRGs at large scales, we attribute the discrepancies to failures of the galaxy formation model rather than non-Gaussian initial conditions. The semi-analytic models can also predict the topology of color subsets, but none of them fully captures the observed topology differences between red and blue galaxies. In future work, we will investigate models with non-Gaussian initial conditions to see what levels of primordial non-Gaussianity can be ruled out by our measurements. In principle, the high-precision topology measurements presented here and by Gott et al. (2009) can provide valuable constraints on non-standard inflationary models or alternative hypotheses for the origin of primordial fluctuations. Appendix A details our estimates of systematic biases in the genus curve measurements, demonstrating that the dominant effect is peculiar velocity distortions in redshift space rather than sample geometry or boundary effects. Except where noted otherwise, observational measurements in this paper are corrected for these biases, so they can be compared to theoretical predictions in real space with periodic boundaries. Appendix B investigates error covariances, showing that while the individual points on the genus curve have strongly covariant errors, the statistics $G$, $\\Delta\\nu$, $A_V$, and $A_C$ are approximately independent. Appendix C tabulates the full genus curves for our best samples, complementing the statistics recorded in earlier tables. While future surveys will use luminous galaxies and emission-line galaxies to probe structure in the distant universe, the SDSS DR7 sample is likely to remain the definitive map of large scale structure at low redshift traced by a broad spectrum of galaxy types, for the foreseeable future. The measurements in this paper characterize the topology of this definitive sample, attaining unprecedented statistical precision and providing a valuable test for future models of primordial fluctuations and galaxy formation physics." }, "1005/1005.5645_arXiv.txt": { "abstract": "The fundamental plane of early-type galaxies is a rather tight three-parameter correlation discovered more than twenty years ago. It has resisted a both global and precise physical interpretation despite a consequent number of works, observational, theoretical or using numerical simulations. It appears that its precise properties depend on the population of galaxies in study. Instead of selecting a priori these populations, we propose to objectively construct homologous populations from multivariate analyses. We have undertaken multivariate cluster and cladistic analyses of a sample of 56 low-redshift galaxy clusters containing 699 early-type galaxies, using four parameters: effective radius, velocity dispersion, surface brightness averaged over effective radius, and Mg2 index. All our analyses are consistent with seven groups that define separate regions on the global fundamental plane, not across its thickness. In fact, each group shows its own fundamental plane, which is more loosely defined for less diversified groups. We conclude that the global fundamental plane is not a bent surface, but made of a collection of several groups characterizing several fundamental planes with different thicknesses and orientations in the parameter space. Our diversification scenario probably indicates that the level of diversity is linked to the number and the nature of transforming events and that the fundamental plane is the result of several transforming events. We also show that our classification, not the fundamental planes, is universal within our redshift range (0.007 -- 0.053). We find that the three groups with the thinnest fundamental planes presumably formed through dissipative (wet) mergers. In one of them, this(ese) merger(s) must have been quite ancient because of the relatively low metallicity of its galaxies, Two of these groups have subsequently undergone dry mergers to increase their masses. In the k-space, the third one clearly occupies the region where bulges (of lenticular or spiral galaxies) lie and might also have formed through minor mergers and accretions. The two least diversified groups probably did not form by major mergers and must have been strongly affected by interactions, some of the gas in the objects of one of these groups having possibly been swept out. The interpretation, based on specific assembly histories of galaxies of our seven groups, shows that they are truly homologous. They were obtained directly from several observables, thus independently of any a priori classification. The diversification scenario relating these groups does not depend on models or numerical simulations, but is objectively provided by the cladistic analysis. Consequently, our classification is more easily compared to models and numerical simulations, and our work can be readily repeated with additional observables. ", "introduction": "Physical understanding of astrophysical objects most often uses correlation diagrams. For early-type galaxies, such scaling laws have been for instance established on one hand between optical luminosity and central velocity dispersion $\\sigma$ \\citep{FaberJackson1976}, and on the other hand between their effective radius $R_e$ and surface brightness averaged over effective radius $<\\mu_e>$ \\citep{Kormendy1977}. These correlations are rather tight, but the scatter is still reduced using a three-parameter relation of the form : $\\log{R_e} = a \\log\\sigma + b <\\mu_e> +\\ c$ \\citep{Dressler1987,Djorgovski1987}. This relation extends to faint and low-mass galaxies \\citep[e.g.][]{Nieto1990}. This is the fundamental plane (hereafter FP) of early-type galaxies. A long-running difficulty is the so-called ``tilt'' of the FP with respect to the ``virial plane'' obtained with the virial theorem and some simple assumptions about the population of early-type galaxies. Indeed, this tilt is different for different types of galaxies, like disk ones \\citep[e.g.][]{Robertson2006,Hopkins2008}. Many studies have been devoted to this problem without a clear solution. The motivation is to obtain pure correlations both to constrain the models better and to use them as probes of characteristics which are difficult to measure or strongly biased. In particular, the FP could in principle be a powerful tool to measure distances. But a proper calibration is required, and this appears difficult with its different tilts depending on the galaxy populations. Interpretation of the FP very often assumes some homology which is defined by \\citet{Gargiulo2009} as: ``systems with density, luminosity and kinematics structures equal over the entire early-type sequence and with constant mass-to-light ratios''. \\citet{vanDokkum2003} define homology in such a way that the evolution of the FP is due only to the evolution of $M/L$. These definitions are certainly linked to the assumption that all early-type galaxies are assembled in the same way, like dissipational mergers \\citep{Robertson2006,Hopkins2008}. Reality is however more complicated \\citep[see e.g.][]{Bender1992, Jorgensen1996, Borriello2003, vanderMarel2007} and early-type galaxies are very probably not all the result of the same formation process. Basically, what is needed is some invariant that allows us to trace a given object or class of objects through changes due to evolution. This invariant has been hoped to be the FP relation with some universality among a given population, universality which is implied by the definition of homology above and characterized essentially by $M/L$. This provides a rough way of simplifying the many variables that can evolve and may hide important sides of galaxy evolution. Models assuming only one homologous population have apparently failed to fit the FP in its entire extent. Selection criteria were proposed in order to obtain more homogeneous samples and thus define what could be called a ``purer'' fundamental plane. However, these criteria are necessarily arbitrary, subjective and/or model dependent. The difficulty is that many parameters are known to influence the global shape of the FP. Tilt of the FP \\citep{Robertson2006}, warps, dispersion, changes with redshift, dependence of the mass-distribution on mass \\citep{Nigoche-Netro2009} among others, show that even though the FP looks tight in the $\\log\\sigma$, $<\\mu_e>$ and $\\log{R_e}$ space, additional parameters could still play a role. The universality of the FP is also questioned: is it a bent plane \\citep{Gargiulo2009} or a bent surface approximated by a collection of planes \\citep{DOnofrio2008}? Obviously, this problem is related to the choice of the sample, that is to the definition of homology. Theorists cope with many parameters that may influence the evolution of a given galaxy, especially when mergers are considered \\citep{Robertson2006}. Testing parameters one after the other, both theoretically and observationally, takes a lot of time, and might partly explain why after so many years, this tight correlation keeps most of its mysteries. We think that it is time to explore new methodologies to better characterize and understand the FP relation. Multivariate clustering approaches are more objective in selecting really multivariate ``homologous'' sub-populations of galaxies. This requires to explicitly assume that the ``fundamental plane'' is a priori not universal, and to understand it as a correlation in the $(\\log{R_e}, \\log\\sigma, <\\mu_e>)$ space that could depend on the sub-population. Since galaxies are evolutive objects, homology can be more rigorously defined by 'similarity due to same class of progenitor'. It is the astrophysical equivalent of 'similarity by common ancestry' in cladistics, a statistical method designed to relate evolutionary objects and developed mainly by biologists. The use of many parameters is necessary to find the true homology, and to prevent analogy (same characteristics due to convergent evolution) to yield false lineages of galaxies. Cladistics does not \\emph{assume a priori} properties linked to homology, but rather relies on all the pertinent parameters to \\textit{construct} homologous groups. In the present paper, we have performed multivariate classifications with two independent approaches, cluster analysis and cladistic analysis, of a carefully chosen sample from the literature. For this first kind of study, we consider only the three parameters of the FP, $\\log\\sigma$, $<\\mu_e>$, $\\log{R_e}$, plus $Mg_2$ that are all given for this homogeneous sample. The first justification is that the number of parameters is small, so it is not necessary to use PCA because we can physically discuss the variation in the light of these few parameters easily. The second advantage is that we are interested to study the evolution of early-type galaxies in the light of their FP properties, and most of the authors have discussed evolution of galaxies on the basis of global FP. The first approach we have used is by a multivariate technique known as K-means Cluster Analysis using the parameters above. The second approach, known as astrocladistics, is based on the evolutionary nature of both galaxies and their properties \\citep{FCD06,jc1, jc2,DFB09,FDC09}. The clustering technique compares objects for their global similarities, while astrocladistics gathers objects according to their ``histories''. The two techniques are indeed complementary, the first one identifying coherent groups, the second one establishing an evolutionary scenario among groups of objects. They are also totally independent, so that the comparison of their results is extremely instructive from a statistical point of view. This paper is organized as follows. The data and the different multivariate analyses we performed are described in Sect.~\\ref{multivanal}. Comparison of the structures resulting from these analyses, characterization of the fundamental planes of individual groups, evolution properties within the global FP and properties of the groups, are presented in Sect.~\\ref{results}. The discussion on the meaning of the groups we have found is given in Sect.~\\ref{discussion}, before our conclusion in Sect.~\\ref{conclusion}. We present detailed descriptions of the cluster and cladistic analyses respectively in Appendix~\\ref{appendClus} and Appendix~\\ref{appendClad}, and provide additional diagrams in Appendix~\\ref{otherfigures}. ", "conclusions": "\\label{conclusion} In this paper, we have reconsidered the study of the so-called fundamental plane of early-type galaxies. We have used two different multivariate clustering tools, cluster analysis and cladistic analysis, to explore the 4-parameter space $\\log\\sigma$, $<\\mu_e>$, $\\log{R_e}$ and $Mg_2$. With both methods, we used both the three first observables, those of the fundamental plane correlation, and the four altogether. The sample used in our analysis, taken from \\citet{hudson2001}, has 699 objects spread into 56 galaxy clusters. In all the analyses, 4 to 7 groups are found, which are consistent from both statistical and physical arguments. The very good agreement between our different analyses provides a good confidence in the existence of structures \\textit{within} the FP. We emphasize that no a priori criterion is used to select groups of objects, even in the multivariate space. In this paper, we focus on the 7 groups (labelled C1 to C7) defined by the cladistic analysis because it additionally provides a diversification scenario linking the groups. Note that since we are in a multivariate space, the wording ``diversification'' is more appropriate than ``evolution'' that applies to a single parameter or that has the general meaning of ``transformation with time''. The groups define separate regions on the global fundamental plane, not across its thickness. In fact, each group shows its own fundamental plane, which is more loosely defined for less diversified groups. We conclude that the global FP is not a bent surface, but made of a collection of several groups characterizing several fundamental planes with different thicknesses and orientations in the parameter space. In addition, since all groups are present in all galaxy clusters, we conclude that our classification, not the FPs, is universal within the redshift range of the sample (0.007 -- 0.053). By design in the cladistic analysis, each group supposedly gathers objects sharing a similar history. They are also related by evolutionary relationships that represent a diversification scenario. By rooting the scheme with the group of least metallicity, we find that the two most diversified groups (C6, C7) have the thinnest FP, together with an intermediate one (C3). It probably indicates that the level of diversity is linked to the number and the nature of transformation events like collapse, accretion, interaction and merging, dry and wet, and that the fundamental plane is the result of several such transforming events. Our groups have distinct multivariate properties that can thus be interpreted in the light of our current knowledge and understanding of galaxy evolutionary processes. Three groups (C1, C2 and C4) probably did not form by major mergers and have been strongly affected by interaction, some of the gas in C2 objects having possibly been swept out. Three other groups, C3, C6 and C7 have been formed by dissipative (wet) mergers because they follow a tight FP relation. In C3, this(ese)) merger(s) must have been quite ancient because of the relatively low metallicity of its galaxies. But contrarily to C7, both C3 and C6 have subsequently undergone dry mergers to increase their masses, more violent in C6 than in C3. Also, in the k-space, the C7 group clearly occupies the region where bulges (of lenticular or spiral galaxies) lie. It could have formed through minor mergers and accretions. It has been recognized that the properties of the FP depends much on the sample. Our approach strongly associates the ``sample'' with the ``evolutionary group''. Since galaxies are evolutive objects, homology, a concept often used for FP studies, can be more rigorously defined by 'similarity due to same class of progenitor'. Since galaxy histories are so complex, only multivariate studies can objectively construct homologous groups. Cladistics is indeed designed to build homologous groups and provides an evolutionary scenario that relates them. It must be understood that the assembly and interaction histories of nearby galaxies are necessarily complex and comprises often several significant transforming events \\citep[e.g.]{jc2}. The interpretation, based on specific assembly histories of galaxies, of our seven groups shows that they are truly homologous. Our work also shows that multivariate cluster analysis is able to find homologous groups even though it cannot predict the evolutionary relationships. The properties of the 7 groups clearly reveal that they differ in assembly histories. Since they have been obtained directly from several observables, the interpretation of the result does not depend on any a priori classification. In particular, it partly matches refined morphological classification because we used 3 parameters (among 4) that are linked to structural properties of galaxies. However our classification is more easily compared to models and numerical simulations. Our work can be readily repeated with additional observables. In addition, the diversification scenario relating these groups does not depend on models or numerical simulations. The astrocladistic analysis was based on the assumption that the four parameters $\\log{R_e}$, $<\\mu_e>$, $\\log\\sigma$ and $Mg_2$ are evolutive characters, that evolve and can characterize states of evolution of galaxies. The astrophysical interpretation of the diversification scenario and of the evolutionary groups demonstrates a posteriori that this assumption is correct. This is another proof that cladistics can be applied in astrophysics. Since both cluster and cladistic analyses are only valid for the sample in study and the variables used, the present study will be extended to other galaxy samples with more parameters." }, "1005/1005.2253_arXiv.txt": { "abstract": "{ We report on the environmental dependence of galaxy properties at $z=2.15$. We construct multi-band photometric data sets in the (proto-)cluster PKS1138-26 field and in the GOODS field. We then fit spectral energy distributions of the galaxies with model templates generated with the latest stellar population synthesis code and derive physical properties of galaxies from the fits. To quantify the environmental dependence of galaxy properties, a special care is taken of systematic errors -- we use data sets that have almost the same wavelength samplings, use the same code to fit SEDs with the same set of templates, and compare {\\it relative} differences between the two samples. We find that the PKS1138 galaxies have similar ages, shorter star formation time scales, lower star formation rates, and weaker dust extinction compared to the GOODS galaxies at $z\\sim2$. This trend is similar to that observed locally, suggesting that the environmental dependence of galaxy properties is already partly in place as early as $z=2.15$. We show that the PKS1138 galaxies assemble the bulk of their masses $\\sim1$ Gyr earlier than field galaxies, i.e., the galaxy formation depends on environment. Galaxy mergers should frequently occur during the first collapse of clusters and they might play an important role in driving the observed environmental dependence of galaxy properties at $z=2.15$. }{}{}{}{} ", "introduction": "The formation and evolution of galaxies in the Universe are dependent on environment in which galaxies live. In the local Universe, red early-type galaxies are the dominant population in rich galaxy clusters, while blue late-type galaxies are the dominant population in the low-density field. Not only galaxy properties, but the formation epoch of galaxies also depends on environment in the sense that cluster galaxies form earlier than field galaxies (e.g., \\citealt{kuntschner02,gebhardt03,thomas05}). Earlier studies concentrated on nearby galaxies, but with the recent advent of large telescopes, environment studies at $z\\sim1$ became possible. Interestingly, the environmental dependence of galaxy properties observed at $z\\sim1$ is already strong; clusters at $z\\sim1$ are dominated by red early-type galaxies (e.g., \\citealt{blakeslee03,nakata05,postman05,lidman08,mei09}, but see also \\citealt{cucciati06}). Also, the formation epoch of cluster galaxies measured at $z\\sim1$ is consistent with that observed locally \\citep{gobat08} . Although there is a clear sign of galaxy evolution between $z=1$ and 0 (e.g., \\citealt{elbaz07,cooper08}), one has to observe galaxies at even higher redshifts to fully quantify the environmental dependence of galaxy formation and evolution. Because of observational difficulties, only a few high redshift clusters are known so far, the highest redshift cluster being at $z=1.45$ \\citep{stanford06}. Higher redshift galaxies appear fainter and their rest-frame optical light migrates to the near-IR, where the sky background is brighter and it is challenging to observe faint galaxies. Furthermore, high redshift clusters are poor clusters as they are still fast growing according to the dark matter halo growth models \\citep{press74,springel05}. Such poor clusters are difficult to locate due to their weak density contrasts to the general field. There are a number of ways to find high redshift clusters, but one of the proven techniques is to look around high redshift radio galaxies \\citep{miley08}. While not all the radio galaxies are in over-density regions, many of them host clear over-densities of galaxies around them \\citep{venemans07} and they are called proto-clusters. Although over-densities of red massive galaxies are not necessarily confirmed around them, they likely virialize and evolve to clusters at lower redshifts. Among the several proto-clusters reported so far, PKS1138-26 at $z=2.15$ is one of the most promising proto-clusters for its clear over-density of spectroscopically confirmed galaxies by previous studies \\citep{miley08}. Early studies of the the PKS1138-26 radio galaxy at $z=2.15$ were performed by \\citet{pentericci97,pentericci98}, who reported a clumpy morphology of the radio galaxy. Followed by these initial observations, \\citet{kurk00} first reported an over-density of star forming galaxies around the radio galaxy. This region was then followed up by several authors. \\citet{pentericci00} performed spectroscopic follow-up observations of Lyman $\\alpha$ emitters reported in \\citet{kurk00} and confirmed 14 galaxies close to the radio galaxy redshift. \\citet{kurk04a,kurk04b} carried out further imaging and spectroscopic observations of the field targeting H$\\alpha$ emitters and confirmed another 10 galaxies at the cluster redshift. An X-ray observation has also been performed \\citep{pentericci02} and follow-up spectroscopy confirmed at least 5 X-ray sources at the cluster redshift \\citep{croft05}. By now, there are more than 20 objects confirmed at the cluster redshift. Detailed analyses of the radio galaxy and the surrounding region with the superb resolution imaging with HST have also been performed \\citep{miley06,zirm08,hatch08,hatch09}, which added further lines of evidence for the forming (proto-)cluster. In fact, \\citet{zirm08} reported on the forming cluster red sequence. Recently, \\citet{doherty09} carried out a near-IR spectroscopic follow-up observation and confirmed two massive red galaxies at the cluster redshift. Given the the convincing over-density of galaxies and wealth of imaging and spectroscopic data available in the field, PKS1138 is an ideal sample to study the environmental dependence of galaxy properties at this high redshift. In this paper, we perform an extensive analysis of galaxies around PKS1138 to quantify the environmental dependence of galaxy evolution and formation at $z=2.15$. The layout of the paper is as follows. We summarize our data in Section 2. We then describe details of our method of fitting spectral energy distributions of galaxies in Section 3. Before presenting our results, we perform extensive sanity checks in Section 4. Section 5 presents physical parameters of galaxies obtained from the fits as a function of environment at $z\\sim2$ and Section 6 discusses implications of our results for galaxy formation. Finally, we summarize the paper in Section 7. Unless otherwise stated, we adopt H$_0=70\\rm km\\ s^{-1}\\ Mpc^{-1}$, $\\Omega_{\\rm M}=0.3$, and $\\Omega_\\Lambda =0.7$. Magnitudes are on the AB system. We use the following abbreviations: IMF for initial mass function, SED for spectral energy distribution and SFR for star formation rate. ", "conclusions": "From the extensive SED fits of galaxies in PKS1138 and GOODS, we have found that massive galaxies in PKS1138 tend to have \\begin{enumerate} \\item similar ages \\item shorter star formation time scales \\item lower star formation rates \\item lower amounts of dust \\end{enumerate} \\noindent compared to those in GOODS. The combination of first and second points is interesting. Galaxies in PKS1138 and GOODS start forming stars at a similar epoch in a statistical sense, but PKS1138 galaxies form more rapidly. We recall that our definition of age is the time since the onset of star formation. The third and fourth points are basically by-products of the first and second points. At $z=2$, PKS1138 galaxies have already undergone intense star formation and their SFRs are rapidly declining, while GOODS galaxies are still actively forming stars due to their longer star formation time scales. This results in the lower SFRs in PKS1138 (third point). The fourth point can then be easily understood given the correlation between between SFR and dust amount (lower SFR galaxies have less dust, e.g., \\citealt{hopkins03}). Cluster galaxies have a shorter star formation time scale -- this is the same trend as observed in $z\\lesssim1$ clusters. For example, \\citet{gobat08} showed the same trend in a $z=1.2$ cluster based on a photo-spectroscopic analysis. It is striking that the trend holds even at $z=2.15$. The difference in the star formation time scale suggests that cluster and field galaxies may form in different ways. Let us discuss the formation and evolution of cluster and field galaxies in detail. We first introduce two parameters to further quantify the galaxy formation. We then extend the discussion and address the origin of the environmental dependence of galaxy properties observed at lower redshifts. We note a caveat here that our results are based on one (proto-)cluster only and the trends we observe may not represent global trends at $z=2$. A larger sample of $z=2$ clusters should be investigated to draw a global picture. \\subsection{The environmental dependence of galaxy formation} We have age, star formation time scale, and SFR for each galaxy. Assuming the exponentially decaying star formation histories, we can estimate the SFR at the onset of star formation, which we call initial SFR ($\\rm SFR_0$; see Fig. \\ref{fig:tau_model}). Fig. \\ref{fig:sfr0_mstar} plots $\\rm SFR_0$ as a function of stellar mass. The cluster galaxies tend to have higher $\\rm SFR_0$ -- the fractions of $\\rm SFR_0>1000M_\\odot\\ yr^{-1}$ are $0.61\\pm0.21$, and $0.39\\pm0.15$ in PKS1138 and GOODS, respectively. The Mann-Whitney test supports this difference (the median probability is $2$\\% and we obtain $<5$\\% probability in 74\\% of the realizations). PKS1138 galaxies have experienced more intense formation histories. Of course, all this is based on the assumption of the exponentially declining SFRs, and we have ignored effects of galaxy-galaxy mergers, which we will discuss later. If galaxies continue to form stars following the exponential decay, we can derive the time when the galaxies form bulk of their stars. We plot in Fig. \\ref{fig:m80_mstar} the epoch when galaxies form 80\\% of the stars that they would have at $z=0$ ($t_{0.8}$; see Fig. \\ref{fig:tau_model}). PKS1138 galaxies typically formed around $z\\sim3$ or higher, while GOODS galaxies typically formed below $z\\sim3$. We recall that we obtained $z_f\\sim4$ from the location of the red sequence in PKS1138 in Section 5.1, which is in agreement with what we find here. We find that $0.61\\pm0.21$ of galaxies in PKS1138 form the bulk of their stars at $z>3$, while the fraction is $0.17\\pm0.09$ in GOODS. The Mann-Whitney test suggests that they are likely different -- the median probability is $\\sim8\\%$ and we obtained $<5$\\% probabilities in 38\\% of the realization. Interestingly, the ages of the cluster and field galaxies are not very different (Fig. \\ref{fig:physical_params}a). The difference in $t_{0.8}$ is therefore due to their shorter formation time scales. The formation of cluster galaxies is a more intense event, they form in a shorter time scale, and they assemble the bulk of their stars earlier by $\\sim1-2$ Gyr than the field galaxies. \\begin{figure} \\centering \\includegraphics[width=8cm]{sfr0_mstar.eps} \\caption{ $SFR_0$ is plotted against stellar mass. The meanings of the symbols are the same as in Fig. \\ref{fig:physical_params}. } \\label{fig:sfr0_mstar} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=8cm]{m80_mstar.eps} \\caption{ $t_{0.8}$ is plotted against stellar mass. The meanings of the symbols are the same as in Fig. \\ref{fig:physical_params}. } \\label{fig:m80_mstar} \\end{figure} \\begin{figure} \\centering \\includegraphics[width=8cm]{tau_model3.eps} \\caption{ The median star formation histories of the PKS1138 and GOODS galaxies. The shaded area shows 68 percentile intervals. } \\label{fig:tau_model2} \\end{figure} To summarize, we measure the median $\\tau$ and SFR for $>10^{11}\\rm M_\\odot$ galaxies and illustrate the differences in star formation histories of galaxies in PKS1138 and GOODS in Fig. \\ref{fig:tau_model2}. This will the summary plot of the paper. As discussed above, PKS1138 galaxies have higher $\\rm SFR_0$ and shorter $\\tau$. PKS1138 experience much more intense galaxy formation at early times than GOODS galaxies. The galaxy assemblies are completed in a short time, and by the time we observe them ($\\sim2$ Gyr after the onset of star formation as shown in Fig. \\ref{fig:physical_params}a) their SFRs become low. The plot summarizes the points listed at the beginning of this section (except for the fourth one, which is a result of SFR-dust correlation). The plot indicates that PKS1138 galaxies and GOODS galaxies did not form in the same way. The formation of cluster galaxies is a more intense event. But, how can massive galaxies form on such a short time scale? In general, star formation activities have negative feedback to themselves. Hot, young stars ionize surrounding gas, which prevents further star formation. Also, supernova explosions of massive stars give negative feedback. If galaxies undergo an AGN phase, it might also give negative feedback. The formation of massive galaxies on a very short time scale may be difficult to understand in this respect, and that makes us speculate that these massive galaxies might have taken another route -- mergers. Galaxy formation takes place in high peaks of density fluctuations in the Universe. Clusters that we observe today were regions of large-scale over-densities, where many density peaks were embedded. On the other hand, field galaxies form in more isolated peaks. The current galaxy formation theories predict that small galaxies form first and they form progressively massive galaxies via mergers. In the early Universe, cluster galaxies must have had a lot better chances to merge with other galaxies than field galaxies simply because there are more galaxies around them. This accelerated frequency of mergers might be the cause of the rapid formation of cluster galaxies. A simple scenario of an equal mass merger can be helpful here to show that mergers tend to make $\\rm SFR_0$ higher and push $t_{0.8}$ to higher redshifts. Assuming that two equal mass galaxies merge into a single galaxy, the merger event will double the stellar mass of the galaxy. If the galaxy experiences starburts triggered by the merger, its stellar mass will be more than double and its SFR gets lower after the burst. We then observe this massive galaxy with low SFR, and depending on the time elapsed after the burst, we would fit a short $\\tau$ to this galaxy in order to reproduce its low SFR and large stellar mass. In fact, we find that the secondary burst discussed in Section 4.4 reduces the median $SFR_0$ of the PKS1138 galaxis by $\\sim30\\%$, although it is still high, $SFR_0=1600\\rm M_\\odot\\ yr^{-1}$, which might indicate that galaxies experienced more than one merger. On the other hand, the $SFR_0$ of the GOODS galaxies remain the same within $10\\%$. We are not yet sure if PKS1138 is a collapsed cluster or currently collapsing proto-cluster. But, even if it is a cluster, it must be a young system, and it is unlikely that galaxies had enough time to be affected by 'nurture' effects such as ram-pressure stripping. We may suggest that mergers in the early times of the cluster formation may be one of the key processes to establish the environmental dependence of galaxy properties observed at lower redshifts. We will further pursue this point in the following section. \\subsection{The Nature effects in the environmental dependence of galaxy properties} Galaxy properties such as colors and morphology are known to depend on environment in which galaxies reside. This environmental dependence is shaped by two effects -- nature and nurture effects. That is, both how galaxies form and how they evolve are important. Fig. \\ref{fig:m80_mstar} shows that only $\\sim1$ Gyr has passed since the PKS1138 galaxies formed bulk of their stars to the observed epoch. Assuming that PKS1138 is a virialized system, we apply a $2\\sigma$-clipped gapper method \\citep{beers90} and obtain a velocity dispersion of 400 $\\rm km\\ s^{-1}$ using spectroscopic redshifts within 2 arcmin, corresponding to a physical scale of 1 Mpc, from the literature \\citep{pentericci00,kurk04a,croft05}. The virial radius ($r_{200}$) of the cluster is 0.32 Mpc, giving a crossing time scale of $\\sim 1$ Gyr. Therefore, the time elapsed since the bulk formation of the PKS1138 galaxies is comparable to the crossing time scale of this cluster. This suggests that nurture effects may not have had enough time to fully work. They may have affected a fraction of galaxies, but they probably could not change the average properties of galaxies. We may be witnessing the nature effects in shaping the environmental dependence in PKS1138. We admit that it is not very straightforward to classify nature and nurture effects at this high redshift. Now, let us introduce two kinds of mergers; early-epoch mergers and late-epoch mergers. We refer to mergers occur during the first collapse of clusters as early-epoch mergers, and those occur afterwards as late-epoch mergers. This is to sort out the two effects, early-epoch mergers being nature effects (initial conditions) and late-epoch mergers being nurture effects (environmental effects). Of course, nurture effects also include ram-pressure stripping, harassment, etc. On the other hand, nature effects do not include these processes driven by the deep potential well or intracluster medium. In this classification, the environmental dependence of PKS1138 is likely due to early-epoch mergers because $\\sim1$ Gyr time is probably not enough for nurture effects to fully work. Galaxy clusters at $z\\sim1$ do not give useful information about nature effects because nurture effects have had enough time to fully operate ($\\sim4$ Gyr since the formation epoch) and it is not straightforward to disentangle the two effects. Our results suggest that nature effects are strong and they form the basis of the environmental dependence of galaxy properties. A way to probe the significance of the nature effects is to quantify morphology of the PKS1138 galaxies. If early-epoch mergers are an important effect, then we expect to observe more signs of recent interactions in PKS1138 than in GOODS (see below). Also, we expect to observe post-starburst galaxies if galaxies undergone interaction triggered starbursts. A near-IR spectroscopic follow-up campaign of PKS1138 is currently underway. We will be able to study spectral properties of the PKS1138 galaxies, which allow us to look deeper into star formation histories. To sum up, the strong nature effects may shape the environmental dependence of galaxy properties. An ultimate goal of environment studies will be to quantify the relative contribution of nature and nurture effects. But, that will require statistical work on $z\\sim2$ (proto-)clusters. Although the possible significance of nature effects we suggest here is based only on one (proto-)cluster PKS1138, let us further discuss it. It has an interesting implication for the build-up of the cluster red sequence. \\subsection{The massive end of the cluster red sequence} The sequence of red early-type galaxies is a ubiquitous feature of galaxy clusters. Over the last few years, there is an accumulating amount of evidence in the literature that the cluster red sequence grows from the massive end to the low-mass end (e.g., \\citealt{tanaka05,tanaka07a,koyama07,tanaka08,gilbank08}). The massive end of the red sequence cannot be formed via a simple fading of blue, star forming galaxies because such massive blue galaxies do not exist even at $z\\sim2$ (see Fig. \\ref{fig:cmd} and discussions in \\citealt{faber07}). We need mergers to form it. An interesting point here is that we observe the brightest tip of the red sequence in PKS1138, which was also noted by \\citet{zirm08}. The massive end of the red sequence in a young system -- this might be due to early-epoch mergers. Early-epoch mergers might have formed very massive galaxies during the first gravitational collapse of clusters. Early-epoch mergers should occur more frequently in cluster environments than in the field, and that helps explain why we do not observe red sequence in GOODS. As suggested by \\citet{zirm08}, the red sequence in PKS1138 may be being formed or just formed at the time of observation. This view is further supported by our observation that roughly half of the red sequence galaxies have high SFRs (Fig. \\ref{fig:physical_params}). This formation redshift of $z\\sim2$ is in line with predictions from the build-up of the red sequence observed in lower redshift clusters \\citep{tanaka07a}. We present in Fig. \\ref{fig:acs_blowup} ACS $I$-band images of the bright red galaxies in PKS1138 ($K_s<21$ and $J-Ks>1$). Here we only briefly discuss their morphologies and a detailed study will be presented elsewhere (Zirm et al. in prep). Interestingly, a half of the galaxies show disturbed morphology and/or have nearby companions, lending a support to the picture of the accelerated mergers in clusters. Two out of three apparently disturbed galaxies are detected in MIPS. Early-epoch mergers during the first collapse of clusters may form the brightest end of the red sequence, and at the same time, form the basis of the environmental dependence of galaxy properties. \\begin{figure*}[tbh] \\centering\\noindent \\includegraphics[width=4.5cm]{acs395.eps} \\includegraphics[width=4.5cm]{acs451.eps} \\includegraphics[width=4.5cm]{acs487.eps}\\\\ \\includegraphics[width=4.5cm]{acs495.eps} \\includegraphics[width=4.5cm]{acs589.eps} \\includegraphics[width=4.5cm]{acs632.eps} \\caption{ ACS $I$-band images of the bright red galaxies. } \\label{fig:acs_blowup} \\end{figure*} The low-mass end of the red sequence is built up at later times. Nurture effects may come in there. Once clusters form, intracluster gas and deep potential field, in addition to late-epoch mergers, can affect galaxies and terminate their star formation activities. In this way, moderate-low mass galaxies could become red and form the low-mass end of the red sequence. The combination of nature and nurture effects may have conspired to produce the down-sizing behavior of the observed build-up of the red sequence (e.g., \\citealt{tanaka05}). \\subsection{Proto-clusters at higher redshifts} Finally, we finish the discussion with a forecast for future (proto-)cluster studies. We can go back in time and see how the PKS1138 field looked like at higher redshifts as we have star formation histories of individual galaxies. A caveat of course is that we ignore all the early-epoch mergers occurred before the time of the observation (the only way to recover that information is to resolve the galaxies into individual stars). Another caveat is that we cannot track spatial positions of galaxies back in time. But, PKS1138 is likely a collapsing/collapsed system and it may well have been an over-density region already at $z=4$. Fig. \\ref{fig:lss_evol} shows distribution of galaxies and their star formation rates evolved back to $z=4$. Compared to Fig. \\ref{fig:lss}, galaxies are more actively forming stars on average. Interestingly, galaxies in PKS1138 are more actively forming stars and there are more starbursting galaxies with $\\rm SFR>100\\ M_\\odot\\ yr^{-1}$ than in GOODS. This is in stark contrast to Fig. \\ref{fig:lss}, where we saw that PKS1138 galaxies have lower SFRs than those in GOODS. We still observe a hint of a galaxy over density in PKS1138 at $z=4$. The plots suggest that, as we approach the formation epoch of clusters (i.e., early phase of the gravitational collapse to a massive cluster halo), we expect to observe an over density of starbursting galaxies. Along with starbursting galaxies, low SFR galaxies already appear in PKS1138, while such galaxies are extremely rare in GOODS. At this redshift, low SFR galaxies and starbursting galaxies may co-exist in a forming cluster. It is not easy to show how early-epoch mergers change the picture we see here. But, we still expect to observe starbursting galaxies triggered by early-epoch mergers in collapsing clusters with a higher over-densities of lower-mass, pre-merger galaxies. We deem that observations with existing/future sub-millimeter arrays would be able to discover many forming clusters at high redshifts. In fact, some submm observations of distant radio galaxies have found possible over-densities of dusty starburst populations around them (e.g., \\citealt{debreuck04,greve07}). Full wavelength observations will be essential for future proto-cluster studies at very high redshifts. \\begin{figure*} \\centering \\includegraphics[height=10cm]{lss_evol_1138.eps} \\includegraphics[height=10cm]{lss_evol_goods.eps} \\caption{ Same as Fig. \\ref{fig:lss}, but galaxies are evolved back in time to $z=4$ assuming the star formation histories obtained from the SED fits. The sizes of the symbols correlate with stellar mass. The top and right axes show comoving scales at $z=4$. } \\label{fig:lss_evol} \\end{figure*}" }, "1005/1005.5229_arXiv.txt": { "abstract": "The Lorentz factor (LF) of gamma-ray burst (GRB) ejecta may be constrained by observations of high-energy (HE) spectral attenuation. The recent Fermi-LAT observations of prompt GeV emission from several bright GRBs have leaded to conclusions of unexpectedly large LFs, $\\Gamma>10^3$. Here we revisit this problem with two main concerns. (1) With one-zone assumption where all photons are assumed to be generated in the same region (radius) and time, we {\\em self-consistently} calculate the $\\gamma\\gamma$ optical depth by adopting a target photon spectrum with HE cutoff. We find that this might be important when the GRB LF is below a few hundreds. (2) Recent Fermi-LAT observations suggest that the bulk MeV-range and HE ($\\ga100$~MeV) emission may arise from different regions. We then consider a two-zone case where HE emission is generated in much larger radii than that of the MeV-range emission. We find that the HE emission may be mainly attenuated by MeV-range emission and that the attenuated HE spectrum does not show an exponential spectral cutoff but a slight steepening. This suggests that there may be no abrupt cutoff due to $\\gamma\\gamma$ attenuation if relaxing the one-zone assumption. By studying the spectra of three bright Fermi-LAT GRBs 080916C, 090510 and 090902B, we show that a bulk LF of\\textbf{ $\\Gamma\\sim600$ }can be consistent with observations in the two-zone case. Even lower LFs can be obtained in the multi-zone case. ", "introduction": "Relativistic expansion is a key property of gamma-ray bursts (GRBs), and has been confirmed by measurements of radio afterglow sizes, for examples, the indirect estimation by radio scintillation in GRB 970508 \\citep{wkf98} and direct imaging of nearby GRB 030329 \\citep{taylor04}. These observations revealed mildly relativistic GRB ejecta, $\\Gamma\\sim$a few, in the radio afterglow phase. However, it is well believed that GRB ejecta are ultra-relativistic in the beginning-- this is required to solve the so-called \"compactness problem\" \\citep[e.g.,][]{piran99}. The compact GRB source, suggested by the rapid variabilities in MeV light curves, and the huge luminosity suggest hot, optically thick GRB sources, which is in confliction with the nonthermal and hard GRB spectra. Relativistic expansion of the emission region is introduced to solve this problem. In order for the $\\sim100$~MeV photons, as detected by EGRET in several GRBs, to escape from the emission region, avoiding $\\gamma\\gamma$ attenuation, the bulk Lorentz factor (LF) of the emission region is required to be extremely large, $\\Gamma\\ga10^2$ \\citep[e.g.,][]{ls01,krolik91,fenimore93,wl95,bh97}. Recently, the powerful Fermi satellite reveals in much more detail the high-energy (HE) emission from GRBs. Several bright GRBs are reported to show time-integrated spectra extending up to GeV or even tens GeV, without any signs of spectral cutoff. Assuming the $\\gamma\\gamma$ optical depth for these HE photons are below unity, these observations have leaded to even larger bulk LFs, $\\Gamma>10^3$ \\citep{a09a,a09b,a09c}. This is putting the theoretical problem of relativistic jet formation to extremes. In the previous constraints two assumptions are usually taken. First, all photons, from low to high energy, are produced in the same region and the same time. This \"one-zone\" assumption is not solid, as Fermi observations actually revealed that: the onset of HE emission is delayed relative to MeV emission \\citep[e.g.,][]{a09a,a09b,a09c}; the HE emission lasts longer than MeV emission \\citep[e.g.,][]{a09a,a09b,a09c}; the bulk emission shifts toward later time as the photon energy increases \\citep{a09a} and the shift is longer than the variability times in MeV light curves, as pointed out by \\citep{li10}; some GRBs obviously show distinct HE components with different temporal behaviors \\citep{a09b,a09c}. All these features may imply that different energy photons are produced in different regions. In particular, the bulk $>100$~MeV emission in GRB 080916C shows $\\sim1$~s shifting relative to MeV emission, which is much longer than the MeV variability time, $<100$~ms as revealed by INTEGRAL \\citep{greiner09}, strongly implying that $>100$~MeV emission is produced in a region of much larger radii than MeV emission's \\citep{li10}. As pointed out by \\cite{lw08}, within the framework of internal shock model, the internal collisions at small radii, which would produce the prompt MeV emission, are expected to lead to \"residual\" collisions at much larger radii, which would produce low-frequency emission. The electrons accelerated by residual collisions at larger radii inverse-Compton scattering the MeV photons and/or double scattering the low-frequency photons could produce HE emission \\citep{li10,zhao10}. In this case, MeV and HE photons are produced in different regions. In the comoving frame of HE emission region, the MeV photons would be collimated other than isotropic, thus the $\\gamma\\gamma$ absorption is angular dependent. Second, the target photon spectrum is assumed to be extending to infinity. As pointed out by \\cite{li10}, the calculation of $\\gamma\\gamma$ optical depth taking such a target photon filed is obviously not self-consistent, because the HE spectral end should be cut off due to absorption considered in the calculation. In this paper, we revisit the problem of GRB LF constraint by modifying the above mentioned two assumptions. We consider in \\S2 a one-zone case where the $\\gamma\\gamma$ optical depth is calculated {\\em self-consistently} by assuming a truncated target spectrum, then we consider in \\S3 a simple two-zone case with anisotropic effect on $\\gamma\\gamma$ optical depth taken into account. In \\S4 we studied the spectra of the three bright Fermi-LAT GRBs and constrain their LFs. \\S5 is discussion and conclusions. In the following we assume the concordance universe model with $(\\Omega_m,\\Omega_\\Lambda)=(0.27,0.73)$ and $H_0=71\\rm km\\,s^{-1}Mpc^{-1}$. ", "conclusions": "We have revisited in this work the problem of constraining the GRB LFs by the HE attenuation. Although this problem has been considered by many previous works, two concerns that have been ignored in the previous work have been emphasized here. First, we notice that in the one-zone case in order to self-consistently calculate the $\\gamma\\gamma$ optical depth one needs to consider the target photons with HE spectral cutoff, other than extending to infinity. This concern is important when the LFs are below a few hundreds, or when the luminosity of GRBs are low. Second, we relax the one-zone assumption and consider a simple two-zone case where the beaming of target photons in the emission region should be taken into account. Our results show that in the two-zone case, the $\\gamma\\gamma$ absorption does not lead to an abrupt spectral cutoff but a spectral steepening. If the target photon energy distribution is with a power law with photon index $\\bt$ then the spectral slope is changed by a factor of $\\frac1{\\bt}-1$. This also predicts that there should be no spectral cutoff in the GRB spectra if the prompt emission is not produced in one single region. It should be noted that there are some attempts by other authors to improve the approximation for the optical depth. \\cite{Baring06} concluded that the pair attenuation signature appears as broken power-law rather than exponential cutoff by considering the skin effect and introducing an attenuation descriptor of $1/(1+\\tau)$ instead of $e^{-\\tau}$. \\cite{Granot08} considered the emission zone as a very thin layer producing impulsive emission. They calculated in detail the opacity evolution during a pulse, and claimed that the attenuation signature can be different from that derived from the simple one-zone approximation. Essentially, these two works still concern one-zone problem, with $\\Delta R\\sim R_{\\rm MeV}$. However in the two-zone problem that we considered here, the HE and MeV emission components are emitted at very different radii, with $\\Delta R\\gg R_{\\rm MeV}$, which leads to much smaller optical depth and hence smaller LF at HE emission region. Furthermore, we take our new concerns to analyze the spectra of the three bright GRBs 080916C, 090510 and 090902B and found that in the two-zone case a LF of $\\Gamma\\sim600$ can still be consistent with the observed spectra. This relaxes the strict requirement, $\\Gamma>10^3$, in one-zone assumption. We note that in the present observational situation where only tens to hundreds HE photons detected in one GRB, a slight change of the spectral slope is not easy to be identified. A single power law may still fit the HE spectral tail. We have considered a simple two-zone case here. However the situation can be more complicated. The central engines of GRBs may naturally create variabilities in a wide range of timescales, e.g., from $\\sim1$~ms to $\\sim10$~s. In the framework of internal shock model, this will lead to kinetic dissipation in a wide range of radii. Even in the single-timescale case, the internal collisions will happen as the ejecta expand until the material is distributed with velocity increasing with radius. In such case we will expect multi-zone other than simple two-zone case. The time-integrated spectrum-- note that the time interval with high enough photon statistic is usually much larger than the variability time-- will be contributed by the multiple regions. We also calculate cases with $\\Gamma<600$, the sum of the flux at the HE end can be comparable to the original flux. This means that the spectra can be consistent with a multi-zone case with the LF $\\Gamma<600$. The formation and acceleration of relativistic collimated GRB jets are open questions. In the standard \"fireball\" model, the thermal pressure can only accelerate the gas up to a LF $\\Gamma\\la10^3$ \\citep[see, e.g.,][]{piran99,li10}. On the other hand, simulations of magnetic-driven jets \\citep[e.g.,][]{mhdjet} can generate jets with the product of the LF and jet opening angle being $\\Gamma\\theta_j\\approx10-30$, which is consistent with pre-Fermi GRB observations. However for the bright Fermi-LAT GRBs, \\cite{cenko10} constrained the jet opening angles by their afterglows, which, combined with the large LF, $\\Gamma>1000$ from $\\gamma\\gamma$ attenuation argument, suggests much larger values of $\\Gamma\\theta_j$. We stress here that if relaxing the one-zone assumption for GRB multi-band emission, LFs with \"normal\" values, say, $\\Gamma\\la600$, can still be consistent with observations. This relaxes further the theoretic problem of jet acceleration. Recently, a similar paper, \\cite{Zou10}, considering the same two-zone absorption, is now in preprint. The main difference between two papers is the rest frames for the optical depth calculation, i.e., we consider the comoving frame of the HE emission region while they consider the observer frame. They integrate over the region up to $R_{\\max}$, where the HE photon spatially leaves the MeV front, while our integration corresponding to one dynamical time expansion is equivalent to integration up to $2R$, where the generated HE photon doubles its radius. Because both the number density of the target photons and the angle between the travelling directions of HE photon and the MeV front decrease rapidly with radius, the interaction is strongly dominated by those at small radius, and hence the upper limit of the integration is unimportant-- no matter the upper limit is $R_{\\max}(\\gg R)$ or $2R$ the result is practically the same. Furthermore, they use an \"averaged\" optical depth to constrain the LF, which may not be appropriate since we have shown that no sharp cutoff is expected in the two-zone case. We consider more carefully the spectral profile due to suppression. Finally, they neglect the self absorption in the local region which may contribute significant effect as we show." }, "1005/1005.0310_arXiv.txt": { "abstract": "{ Blue straggler, which are stars that appear to be younger than they should be, are an important population of unusual stars in both stellar clusters and the halo field of the Galaxy. Most formation scenarios evoke either stellar collisions or binary stars that transfer mass or merge. } { We investigate high-velocity stars in the Galactic halo and perform a spectral and kinematical analysis to shed light on their nature and origin. Here we report that SDSSJ130005.62+042201.6 (J1300+0422 for short) is an A-type star of unusually large radial velocity (504.6 $\\pm$ 5 \\kms). } { From a quantitative NLTE (and LTE) spectral analysis of medium-resolution optical spectra, the elemental composition is derived. Proper motion measurements combined with a spectroscopic distance estimate allow us to determine its present space velocity. Its kinematical properties are derived by integrating the equation of motion in the Galactic potential. } { We find J1300+0422 to be metal poor ([M/H]=$-1.2$) and exhibit an $\\alpha$-element enrichment ($0.3-0.4$~dex) that is characteristic of the halo population, as confirmed by a kinematical analysis of its 3D space motions, which places it on a highly eccentric retrograde Galactic orbit. } { The mass of J1300+0422 (1.15 $\\pm$ 0.10 M$_\\odot$) is higher than the globular cluster turn-off masses indicating that it is a halo blue straggler star. At a Galactic rest-frame velocity of $\\approx$467~\\kms, the star travels faster than any known blue straggler but is still bound to the Galaxy. } ", "introduction": "\\label{sec:intro} Blue straggler stars (BSS) were first discovered as an unusual subclass of stars in globular clusters \\citep{1953AJ.....58...61S}. They lie on or near the main sequence but are more luminous than the turn-off stars indicating that they are of higher mass than the latter. Apparently, BSS are present in all the Galactic globular clusters \\citep{2003ASPC..296..263P}. Because the stars in a cluster are believed all to have formed at the same time, whereas the stellar turnoff age decreases with mass, BSs should have evolved off the main sequence to become giants and white dwarfs long ago. It is generally believed that blue stragglers are coeval with the remaining stars in the cluster and originate in normal main-sequence stars that gained mass by means of a recent accretion episode. Most formation scenarios evoke mass transfer in and/or the merger of binary stars or collisions of stars. It remains unclear whether two types of BSS exist in clusters, which may result from the different suggested mechanisms discussed frequently in the literature. The discovery of two distinct sequences of BSS in the globular cluster M~30 \\citep{2009Natur.462..1128N} indicates that indeed both formation mechanisms are at work. In open clusters, this does not seem to be the case, because \\citet{2009Natur.462..1132N} indentified a high frequency of binaries among the blue stragglers in the open cluster of NGC~188, most having eccentric orbits with periods of about 1000 days. \\citet{2009ApJ...690.1639B} argue that most if not all BSS in open clusters arise from multiple star systems. This suggests that blue stragglers are formed in both ways in star clusters, with collisions/mergers becoming more common with increasing cluster density. In the Galactic field, it is more difficult to identify BSS because the stellar age cannot be determined. Field blue stragglers are therefore usually identified as metal-poor main sequence objects that are hotter than globular cluster main-sequence turnoff stars. Nevertheless, many blue metal-poor stars $[Fe/H]<-1$ with main-sequence luminosities have been found to be hotter than the main-sequence turnoff of globular clusters and are therefore considered to be field analogs of the cluster BSS population. These metal-poor stars seem to be so numerous that their specific frequency of appearance relative to regular horizontal branch stars was found to be higher than in globular clusters \\citep{2000AJ....120.1014P}. Since the stellar density is low in the field, binary star evolution and mass transfer is probably the most common path of formation among field BSS. \\cite{2001AJ....122.3419C,2005AJ....129.1886C} pointed out that some BS candidates among metal-poor halo main-sequence stars are not binaries, hence binary star evolution and mass transfer may not be the only path. They argue that the apparently single stars could be the partial remnants of an accreted dwarf satellite galaxy whose star formation continued over a long period of time. These metal-poor main-sequence stars are simply therfore young stars. A detailed spectral analysis of a blue metal-poor star sample in the field by \\cite{2000AJ....120.1014P} measured a very high binary fraction of at least 67\\%, dominated by long period (wide) binaries. This population appears to be very similar to that of the open cluster NGC~188 \\citep{2009Natur.462..1132N}. Assuming a formation by Roche lobe overflow during the red giant branch stage of the primary, \\cite{2000AJ....120.1014P} identified at least half of the blue metal-poor stars in their survey as blue stragglers. This is supported by more recent and much larger surveys for faint blue stars in the halo. For example, \\cite{2008ApJ...684.1143X} found that blue stragglers account for half of their sample of more than 10\\,000 A-type stars. We have embarked on a search for so-called hyper-velocity stars \\citep[HVS,][]{2009A&A...507L..37T} based on the work of \\citet{2008ApJ...684.1143X}, who presented radial velocities for a large sample of blue stars. Their sample is a mix of blue horizontal branch (BHB), blue straggler, and main-sequence stars with effective temperatures roughly between 7000 and 10\\,000\\,K according to their colours. To study the stellar motions in 3D, we need to derive their tangential velocities from proper motions and spectroscopic distances. The measurement of proper motions for faint high-velocity stars is the most challenging part of this project because the stars' distances are large and therefore require highly accurate proper motion measurements. We focused on the fastest stars in terms of large Galactic rest-frame (GRF) velocities with the aim of determining their nature, distance, and kinematics from detailed quantitative spectral analyses and astrometry. First results have already been reported \\citep{2009A&A...507L..37T,2009_J1539}. We found J0136+2425 to be an A-type main-sequence star travelling at $\\approx$590~\\kms, possibly unbound to the Galaxy \\citep{2009A&A...507L..37T}, which makes it an excellent HVS candidate. More importantly, it was inferred to have originated in the outer Galactic rim, nowhere near the Galactic centre, which would be the favoured place of origin if a supermassive black hole acts as a slingshot as suggested by \\citet[][see also \\cite{2009ApJ...690.1639B}]{1988Natur.331..687H}. Amongst the stars of negative GRF velocity, we discovered that J1539+0239 is a BHB travelling with the largest space velocity of any BHB star known so far, which allowed us to place a lower limit on the mass of the Galactic halo \\citep{2009_J1539}. Here we report that SDSSJ130005.62+042201.6 (J1300+0422 for short) is a metal-poor blue straggler of Population II on a wide retrograde halo orbit that is only marginally bound to the Galaxy in a standard Galactic potential. ", "conclusions": "\\label{sec:conclusion} We have presented a quantitative spectral analysis of a high-velocity star from the sample of faint blue stars in the halo of \\citet{2008ApJ...684.1143X}. Its radial velocity, proper motion, and spectroscopic distance were derived and a detailed kinematical analysis was performed using the Galactic potential of \\citet{1991RMxAA..22..255A} as well as a potential modified one that assumes a more massive dark matter halo. The metal-poor A-type star J1300+0422 was identified as a blue straggler of 1.15~M$_\\odot$\\ due to its main-sequence gravity. A detailed NLTE analysis was performed, which we compared to the standard LTE approach. Significant differences were found especially for \\ion{C}{i} and \\ion{O}{i}. With its low metallicity of $[Fe/H]=-1.2$ and characteristic enhancement of $\\alpha$-elements, it would fit perfectly into the sample of \\cite{2000AJ....120.1014P}, apart from the huge space velocity of the star. The kinematic characteristics ($U$, $V$, $e$, $J_{\\rm Z}$) confirm the halo membership of J1300+0422 beyond any doubt. In addition, its trajectory continues far out into the halo. Many blue stragglers were found to be long period binaries \\citep[with periods of several $100$\\ to $1000$~days;][]{2005AJ....129.1886C,2009Natur.462..1132N} with low radial velocity semi-amplitudes ($K\\sim 5-10$~\\kms, \\citealp{2001AJ....122.3419C,2005AJ....129.1886C} ). Whether or not J1300+0422 is such a binary needs to be verified by an extensive radial velocity study. Our kinematical result is limited by the errors in both the spectroscopic distance and the proper motion. ESA's upcoming astrometry mission GAIA will improve the situation because it will provide a parallax measurement with which to check the spectroscopic distance and improve the proper motion of J1300. GAIA will also have an enormous impact on research into blue stragglers in a more general sense as it will provide astrometry of thousands of halo BSS." }, "1005/1005.0242_arXiv.txt": { "abstract": "While usually cosmological initial conditions are assumed to be Gaussian, inflationary theories can predict a certain amount of primordial non-Gaussianity which can have an impact on the statistical properties of the lensing observables. In order to evaluate this effect, we build a large set of realistic maps of different lensing quantities starting from light-cones extracted from large dark-matter only N-body simulations with initial conditions corresponding to different levels of primordial local non-Gaussianity strength $f_{\\rm NL}$. Considering various statistical quantities (PDF, power spectrum, shear in aperture, skewness and bispectrum) we find that the effect produced by the presence of primordial non-Gaussianity is relatively small, being of the order of few per cent for values of $|f_{\\rm NL}|$ compatible with the present CMB constraints and reaching at most 15-20 per cent for the most extreme cases with $|f_{\\rm NL}|=1000$. We also discuss the degeneracy of this effect with the uncertainties due to the power spectrum normalization $\\sigma_8$, finding that an error in the determination of $\\sigma_8$ of about 3 per cent gives differences comparable with non-Gaussian models having $f_{\\rm NL}=\\pm 1000$. These results suggest that the possible presence of an amount of primordial non-Gaussianity corresponding to $|f_{\\rm NL}|=100$ is not hampering a robust determination of the main cosmological parameters in present and future weak lensing surveys, while a positive detection of deviations from the Gaussian hypothesis is possible only breaking the degeneracy with other cosmological parameters and using data from deep surveys covering a large fraction of the sky. ", "introduction": "In recent years, the interest for an accurate measurement of the amount of non-Gaussianity present in the primordial density field has largely increased. The main reason is that this test is now considered not only a general probe of the inflationary paradigm, but also a powerful tool to constrain the plethora of its different variants. Only the most standard slow-rolling models based on a single field produce in fact almost uncorrelated fluctuations, which is the motivation of the common assumption (and large simplification) that their distribution is Gaussian. In general, small deviations from Gaussianity are predicted even for the simplest inflationary models, while non-standard models, like the scenarios based on the curvaton, on the inhomogeneous reheating and on the Dirac-Born-Infeld inflation allow much more significant departures \\citep[see][and references therein]{Bartolo2004}. It has become common to quantify the level of primordial non-Gaussianity adopting the dimensionless non-linearity parameter $f_{\\rm NL}$ \\citep[see, e.g.,][]{Salopek1990,Gangui1994,Verde2000,Komatsu2001}, that measures the importance of the quadratic term in a sort of Taylor expansion of the gauge-invariant Bardeen potential $\\Phi$\\footnote{We recall that on scales smaller than the Hubble radius $\\Phi$ corresponds to the usual Newtonian peculiar potential (but with changed sign).}: \\begin{equation} \\Phi=\\Phi_{\\rm L}+f_{\\rm NL} (\\Phi^2_{\\rm L}-\\langle \\Phi^2_{\\rm L}\\rangle)\\ ; \\end{equation} here $\\Phi_{\\rm L}$ represents a Gaussian random field. Hereafter we will adopt the so-called large-scale structure (LSS) convention, where $\\Phi$ is linearly extrapolated to the present epoch\\footnote{With the cosmological parameters adopted in this paper, this corresponds to values for $f_{\\rm NL}$ larger by a factor of $\\approx 1.3$ with respect to the so-called CMB convention, where $\\Phi$ is instead extrapolated at $z=\\infty$.}. Moreover we will consider only the so-called local shape for non-Gaussianity, in which the bispectrum signal is larger on squeezed triangle configurations. For more details about other possible shapes we refer to \\cite{Bartolo2004,Verde2009,Bartolo2009.1} and references therein. At present the most stringent constraints on $f_{\\rm NL}$ come from the cosmic microwave background (CMB) data. Their discriminating power derives from the fact that its temperature fluctuations trace the density perturbations before the gravitational non-linearities modify their original distribution. Whatever is the specific test adopted, all CMB analysis consistently allow only very small deviations from Gaussianity: for instance, analyzing the recent WMAP data, \\cite{Komatsu2010} found that $f_{\\rm NL}$ varies between -13 and 97, while \\cite{Smith2009} found $-530$ (approximately 20000 deg$^2$). The possibility of using the weak lensing signals to constrain also the amount of primordial non-Gaussianity has been already explored by different authors. \\cite{Amara2004} used a generalized halo model to study the impact on the estimates of the power spectrum normalization $\\sigma_8$ of primordial non-Gaussianity, modeled assuming various lognormal distributions for the density field. More recently, \\cite{Fedeli2010} computed the power spectrum of the weak cosmic shear for non-Gaussian models with different values of $f_{\\rm NL}$. In particular, they improved the halo model including more accurate prescriptions for its ingredients (mass functions, bias and halo profile), calibrated on the last generation of non-Gaussian N-body simulations. The application of this model to a survey having the expected characteristics of the EUCLID project showed the possibility of a significant detection of non-Gaussianity at the level of $|f_{\\rm NL}|\\approx$ few tens, once the remaining parameters are held fixed. In this paper we investigate weak lensing statistics in non-Gaussian scenarios using numerical rather than analytical tools. Specifically, we will create weak lensing maps performing ray-tracing simulations through very deep light-cones extracted from high-resolution N-body simulations. The advantage of this approach is twofold. First of all, N-body simulations permit to fully account for the non-linear evolution which is usually modeled less accurately by analytical means. Second, numerical experiments allow us to extract a large set of realistic weak lensing maps that can be used for better evaluating the statistical robustness of the results. The main goal of our numerical work is to figure out what are the observational evidences of the presence of some level of primordial non-Gaussianity, as quantified by the $f_{\\rm NL}$ parameter. In particular we will compute a large set of weak lensing statistics in models with different $f_{\\rm NL}$ and we will quantify the deviations from the corresponding results in the Gaussian scenario. This is important not only to address the possibility of a positive detection with future data, but also to establish at which level an amount of primordial non-Gaussianity compatible with the present observational constraints can hamper an accurate measurement of the other cosmological parameters. We must recall that Gaussian initial conditions are virtually always assumed in their practical derivation. The plan of this paper is as follows. In Section~\\ref{sect:lensing} we review the basis of the lensing formalism necessary to the present work. In Section~\\ref{sect:simulations} we describe the cosmological simulations and the numerical procedure to build the lensing maps. In Section~\\ref{sect:results} we present our main results about the statistical properties of the different lensing quantities investigated: the probability distribution function, the third-order moment, the power spectrum and the bispectrum. In Section~\\ref{sect:sigma8} we compare the effects produced by primordial non-Gaussianity to the uncertainties related to power spectrum normalization $\\sigma_8$. Finally, in Sect.~\\ref{sect:conclusions} we draw our conclusions. ", "conclusions": "\\label{sect:conclusions} In this work, we used the outputs of N-body simulations to create a large set of realistic mock maps for several lensing quantities (deflection angle, effective convergence, shear and the two components of the flexion) in the framework of cosmological models with different amount of primordial non-Gaussianity, quantified using the dimensionless parameter $f_{\\rm NL}$. In particular we considered several statistical properties (PDF, power spectrum, shear in aperture, skewness and bispectrum) and compared the results with the corresponding ones obtained in the Gaussian case. Our main results can be summarized as follows. \\begin{itemize} \\item For all quantities here considered the effect produced by the presence of primordial non-Gaussianity is relatively small, amounting to differences of few per cent for $|f_{\\rm NL}|=100$, 5-10 per cent for $|f_{\\rm NL}|=500$, and 15-20 per cent for $|f_{\\rm NL}|=1000$. These results are in good agreement with the analytic predictions presented in \\cite{Fedeli2010}. \\item The largest effects are visible on small scales (i.e. for large multipoles $l>1000$), where, however, also non-linearity can produce strong effects which have to be accurately modeled. \\item The most promising statistical tests to search for imprints of primordial non-Gaussianity are the (convergence and shear) power spectra and the (convergence) bispectrum, thanks to the smaller size of their error bars at the relevant scales. \\item The differences of the various PDFs in both rare-event tails are also important, but their discriminating power is reduced by the poor statistics and by the high-level of noise. \\item We compared the effects produced by the primordial non-Gaussianity with the uncertainties due to the power spectrum normalization: an error in the determination of $\\sigma_8$ of about 3 per cent gives differences comparable with the non-Gaussian models with $f_{\\rm NL}=\\pm 1000$, while for more realistic non-Gaussian models with $f_{\\rm NL}=\\pm 100$ the effects is even larger than the one induced by $\\Delta \\sigma_8=0.01$. \\end{itemize} As said, a significant covariance exists between primordial non-Gaussianity and fundamental cosmological parameters, especially with $\\sigma_8$. The smallness of non-Gaussian effects found in our analysis means that one can obtain a precise estimate of these parameters despite their covariance with $f_{\\rm NL}$. As an example, ignoring the possible presence of a primordial non-Gaussianity with $f_{\\rm NL}=100$, consistent with current observational constraints, would induce a mere 0.2 per cent uncertainty in the estimate of $\\sigma_8$, i.e. well below the current 1-$\\sigma$ error. The same argument, in reverse, is telling us that the search for non-Gaussian features in the weak-lensing statistics is a very challenging task. According to our results, it can be successfully completed only if errorbars can be significantly reduced. For the purpose of discriminating among competing models the reduction needs not to be dramatic. Indeed, as pointed out by \\cite{Fedeli2010}, the fact that deviations from Gaussianity are small but systematics allows one to estimate $f_{\\rm NL}$ by adding the statistical information from different angular bins. A more significant reduction in the errorbars is required to break the degeneracy between non-Gaussianity with cosmological parameters, say $\\sigma_8$. Indeed, as shown in Fig.~\\ref{fig:sigma8}, removing such degeneracy requires to compare information from a limited number of bins at very different scales. As shown by \\cite{Fedeli2010}, both tasks could be achieved with next-generation, all-sky surveys. A good example is certainly represented by the ESA mission EUCLID \\citep{Laureijs2009}. According to the analysis made by \\cite{Fedeli2010} and based on analytic predictions for the power spectrum quantitatively confirmed in this numerical work, we expect that the quality and the quantity of the EUCLID data will allow to constrain $f_{\\rm NL}$ at the level of few tens, opening the possibility of discriminating between the various inflationary models." }, "1005/1005.2592_arXiv.txt": { "abstract": "{We demonstrate the unique capabilities of {\\it Herschel} to study very young luminous extragalactic young stellar objects (YSOs) by analyzing a central strip of the Large Magellanic Cloud obtained through the HERITAGE Science Demonstration Program. We combine PACS 100 and 160, and SPIRE 250, 350, and 500 $\\mu$m photometry with 2MASS (1.25-2.17 $\\mu$m) and {\\it Spitzer} IRAC and MIPS (3.6-70 $\\mu$m) to construct complete spectral energy distributions (SEDs) of compact sources. From these, we identify 207 candidate embedded YSOs in the observed region, $\\sim$40\\% never-before identified. We discuss their position in far-infrared color-magnitude space, comparing with previously studied, spectroscopically confirmed YSOs and maser emission. All have red colors indicating massive cool envelopes and great youth. We analyze four example YSOs, determining their physical properties by fitting their SEDs with radiative transfer models. Fitting full SEDs including the {\\it Herschel} data requires us to increase the size and mass of envelopes included in the models. This implies higher accretion rates ($\\gtrsim$10$^{-4}$M$_\\odot$yr$^{-1}$), in agreement with previous outflow studies of high-mass protostars. Our results show that {\\it Herschel} provides reliable longwave SEDs of large samples of high-mass YSOs; discovers the youngest YSOs whose SEDs peak in {\\it Herschel} bands; and constrains the physical properties and evolutionary stages of YSOs more precisely than was previously possible.} ", "introduction": "The proximity of the Magellanic Clouds offers a unique opportunity to analyze the complete inventory of luminous YSOs over an entire galaxy. With known YSO distances, luminosities, masses, and mass accretion rates can all be well-defined. Comparison of the properties of YSOs in the Magellanic Clouds and in the Milky Way can reveal differences in star formation physics due to metallicity and environment. Using the {\\it Spitzer} SAGE (\\textquotedblleft Surveying the Agents of Galaxy Evolution\\textquotedblright) Survey of the Large Magellanic Cloud \\citep[LMC;][]{meixner06}, Whitney et al. (2008; W08) and Gruendl \\& Chu (2009; GC09) discovered $\\sim$1800 massive YSO candidates in the LMC (a 90-fold increase over previous work). {\\it Spitzer} studies selected sources using colors and SEDs at wavelengths $\\leq$24$\\;\\mu$m (where {\\it Spitzer} can resolve individual YSOs), requiring a detection at 4.5$\\;\\mu$m or shorter in most cases. These surveys thus missed the youngest, most embedded YSOs that can only be detected at longer wavelengths. The {\\it Herschel} Space Observatory \\citep{pilbratt10} has the spatial resolution required to study individual sources at $\\lambda\\gtrsim$50 $\\mu$m (from $\\sim$1.3 pc at 70 $\\mu$m to $\\sim$8.7 pc at 500 $\\mu$m for a distance of 50 kpc, \\citealt{schaefer2008}). The least-evolved massive protostars are characterized by cold dust temperatures probed at far-infrared (far-IR) wavelengths, and are expected to be $\\sim$10$^3$ times brighter at 100 $\\mu$m than at 5 $\\mu$m \\citep{whitney04,molinari08}, making {\\it Herschel} extremely effective at detecting those youngest YSOs. With {\\it Herschel}, we not only discover new objects but also better characterize {\\it Spitzer}-identified YSO candidates. The {\\it Herschel} data constrain the models of these sources and improve estimates of such physical parameters as total luminosity, stellar mass, and total dust mass. We demonstrate these capabilities by studying a strip across the LMC observed as part of the Science Demonstration Program (SDP) -- the first part of the {\\it Herschel} Key Program \\textquotedblleft HERschel Inventory of the Agents of Galaxy Evolution\\textquotedblright\\ (HERITAGE; \\citealt{meixner10}) in the Magellanic Clouds. The strip was mapped in the PACS 100 and 160$\\;\\mu$m bands \\citep{poglitsch10} and SPIRE 250, 350, and 500$\\;\\mu$m bands \\citep{griffin10}. ", "conclusions": "We show that {\\it Herschel} far-IR photometry is very effective in identifying YSOs in the LMC. Adding {\\it Herschel} data to existing {\\it Spitzer} and near-IR observations results in significantly improved analysis of YSOs, as summarized in these key findings: \\begin{list}{$\\circ$}{\\setlength{\\topsep}{0in}\\setlength{\\parsep}{0in}} \\item Nearly all sources detected by {\\it Herschel} are Stage~0/I, very young, with a high ratio of circumstellar to stellar mass. \\item Previously studied warm sources such as YSO-1 require more circumstellar dust to fit the {\\it Herschel} data, implying a less evolved state than would be inferred from {\\it Spitzer} data alone. \\item {\\it Herschel} photometry significantly constrains our SED fits, decreasing the range of circumstellar dust masses and evolutionary states consistent with the measurements. \\item Many sources require even more cold circumstellar dust than is present in our original model grid, motivating improvements to our models. \\end{list} Our observations prove that {\\it Herschel} can offer us, for the first time, an inventory of the earliest stages of protostellar formation throughout an entire galaxy. \\\\" }, "1005/1005.2840_arXiv.txt": { "abstract": "{% V391~Peg (HS~2201+2610) is an extreme horizontal branch subdwarf B (sdB) star, it is an hybrid pulsator showing $p$- and $g$-mode oscillations, and hosts a 3.2/sin$i$ M$_{Jup}$ planet at an orbital distance of about 1.7 AU. In order to improve the characterization of the star, we measured the pulsation amplitudes in the {\\it u'g'r'} SLOAN photometric bands using ULTRACAM at the William Herschel 4.2 m telescope and we compared them with theoretical values. The preliminary results presented in this article conclusively show that the two main pulsation periods at 349.5 and 354.1~s are a radial and a dipole mode respectively. This is the first time that the degree index of multiple modes has been uniquely identified for an sdB star as faint as V391~Peg (B=14.4), proving that multicolor photometry is definitely an efficient technique to constrain mode identification, provided that the data have a high enough quality.} ", "introduction": "About half of field sdB stars, which reside in binary systems, can form through common envelope ejection or stable Roche lobe overflow (Han et al. 2002, 2003). It is more difficult to explain the formation of a single sdB star. Two scenarios are possible: the merger of two low-mass helium white dwarfs and an early hot helium flash; but both are not fully consistent with the observations. A recent review on these arguments is given by Heber (2009). Another possibility, suggested by Soker (1998), is that the huge mass loss needed to form an sdB star is triggered by low-mass bodies, planets or brown dwarfs (BDs). Although this possibility has not been tested by detailed models yet, a planet to the pulsating sdB star V391~Peg (Silvotti et al. 2007) and three circumbinary planets/BDs to the eclipsing sdB+M systems HW~Vir (Lee et al. 2009) and HS~0705+6700 (Qian et al. 2009) suggest that sdB planets/BDs could be a relatively common phenomenon (see also Bear \\& Soker 2010). A systematic search for sdB substellar objects around 4 sdB stars using the timing method is the main goal of the EXOTIME project (Schuh et al. 2010, Benatti et al. 2010). \\vspace{1mm} V391~Peg (HS~2201+2610) is a particularly interesting system formed by an sdB star and a 3.2/sin$i$ M$_{Jup}$ planet orbiting the host star in 3.2 years at a distance of about 1.7 AU (Silvotti et al. 2007). The sdB star is a hybrid pulsator showing $p$ and $g$-mode oscillations at the same time ({\\O}stensen et al. 2001, Lutz et al. 2009), offering a unique opportunity to characterize the host star through asteroseismic methods. A preliminary mode identification of the higher pulsation frequencies ($p$-modes) was proposed in Silvotti et al. (2002): the two main pulsation periods of 349.5 and 354.1 s could be reproduced with $l$=0 ($k$=1) and $l$=1 ($k$=1) respectively. However this solution was not unique due to the small number of detected frequencies and other solutions could not be excluded. \\vspace{13.6mm} ", "conclusions": "Thanks to the high quality of the data, this is the first time that the mode degree index has been uniquely identified from multicolor photometry for the two main modes of a star as faint as V391~Peg (V=14.6). To our knowledge, conclusive results were obtained only for two brighter stars: KPD~2109+4401 (V=13.4, Randall et al. 2005, see also Jeffery et al. 2004) and Balloon 090100001, the brightest known sdBV star with V=12.1 (Baran et al. 2008, Charpinet et al. 2008). The results reported in this article confirm that multicolor photometry can set useful identification constraints on the pulsation modes of sdB rapid pulsators, provided that the data have a high enough quality. ULTRACAM on a 4~m class (or larger) telescope is an ideal instrument for such studies." }, "1005/1005.3594_arXiv.txt": { "abstract": "{A phenomenological relationship between oscillations in a sunspot and quasi-periodic pulsations in flaring energy releases at an active region above the sunspot, is established. The analysis of the microwave emission recorded by the Nobeyama Radioheliograph at 17~GHz shows a gradual increase in the power of the 3-min oscillation train in the sunspot associated with AR 10756 before flares in this active region. The flaring light curves are found to be bursty with a period of 3 min. Our analysis of the spatial distribution of the 3-min oscillation power implies that the oscillations follow from sunspots along coronal loops towards the flaring site. It is proposed that quasi-periodic pulsations in the flaring energy releases can be triggered by 3-min slow magnetoacoustic waves leaking from sunspots. } ", "introduction": "The problem of the transfer of energy, momentum, and information from sub-photospheric solar regions to both the corona and the solar wind is one of the most difficult in solar and stellar physics. Magnetohydrodynamic (MHD) waves are believed to play a key role because the waves are natural carriers of energy, momentum, and information (e.g., Erd{\\'e}lyi 2006 for a recent review). In addition, the guided nature of the wave propagation opens up very interesting perspectives for fixing and tracing the energy transfer channels highlighted by the waves. In general, MHD waves can be guided by inhomogeneities in the characteristic MHD speeds (Alfv\\'en, fast and slow) as well as the magnetic field itself, which are found in all regions of the solar atmosphere. At the chromospheric level, one of the most pronounced wave phenomena are the 3-min oscillations over sunspots (see, e.g. Bogdan \\& Judge 2006 for a recent review), usually detected as intensity oscillations in visible light, UV, and EUV spectral lines, as well as in microwave band (e.g., Shibasaki 2001, Gelfreikh et al. 1999, Nindos et al. 2002) and in the dm-radio flux records (M\\'esz\\'arosov\\'a et al. 2006). Three-minute oscillations in sunspots are believed to be associated with slow magnetoacoustic waves (e.g., Zhugzhda 2008). Outwardly propagating compressible waves of the same periodicity are also seen in both the EUV 171\\AA\\ and 195\\AA\\ bandpasses in the magnetic fan structures situated over sunspots (e.g., De Moortel 2006 for a review). The projected phase speed of these waves is subsonic, and the waves are seen to propagate along the plasma channels elongated along the coronal magnetic field lines, and hence are interpreted as slow magnetoacoustic waves. Compressible 3-min waves observed at the same location in both EUV bandpasses show a high degree of correlation (King et al. 2003). The relationship between these waves and 3-min oscillations in sunspots remains unclear. The understanding of the propagation of 3-min oscillations through the solar atmosphere is one of the most important problems of solar physics. Its understanding will probably indicate the nature and properties of the plasma channels that transfer these waves into the corona, and hence the connectivity of different layers of the atmosphere. The role played by 3-min oscillations in the corona is also of interest, in particular, the relationship between 3-min oscillations in sunspots and the flaring activity in the active regions (AR) above the sunspots. A possible indication of such a relationship was mentioned in Gelfreikh (2002). Wave and oscillatory phenomena in various parts of the solar atmosphere can trigger and modulate bursty energy releases, e.g., solar flares. In this case, the periodicity of the oscillations will be evident in the flaring light curves as quasi-periodic pulsations (QPP). This can be achieved by several mechanisms. In the scenario proposed by Nakariakov et al. (2006), energy of transverse (kink or sausage) oscillations of coronal loops can periodically leak to a magnetic neutral point or line situated nearby. The incoming fast magnetoacoustic wave refracts towards the neutral point, experiencing focussing and steepening. This periodically generates very sharp spikes of electric current density in the vicinity of the neutral point, which in turn can be affected by current driven plasma micro-instabilities. The instabilities can cause the onset of micro-turbulence and hence enhance the plasma resistivity by several orders of magnitude. This would lead to periodic triggering of magnetic reconnection and hence the manifestation of the loop oscillations as periodic variation in the flaring light curve. A compressible wave can periodically trigger magnetic reconnection not only by periodic current density spikes, but also by the variation in the plasma density in the vicinity of the reconnection site. This possibility was modelled numerically by \\cite{2006SoPh..238..313C} in interpreting 3--5~min periodicity detected in repetitive bursts of explosive events in the transition region (Ning et al. 2004). Density variations result in a periodic variation in the electron drift speed. Depending upon the ratio of electron to proton temperatures, the value of the speed controls the onset of the Buneman or ion-acoustic instabilities and hence anomalous resistivity. The periodic onset of the anomalous resistivity triggers periodic energy releases. Transverse compressible waves may also directly trigger magnetic reconnection causing transition region explosive events, by changing the magnetic field strength (Doyle et al. 2006). Longitudinal, e.g., acoustic waves can also modulate flaring energy releases, either directly by the modulation of the drift velocity or the modulation of gyrosynchrotron emission efficiency (Nakariakov \\& Melnikov 2006), or indirectly e.g., by means of centrifugal conversion into fast magnetoacoustic waves on the curved magnetic field lines (Zaitsev \\& Stepanov 1989). Kislyakov et al. (2006) analysed 15 flares observed in the 37~GHz band with the Mets\\\"ahovi radio telescope (Finland) with the use of the \\lq\\lq sliding window\" Fourier transform and the Wigner--Ville nonlinear transform. The telescope spatial resolution is 2.4~arc min, the sensitivity is about 0.1 sfu, and the time resolution was higher than 0.1~s. During 13 events (about 90\\%), a 5-min periodic modulation of the emission intensity was detected with the frequency of $3.2\\pm 0.37$~mHz. In addition, a shorter period (about 1~s) signal was detected, which was found to be frequency modulated with the same 5-min period. In the development of this study (Za{\\u i}tsev \\& Kislyakov 2006), simultaneous modulation of the microwave emission by three low frequency signals with periods of 3.3, 5, and 10~min was observed in 30\\% of the analysed outbursts. It was suggested that the detected modulation was caused by the parametric resonance between 5-min velocity oscillations in the solar photosphere and natural acoustic oscillations of coronal magnetic loops modulating the microwave emission. The detected periods of 5, 10, and 3 min were interpreted to correspond to the pumping frequency, its subharmonic, and its first upper frequency of parametric resonance, respectively. Confirmation of this finding with a different instrument is required. The physical mechanisms responsible for the appearance of 3 and 5 min flaring QPP cannot be determined without spatial information. In general, spatial information, such as the spatial size and shape of the region occupied by an oscillation, and the distribution of the oscillation power, phase, and spectrum over the source, is crucial to establishing the nature of the QPP (e.g., Grechnev et al. 2003; Melnikov et al. 2005). Novel imaging data analysis techniques developed in solar physics have been shown to allow one to exploit the full potential of the spatially resolving observations. Grechnev (2003) proposed creating a 2D broadband variance map, representing the overall dynamics of an analysed event, from microwave correlation data cubes obtained with NoRH. This approach allows us to ascertain the spatial locations of faint variable microwave emission sources. Nakariakov \\& King (2007) designed a coronal-periodmapping technique, which reduces 3D imaging data cubes (2D in space and time) to a sequence of static maps inferring collective oscillations of extended (larger than the pixel) coronal structures. This approach was successfully applied by Inglis et al. (2008) to the study of single-periodic pulsations in a large off-limb flaring loop seen in the microwaves with NoRH. However, period-mapping does not provide any phase information, and does not allow detailed studying of multi-periodic or non-stationary phenomena. Complex temporal and spatial features of oscillatory processes in imaging datasets can be studied with the pixelised wavelet filtering (PWF) technique (Sych \\& Nakariakov 2008). This approach produces 4D (2D in space, time and frequency) data cubes providing information about the time modulation of oscillatory signals, their coupling, and their evolution. In particular, this method allows one to obtain information about the spatial structure of narrowband and broadband time signals, as well as the analysis of the signal integrated over the whole field-of-interest. The practical implementation of PWF consists of several steps: \\begin{itemize} \\item The images in the time sequence are coaligned, removing the spatial mismatch between consecutive images. \\item The object of investigation (e.g., the period of the oscillations) and the field-of-interest (FOI) are selected. \\item Construction of the variance map of the FOI, determining the spatial distribution of the integrated power of the time signal. \\item Direct wavelet transform in the time domain, filtering out the spectral components of interest, and inverse wavelet transform. The resultant data cube contains the time variation in the signal in the prescribed spectral band. \\end{itemize} In addition, global wavelet spectra can be calculated for each pixel, resembling the construction of a periodmap. PWF is a convenient tool for establishing a relationship between different spatially-separated oscillatory processes. Analysis of the possible relationship between flaring energy releases and dynamical processes in the lower regions of the solar atmosphere can shed light on the triggering of the energy releases and hence the basic physical processes responsible for them. The spatially-resolved analysis of this relationship can also infer the atmospheric connectivity channels. The aim of the paper is to establish a phenomenological relationship between dynamical processes occurring in a sunspot and in flares, within the active region (AR) linked to the sunspot, and identify the channels of the connectivity. We demonstrate that 3-min oscillations of a sunspot appear to be present in the microwave emission associated with the flaring activity over the sunspot. ", "conclusions": "The microwave light curves of solar flares on 2005 April 28 and 2005 May 4 contain pronounced variations with periods of about 3 min. This behaviour indicates that there is an apparent relationship with 3-min oscillations in the sunspot situated close to the flare sites. The aim of this paper was to understand this relationship. Our analysis of dynamical features in the microwave, EUV, white light, and X-ray imaging data of AR~10756 acquired during its passage through the solar disk from 2005 April 28 to 2005 May 4, inferred the dynamical morphology of the active region. The 3-min narrowband signals detected over the sunspot and in the flare site are all well localised, which excludes their possible link with the instrumental artifacts, such as sidelobes of the image synthesis, and hence are natural. The 3-min narrowband maps of the active region, constructed with the use of PWF show the presence of extended V-shaped sources situated over the sunspot, with arms extended towards the flare site. We interpret these arms as evidence of the magnetic plasma channels that link the sunspot and the flare site by guiding magnetohydrodynamic waves. The 3-min periodicities of energy releases are then triggered by the 3-min oscillations leaking out from the sunspot along the magnetic structures. On the basis of our findings, we deduce that the physical mechanism responsible for the relationship between 3-min sunspot oscillations and 3-min QPP in nearby flares can be as follows. The energy of 3-min oscillations leaks out of the sunspots in the form of field-aligned slow magnetoacoustic waves, which are often seen as compressible variations in the EUV radiation in the magnetic fan structures over sunspots (e.g., De Moortel 2006). In our study, these waves are seen as a 3-min modulation of the microwave radiation. The spatial distribution of the emission highlights the waveguiding plasma structures, which are the V-shaped microwave sources found in both flares discussed above. Because of the curvature of the magnetic field lines, the centrifugal force associated with the periodic longitudinal field-aligned wave motions produce periodic transverse kink-like perturbations of the magnetic structures (Zaitsev \\& Stepanov 1989). The induced transverse motions are fast magnetoacoustic waves that can carry energy and information across the magnetic field. These kink waves can trigger flaring energy releases (e.g., by the mechanism proposed by Nakariakov et al. 2006) provided that the waveguiding channel is situated close to the magnetic null-point, while is not necessarily linked magnetically. The modulation depth of the flaring light curves can be significantly stronger than in the modulating signal. Hence, the leaking 3-min wave can either trigger the energy release in a form of a short aperiodic spike or spike-like pulse, as seen on 2005 April 30, or lead to periodic triggering (or modulation) of energy releases in longer duration bursts, as seen on 2005 April 28 and 2005 May 4. In the first case, the triggered energy release uses up the energy stored in the magnetic configuration, and the next maximum in the wave cannot trigger another energy release. In the second case, either not all stored energy is liberated in the previous releases, or the next period of the triggering wave causes the energy release at another spatial location (see the discussion in Nakariakov et al. 2006). In both analysed bursts, there is observational evidence that before the flares the energy of 3-min oscillations in the sunspot is enhanced significantly. In both cases, the amplitude of 3-min wave trains was highest just before the onset of the burst. This provides interesting perspectives on the use of the increase in the power of 3-min oscillations just before the flare as a flare precursor. However, the relationship between the amplitude of 3-min oscillations in the sunspot and energy releases nearby, requires statistical proof, and should be subject to a dedicated study. A possible interpretation of the preferential appearance of the spike-like events when the analysed active region is in the vicinity of the central meridian, mentioned in Sect.~\\ref{sec:obs}, can be interpreted as either just the time coincidence with the appearance of certain physical conditions for the energy releases (e.g., the emergence of a new loop or arcade of loops from the sunspot to the burst source) or the preferential observability conditions (e.g., connected with the line-of-sight angle). The latter issue could be indicative of the similarity between the observed pulses and well-known microwave spikes: Altyntsev et al. (1996) found out that the spatial size of the microwave spike source is systematically larger in the vicinity of the limb. This was interpreted in terms of the scattering of the microwave emission across the coronal plasma. Hence, some events situated close the limb can be of lower intensity due to the scattering and hence be below the detection threshold and missed by observations. This may explain the preferential appearance of the short pulses in the vicinity of the central meridian." }, "1005/1005.3077_arXiv.txt": { "abstract": "We conducted the observational tests of a phase correction scheme for the Atacama Compact Array (ACA) of the Atacama Large Millimeter and submillimeter Array (ALMA) using the Submillimeter Array (SMA). Interferometers at millimeter- and submillimeter-wave are highly affected by the refraction induced by water vapor in the troposphere, which results as phase fluctuations. The ACA is planning to compensate the atmospheric phase fluctuations using the phase information of the outermost antennas with interpolating to the inner antennas by creating a phase screen. The interpolation and extrapolation phase correction schemes using phase screens are tested with the SMA to study how effective these schemes are. We produce a plane of a wavefront (phase screen) from the phase information of three antennas for each integration, and this phase screen is used for the interpolation and extrapolation of the phases of inner and outer antennas, respectively. The interpolation scheme obtains apparently improved results, suggesting that the ACA phase correction scheme will work well. On the other hand, the extrapolation scheme often does not improve the results. After the extrapolation, unexpectedly large phase fluctuations show up to the antennas at the distance of $\\sim140$~m away from the center of the three reference antennas. These direction vectors are almost perpendicular to the wind direction, suggesting that the phase fluctuations can be well explained by the frozen phase screen. ", "introduction": "\\label{sect-intro} The Atacama Large Millimeter and submillimeter Array (ALMA), the largest millimeter and submillimeter interferometer ever built, is currently under construction in the northern Chile with the collaboration between East Asia, Europe, and North America \\citep{woo09}. The ALMA is composed of up to eighty high-precision antennas, located on the Chajnantor plain of the Chilean Andes in the District of San Pedro de Atacama, 5000~m above sea level. The ALMA covers the wavelength range from 0.3~mm to 9~mm with an angular resolution of up to \\timeform{0''.004}. The Atacama Compact Array (ACA) is designed to improve the short baseline coverage of the ALMA, especially for the observations of extended structures at submillimeter wavelength \\citep{igu09}. The ACA consists of four 12~m telescopes to obtain the single-dish data and twelve 7~m telescopes to obtain short baseline interferometric data. However, ground-base astronomical observations in the millimeter and submillimeter ranges are strongly affected by the fluctuation of the tropospheric water vapor distribution (e.g., \\cite{tho01}). Therefore, the correction of phase fluctuations due to the spatial and temporal variations of the tropospheric water vapor content is extremely important in millimeter and submillimeter interferometry. Several kinds of techniques have been proposed and performed for reducing phase fluctuations in millimeter and submillimeter interferometry (see \\cite{car99} for a review), such as fast switching phase calibration \\citep{hol92,hol95a,car97,mor00}, paired array phase calibration \\citep{hol92,car96,asa96,asa98}, and radiometric phase calibration \\citep{lay97,car98,car99,mar98,del00,wie01,mat02}. Although the ALMA site is one of the best sites for the millimeter and submillimeter astronomy \\citep{mat98,mat99,pai00,mat03}, atmosphere in this area is often affected by the phase fluctuations due to water vapor \\citep{rad96,hol97,but01}. The phase correction methods for the ALMA and the ACA are therefore highly needed to be considered and tested. \\citet{asa05} conducted a series of simulations of a phase correction scheme for the ACA using water vapor radiometers (WVRs). The WVRs can measure the tropospheric water vapor content directly with observing a water vapor line in centimeter or millimeter wavelength. In the proposed ACA phase calibration scheme, the WVRs are attached to the 12~m antennas at the four corners of the twelve 7~m array (hereafter we call this as the reference rectangle). The changes of the tropospheric water vapor content aloft measured with the WVRs is transferred into the excess path lengths of the arriving radio waves. The excess path lengths measured at the four corners of the reference rectangle are then fitted to a simple two-dimensional slope or a screen. Then the phases of the antennas inside the reference rectangle can be compensated and calibrated. Their simulation succeeded to compensate the atmospheric phase well, which strongly support the use of the proposed phase correction scheme. To confirm this simulation study observationally and discuss further, we performed the proposed phase correction scheme for the ACA using the Submillimeter Array (SMA; \\cite{ho04}). Here we present measurements with the SMA at 230~GHz, analyze the datasets under the proposed scheme, and discuss the results of the corrected phase variations. Our experiment is to clarify how effectively the proposed compensation scheme works under the conditions of the real atmosphere. We construct a reference triangle composed of three antennas, and make a flat phase plane or a screen with observing a strong point source (section~\\ref{sect-mdr}). The phases of antennas inside the reference triangle can be interpolated, while the phases of antennas outside can be extrapolated. We then compare the observed and the predicted phases of the point source. Standard deviations and temporal structure functions of the observed (uncorrected) and the corrected phases are also compared (section~\\ref{sect-res}). Finally, we discuss the usefulness of the interpolation and extrapolation phase correction schemes using a phase screen, and also discuss the validity of the ``frozen-flow'' hypothesis (section~\\ref{sect-dis}). ", "conclusions": "We performed an interferometric phase correction with the interpolation or extrapolation of the phase screen defined by three reference antennas using the SMA. This interpolation method is proposed for the ACA in the ALMA. According to the comparisons of the standard deviations of corrected and uncorrected phases, relations between the phase standard deviation and the distance from the center of the reference triangle, and the temporal structure functions of the rms phase, the interpolation scheme improves phase fluctuation while the extrapolation scheme does not. This result can be explained by the boundary conditions of phase in these schemes; in case of the interpolation scheme, the phase corrected antenna is inside the triangle of three reference antennas, so the phase inside the triangle can be well defined (more known boundary conditions, more precisions or less phase errors and deviations). The extrapolation scheme, on the other hand, only has partial boundary conditions, and therefore less precision. In the extrapolation scheme results, there is a sudden large phase fluctuation at the distance from the center of the reference triangle of about 140~m. According to the meteorological parameters on those observing dates and the antenna configurations, this ``140~m phase fluctuation'' is occurring only at the antennas located from the center of the reference triangle perpendicular to the wind direction. This ``140~m phase fluctuation'' can be explained by the frozen-flow model. Although there are some differences between the configuration of our experiments and that in the proposed phase correction scheme for the ACA, our results based on the actual observations and the simulation results done by \\citet{asa05} promise the success of the phase correction for the ACA. \\bigskip We would like to thank Kazushi Sakamoto and Koh-Ichiro Morita for fruitful discussion. 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." }, "1005/1005.1072_arXiv.txt": { "abstract": "{We set out to determine the ratio, \\qir, of rest-frame 8--1000-$\\mu$m flux, \\sir, to monochromatic radio flux, \\s1.4, for galaxies selected at far-infrared (-IR) and radio wavelengths, to search for signs that the ratio evolves with redshift, luminosity or dust temperature, \\td, and to identify any far-IR-bright outliers -- useful laboratories for exploring why the far-IR/radio correlation (FIRRC) is generally so tight when the prevailing theory suggests variations are almost inevitable. We use flux-limited 250-$\\mu$m and 1.4-GHz samples, obtained using \\herschel\\ and the Very Large Array (VLA) in GOODS-North (-N). We determine bolometric IR output using ten bands spanning $\\lambda_{\\rm obs}= 24-1250\\,\\mu$m, exploiting data from PACS and SPIRE (PEP; HerMES), as well as {\\it Spitzer}, SCUBA, AzTEC and MAMBO. We also explore the properties of an \\lir-matched sample, designed to reveal evolution of \\qir\\ with redshift, spanning log \\lir\\ = 11--12\\,L$_{\\odot}$ and $z=0-2$, by stacking into the radio and far-IR images. For 1.4-GHz-selected galaxies in GOODS-N, we see tentative evidence of a break in the flux ratio, \\qir, at \\l1.4\\ $\\sim 10^{22.7}$\\,W\\,Hz$^{-1}$, where active galactic nuclei (AGN) are starting to dominate the radio power density, and of weaker correlations with redshift and \\td. From our 250-$\\mu$m-selected sample we identify a small number of far-IR-bright outliers, and see trends of \\qir\\ with \\l1.4, \\lir, \\td\\ and redshift, noting that some of these are inter-related. For our \\lir-matched sample, there is no evidence that \\qir\\ changes significantly as we move back into the epoch of galaxy formation: we find \\qir\\ $\\propto (1+z)^{\\gamma}$, where $\\gamma=-0.04\\pm 0.03$ at $z=0-2$; however, discounting the least reliable data at $z<0.5$ we find $\\gamma = -0.26\\pm 0.07$, modest evolution which may be related to the radio background seen by ARCADE\\,2, perhaps driven by $<$10-$\\mu$Jy radio activity amongst ordinary star-forming galaxies at $z>1$.} ", "introduction": "For samples of local galaxies -- on galactic and $\\sim$100-pc scales -- there is a good correlation between far-IR and radio emission \\citep{dejong85, helou85, condon91, yun01}. The correlation spans many orders of magnitude in luminosity, gas surface density and photon, cosmic-ray and magnetic energy density, and arises because the far-IR and radio wavelength regimes share a common link with luminous, massive stars and their end products -- dust, supernovae (SNe) and cosmic rays. In the simplest models \\citep[dubbed `calorimetry' -- e.g.][]{volk89, lvx96}, dust absorbs all of the ultraviolet radiation from massive stars, re-radiating this energy in the far-IR, and when those massive stars explode as SNe they generate cosmic-ray electrons which lose all their energy in the radio regime, mainly via synchrotron emission. A balance is thereby achieved between far-IR and radio emission, assuming that the starburst timescale is sufficiently long ($>$10$^7$\\,yr). Traditionally, \\lir\\ and \\lrad\\ are both employed to determine star-formation rates, and the far-IR/radio flux density ratio has been useful when estimating the redshift or \\td\\ of a distant starburst, or when defining samples of AGN \\citep{condon92, cy99, ivison02, bell03, chapman05, donley05}, or probing magnetic field strength \\citep{thompson06}. For these reasons, and because of recent observational advances at both far-IR and radio wavelengths, there has been a deluge of FIRRC-related work recently, exploring why the correlation exists and whether it continues to hold at progressively larger look-back times \\citep{garrett02, appleton04, ibar08, seymour09, ivison10, sargent10}. Prevailing theory \\citep[e.g.][]{lacki10} suggests that variations in the far-IR/radio flux ratio should be virtually unavoidable and that the FIRRC thus arises due to a mysterious combination of effects involving bremsstrahlung, inverse Compton cooling, ionisation and the relative fractions of primary/secondary cosmic-ray electrons/protons, as well as the critical synchrotron frequency. Aside from the modelling work of \\citeauthor{lacki10}, recent advances in this field have included the use of luminosity-matched samples (between high and low redshift) to better probe evolution with look-back time \\citep{sargent10b} and the use of measurements spanning the far-IR and radio wavebands to avoid assumptions relating to $k$ corrections \\citep{ivison10}, although \\citet{calzetti10} have argued that bands beyond 24\\,$\\mu$m contain a contribution from dust heated by stars from previous episodes of star formation and so we might not necessarily expect the correlation to improve. In this paper we introduce flux-limited 250-$\\mu$m- and 1.4-GHz-selected samples of galaxies from \\herschel\\ and the VLA, as well as a luminosity-matched sample selected at 24\\,$\\mu$m, spanning $z=0-2$, and determine their spectral energy distributions (SEDs) spanning the entire far-IR spectral region. We then investigate the FIRRC from the perspectives of the 24-, 250-$\\mu$m- and radio-selected samples. ", "conclusions": "\\begin{figure} \\centerline{\\psfig{figure=14552fg2.ps,width=3.3in,angle=270}} \\caption{\\qir\\ versus \\sir\\ for those 250-$\\mu$m-selected galaxies (sample~1) with secure, unambiguous radio ids. Those without plausible radio ids are plotted as stars. The dashed line is the median, \\qir\\ = 2.40; the shaded region represents $\\pm$2$\\sigma_q$ ($\\sigma_q=0.24$).} \\label{q250} \\end{figure} \\begin{figure} \\psfig{figure=14552fg3.ps,width=3.45in,angle=0} \\caption{\\qir\\ versus redshift for our radio-selected galaxies (sample~2), in five bins of $K$-corrected \\l1.4, plus the full sample. Values of \\qir\\ for sample 3 are shown (as circles) for comparison.} \\label{qlrad} \\end{figure} \\begin{figure} \\psfig{figure=14552fg4.ps,width=3.55in,angle=0} \\caption{Median \\qir\\ versus \\l1.4. The local luminosity functions of starbursts and AGN are shown \\citep{ms07}.} \\label{ql1400} \\end{figure} \\qir\\ as utilised here is the logarithmic ratio of the rest-frame 8--1000-$\\mu$m flux, \\sir, and the 1.4-GHz flux density, $S_{\\rm 1.4GHz}$, such that \\qir\\ = log$_{10}$ [(\\sir/$3.75\\times 10^{12}$\\,W\\,m$^{-2}$)/($S_{\\rm 1.4GHz}$/W\\,m$^{-2}$\\,Hz$^{-1}$)], where $S_{\\rm 1.4GHz}$ is $k$-corrected assuming $S_{\\nu}\\propto \\nu^{\\alpha}$, with $\\alpha=-0.8$. We begin with sample (1), those selected at 250\\,$\\mu$m: \\qir\\ is not a strong function of \\sir\\ (Fig.~\\ref{q250}), nor of \\s1.4. We see no evidence of contamination by radio-loud AGN, consistent with the findings of \\citet{yun01}. Some galaxies stand out as potentially far-IR-bright: these include the three galaxies without plausible radio ids, two of which are detected at 70 and/or 160\\,$\\mu$m, so are likely at low redshift with their radio emission resolved away. Only 39/65 sources with unambiguous radio ids have redshifts (20 photometric, 19 spectroscopic; $\\left\\langle z \\right\\rangle = 0.98$; interquartile $z$ = 0.46--1.52, similar to sample~2). Nevertheless, this sub-sample allows us to explore correlations between \\qir\\ and luminosity, redshift and \\td. We find significant ($>$95\\% confidence -- Table~\\ref{trends}) trends for lower \\qir\\ amongst the most radio- and far-IR-luminous galaxies, and the warmest and most distant, though these parameters are likely inter-related. The dependence of \\qir\\ on \\l1.4\\ is the strongest and likely reflects the influence of low-radio-power AGN, of which more later; that of \\qir\\ on \\lir\\ is more puzzling, perhaps reflecting the dependence of \\lir\\ on redshift and/or \\td\\ \\citep[e.g.][]{chapman05}, or selection effects (since this trend is not seen for sample 2 -- see Table~\\ref{trends}). \\begin{table} \\caption{Trends.} \\label{trends} \\centering \\begin{tabular}{l c c} \\hline\\hline \\qir\\ trend& Spear- &Signifi-\\\\ & man $\\rho$&-cance\\\\ \\hline \\multicolumn{3}{l}{Sample 1 (250-$\\mu$m-selected galaxies with redshifts):}\\\\ (5.02$\\pm$0.18) -- (0.105$\\pm$0.008)\\,log\\,\\l1.4 &$-0.48$&99.8\\%\\\\ (6.09$\\pm$0.33) -- (0.092$\\pm$0.008)\\,log\\,\\lir &$-0.32$&95.6\\%\\\\ (2.61$\\pm$0.02) -- (0.081$\\pm$0.007)\\,(1+$z$) &$-0.33$&96.0\\%\\\\ (2.76$\\pm$0.03) -- (0.008$\\pm$0.001)\\,\\td\\ &$-0.33$&96.1\\%\\\\ \\hline \\multicolumn{3}{l}{Sample 2 (radio-selected galaxies with redshifts):}\\\\ (4.92$\\pm$0.21) -- (0.101$\\pm$0.009)\\,log\\,\\l1.4 &$-0.27$&99.9\\%\\\\ (2.74$\\pm$0.35) -- (0.007$\\pm$0.009)\\,log\\,\\lir &$+0.07$&69.1\\%\\\\ (2.55$\\pm$0.02) -- (0.047$\\pm$0.010)\\,(1+$z$) &$-0.15$&96.6\\%\\\\ (2.60$\\pm$0.02) -- (0.002$\\pm$0.001)\\,\\td\\ &$-0.16$&89.2\\%\\\\ \\hline \\end{tabular} \\end{table} Fig.~\\ref{qlrad} shows \\qir\\ versus redshift for our radio-selected galaxies (sample 2), split into five log-spaced bins of \\l1.4. Does \\qir\\ evolve with redshift? One might conclude that it does, based on the bottom panel of Fig.~\\ref{qlrad}, where \\qir\\ $\\propto (1+z)^{\\gamma}$, with $\\gamma=-0.05\\pm 0.01$ (Table~\\ref{trends}). However, we must be aware of some strong selection effects which make this evidence unreliable: radio emission can be due to an AGN and several radio-loud objects with low values of \\qir\\ are obvious in Fig.~\\ref{qlrad}. Such AGN are more common at $z\\sim 2$ than today \\citep[e.g.][]{wall05}; moreover, radio emission from faint starbursts (with $\\alpha=-0.8$, although see \\citealt{ibar10}) becomes more difficult to detect at higher redshifts, such that the fraction of radio-loud AGN in a flux-limited sample will rise, driving down \\qir. Indeed, Fig.~\\ref{ql1400} shows tentative evidence of a break in $\\left\\langle q_{\\rm IR} \\right\\rangle$ at \\l1.4\\ $\\sim10^{22.7}$\\,W\\,Hz$^{-1}$. One might also expect radio-loud objects (those with low \\qir) to contain warmer, AGN-heated dust, giving rise to the weak trend (89.2\\% confidence -- Table~\\ref{trends}) of decreasing \\qir\\ with increasing \\td. \\begin{figure} \\centerline{\\psfig{figure=14552fg5.ps,width=3.4in,angle=270}} \\caption{\\lir\\ (dots; left axis) and \\qir\\ (red circles; right axis) -- the former determined via the models of \\citet{ce01} -- versus redshift for our \\lir-matched sample. The luminosity bounds and redshift bins (dashed lines), the number of galaxies in each bin and their predicted (measured) $\\left\\langle {\\rm log} L_{\\rm IR}\\right\\rangle$ and measured $\\left\\langle q_{\\rm IR}\\right\\rangle$ are all shown. The shaded area represents a $\\pm 1\\sigma$ prediction for \\qir\\ \\citep{swinbank08, ivison10}.} \\label{lirz} \\end{figure} Finally, we turn to our \\lir-matched galaxies (sample 3), illustrated in Fig.~\\ref{lirz}. The $\\delta z=0.5$ bins provide significant numbers of objects at near-constant \\lir\\ spanning $z=0-2$. As well as being matched in \\lir, there is another key difference between our new sample and that used by \\citet{ivison10}: although the new sample is based initially on a flux-limited 24-$\\mu$m catalogue, the final selection is based on \\lir, with model-dependent extrapolations from the mid-IR (accurate to $\\ls2\\times$ across all bins -- Fig.~\\ref{lirz}). This should lead to less contamination by AGN at the blue end of the rest-frame 8--1000-$\\mu$m band, where the relative contribution to \\sir\\ can be substantial \\citep[Figure~11 --][]{ivison10}. Using our new sample, there is no strong evidence that \\qir\\ changes as we move back into the epoch of galaxy formation at $z\\sim 2$, with $\\gamma=-0.04\\pm 0.03$ where \\qir\\ $\\propto (1+z)^{\\gamma}$, consistent with the findings of \\citet{sargent10b}. If we discount the $z<0.5$ data, which comprise only 16 galaxies which are not well matched in \\lir\\ to the higher redshift bins, we find $\\gamma = -0.26\\pm 0.07$. This is similar to the $\\gamma = -0.15\\pm 0.03$ found by \\citet{ivison10} who noted reports that evolution in \\qir\\ could be related to the radio background seen by ARCADE\\,2 \\citep{fixsen10, seiffert10}. Our sample, with $\\left\\langle S_{\\rm 1.4GHz} \\right\\rangle \\ls 10\\,\\mu$Jy at $z\\gs 1$, is consistent with the idea that evolution of the FIRRC might be driven by $<$10-$\\mu$Jy radio activity amongst ordinary star-forming galaxies at $z>1$ \\citep{singal10}." }, "1005/1005.1591_arXiv.txt": { "abstract": "{It has been shown that by means of different physical mechanisms the expansion of H\\,{\\sc{ii}} regions can trigger the formation of new stars of all masses. This process may be important to the formation of massive stars but has never been quantified in the Galaxy. } {We use $\\it{Herschel}$-PACS and -SPIRE images from the $\\it{Herschel}$ Infrared survey of the Galactic plane, Hi-GAL, to perform this study.} {We combine the $\\it{Spitzer}$-GLIMPSE and -MIPSGAL, radio-continuum and sub-millimeter surveys such as ATLASGAL with Hi-GAL to study Young Stellar Objects (YSOs) observed towards Galactic \\HII\\ regions. We select a representative \\HII\\ region, N49, located in the field centered on $l$=30$\\degr$ observed as part of the Hi-GAL Science Demonstration Phase, to demonstrate the importance Hi-GAL will have to this field of research. } {Hi-GAL PACS and SPIRE images reveal a new population of embedded young stars, coincident with bright ATLASGAL condensations. The Hi-GAL images also allow us, for the first time, to constrain the physical properties of the newly formed stars by means of fits to their spectral energy distribution. Massive young stellar objects are observed at the borders of the N49 region and represent second generation massive stars whose formation has been triggered by the expansion of the ionized region.} {The first Hi-GAL images obtained using PACS and SPIRE have demonstrated the capability to investigate star formation triggered by \\HII\\ regions. With radio, submillimeter, and shorter wavelength infrared data from other surveys, the Hi-GAL images reveal young massive star-forming clumps surrounding the perimeter of the N49 \\HII\\ generated bubble. Hi-GAL enables us to detect a population of young stars at different evolutionary stages, cold condensations only being detected in the SPIRE wavelength range. The far IR coverage of Hi-GAL strongly constrains the physical properties of the YSOs. The large and unbiased spatial coverage of this survey offers us a unique opportunity to lead, for the first time, a global study of star formation triggered by \\HII\\ regions in our Galaxy. } ", "introduction": "Ionized (H\\,{\\sc{ii}}) regions are known to trigger the formation of stars by means of various physical mechanisms (see Elmegreen~\\cite{elm98} and Deharveng et al.~\\cite{deh05} for a review). Several H\\,{\\sc{ii}} regions have been studied individually in the context of triggered star formation, focusing on the associated neutral material and the young stellar population (Zavagno et al.~\\cite{zav06}; Deharveng et al.~\\cite{deh09}; Pomar\\`es et al.~\\cite{pom09}; Bieging et al.~\\cite{bie09}). These studies have shown that the expansion of H\\,{\\sc{ii}} regions can trigger the formation of new stars of all masses. The $\\it{Spitzer}$-GLIMPSE survey of the Galactic plane (Benjamin et al.~\\cite{ben03}) detected nearly 600 bubbles (Churchwell et al. \\cite{chu06}). Deharveng et al. (\\cite{deh10}) selected a series of 102 ionized bubbles and studied the star formation in their surroundings using $\\it{Spitzer}$-GLIMPSE and MIPSGAL (Carey et al. \\cite{car09}), radio (MAGPIS; Helfand et al. \\cite{hel06} and VGPS; Stil et al. \\cite{sti06}), and ATLASGAL (Schuller et al. \\cite{sch09}) data. They show that 86\\% of these bubbles enclose H\\,{\\sc{ii}} regions, and that more than 20\\% of 64 bubbles (for which the ATLASGAL angular resolution is sufficient to resolve the spatial distribution of cold dust) show massive star formation on their borders. This indicates that triggering is important in the creation of massive stars and that hot photodissociation regions (PDRs) are a good place to study the earliest phases of massive-star formation. Our long-term aim is to use Hi-GAL (Molinari et al.~\\cite{mol10}), combined with existing infrared, submillimeter, and radio surveys, to study the influence of H\\,{\\sc{ii}} regions on triggering the formation of stars in our Galaxy. Hi-GAL's extended wavelength coverage towards the far-infrared and its unprecedented sensitivity offer a unique opportunity to detect an embedded population of young sources that are not detected at shorter wavelengths. This allows us to observe intermediate and high-mass YSOs over the complete range of evolutionary stages. The unprecedented resolution of Hi-GAL also offers the opportunity to accurately characterize the physical nature of the sources by means of a detailed fit to their spectral energy distribution (SED). In this Letter we study the bubble-shaped ionized region, N49, from the Churchwell et al.~(\\cite{chu06}) catalogue to illustrate the purpose of our project. The N49 bubble was studied by Watson et al.~(\\cite{wat08}) and by Deharveng et al.~(\\cite{deh10}). However, these studies had no information in the 70--500\\,$\\mu$m range. This information, obtained with the PACS and SPIRE data presented here, allows us to discuss the star formation of this region in detail. ", "conclusions": "We have presented the first PACS and SPIRE images from Hi-GAL of the bubble-shaped Galactic \\HII\\ region N49. This region is used as an illustration of the study dedicated to the star formation triggered by Galactic \\HII\\ regions that we plan to lead for the whole survey, combining the Hi-GAL results with other infrared and radio surveys of the Galactic plane. We have shown that: \\begin{itemize} \\item{The Hi-GAL SPIRE and PACS images allow us to study the distribution of young sources towards N49. The far-IR fluxes have been measured and strongly constrain the spectral energy distribution of these sources. This allows us to characterize their properties. } \\item{The PACS images reveal the existence of red young stellar objects towards two ATLASGAL condensations, sources that had not been previously detected at shorter wavelengths. The ultracompact \\HII\\ region is not clearly seen in the $\\it{Herschel}$ range. The bright YSO\\#3 Watson et al. observed to be coincident with condensation 1 dominates the emission at longer wavelengths.} \\item{SED fits for the 3 sources detected by $\\it{Herschel}$ towards millimeter condensations using the Robitaille et al.~(\\cite{rob07}) model show that these sources are young and massive. However, their age has not been constrained and other indicators are needed to refine the discussion of star formation history in this region.} \\item{Five massive stars are forming in the N49 PDR. The high star formation efficiency in N49 may be due to the presence of winds from the first generation massive star.} \\end{itemize} The study of N49 using a multiwavelength approach shows that Hi-GAL enables measurements for the crucial far-IR range to be made that are essential to infer the properties of the YSOs. Hi-GAL clearly provides important insight into star formation triggered by expanding \\HII\\ regions. Seventy-six \\HII\\ regions of all shapes are detected in the $l$=30$\\degr$ field and a higher density of Hi-GAL sources is clearly observed towards these regions, indicating that star formation triggered by \\HII\\ regions may be an important process." }, "1005/1005.1244_arXiv.txt": { "abstract": "{It is difficult to determine masses and test formation models for brown dwarfs, because they are always above the main sequence, so that there is a degeneracy between mass and age. However, for brown dwarf companions to normal stars, such determinations may be possible, because one can know the distance and age of the primary star. As a result, brown dwarf companions are well-suited to testing formation theories and atmosphere models. } {With more adaptive optics images available, we aim at detecting orbital motion for the first time in the system TWA 5 A+B. } {We measured separation and position angle between TWA 5 A and B in each high-resolution image available and followed their change in time, because B should orbit around A. The astrometric measurement precision is about one milli arc sec. } {With ten year difference in epoch, we can clearly detect orbital motion of B around A, a decrease in separation by $\\sim 0.0054^{\\prime \\prime}$ per year and a decrease in position angle by $\\sim 0.26^{\\circ}$ per year. } {TWA 5 B is a brown dwarf with $\\sim 25$ Jupiter masses (Neuh\\\"auser et al. 2000), but having large error bars (4 to 145 Jupiter masses, Neuh\\\"auser et al. 2009). Given its large projected separation from the primary star, $\\sim 86$ AU, and its young age ($\\sim 10$ Myrs), it has probably formed star-like, and would then be a brown dwarf companion. Given the relatively large changes in separation and position angle between TWA 5 A and B, we can conclude that they orbit around each other on an eccentric orbit. Some evidence is found for a curvature in the orbital motion of B around A - most consistent with an elliptic (e=0.45) orbit. Residuals around the best-fit ellipse are detected and show a small-amplitude ($\\sim 18$ mas) periodic sinusoid with $\\sim 5.7$ yr period, i.e., fully consistent with the orbit of the inner close pair TWA 5 Aa+b. Measuring these residuals caused by the photocenter wobble - even in unresolved images - can yield the total mass of the inner pair, so can test theoretical pre-main sequence models. } \\titlerunning{Orbital motion of TWA 5 B} ", "introduction": "The star TWA 5 is one of the five original members of the TW Hya association (TWA), a group of 5 to 12 Myr young stars (Kastner et al. 1997), where no gas clouds are left from the star formation process (Tachihara et al. 2009); see Torres et al. (2008) for a recent review on TWA. TWA 5 is an M1.5 weak-line T Tauri star (Webb et al. 1999) with variable H$\\alpha$ emission, hence still ongoing accretion (Mohanty et al. 2003). The central star itself is either a very close ($\\le 66$ milli arc sec or mas) binary (Konopacky et al. 2007, henceforth K07) or even triple (Torres et al. 2003). The close inner pair TWA 5 Aa+b has a total dynamical mass of $0.71 \\pm 0.14$~M$_{\\odot}$ (assuming 44 pc as distance) and an orbital period of $5.94 \\pm 0.09$ years (K07). The wide companion TWA 5 B was originally discovered by Webb et al. (1999) and Lowrance et al. (1999) and confirmed as co-moving with TWA 5 A by Neuh\\\"auser et al. (2000). The spectral type of TWA 5 B is M8-9 (Webb et al. 1999, Lowrance et al. 1999, Neuh\\\"auser et al. 2000, Mohanty et al. 2003). The mass of the companion is between 15 and 40 Jupiter masses just from temperature, luminosity, and theoretical hot-start model tracks (Neuh\\\"auser et al. 2000). The mass lies anywhere between 4 and 145 Jupiter masses, if calculated from temperature ($2800 \\pm 100$ K), luminosity ($\\log(L_{bol}/L_{\\odot}) = -2.62 \\pm 0.30$ at $44 \\pm 4$ pc), and gravity ($\\log g = 4.0 \\pm 0.5$), as obtained by comparison of a Sinfoni K-band spectrum with Drift-Phoenix model atmospheres (Neuh\\\"auser et al. 2009). The system TWA 5 A+B was observed by several teams with ground-based adaptive optics (AO) and/or the Hubble Space Telescope (HST). We obtained two more recent images, so that we can now investigate possible orbital motion of B around A with a 10 year difference in epoch (first preliminary results in Schmidt et al. 2008). We can then also try to detect the orbital motion of Ab and Aa around each other as residuals of the much longer orbit of B around A due to a periodic wobble of the photocenter of the close Aa+Ab pair. ", "conclusions": "" }, "1005/1005.5301_arXiv.txt": { "abstract": "A high-order scheme for direct numerical simulations of turbulent combustion is discussed. Its implementation in the massively parallel and publicly available {\\sc Pencil Code} is validated with the focus on hydrogen combustion. Ignition delay times (0D) and laminar flame velocities (1D) are calculated and compared with results from the commercially available Chemkin code. The scheme is verified to be fifth order in space. Upon doubling the resolution, a 32-fold increase in the accuracy of the flame front is demonstrated. Finally, also turbulent and spherical flame front velocities are calculated and the implementation of the non-reflecting so-called Navier-Stokes Characteristic Boundary Condition is validated in all three directions. ", "introduction": "Modeling of turbulence is one of the largest research areas within flow mechanics. Turbulent combustion inherits all the properties of non-reacting turbulent flow. The most important addition is linked to the highly nonlinear reaction processes, and models for this are called combustion models. Two additional challenges in turbulent combustion are the very sharp changes in density and differential diffusion of mass and heat. For combustion processes it is crucial to be able to simulate the mixing of the combustible species correctly. Traditionally this has been done by means of mixing models in Reynolds Averaged Navier Stokes (RANS) codes by combining, e.g., the $k$-$\\epsilon$ turbulence model and the eddy dissipation concept (or EDC) ``mixing'' model \\citep{MH76,M81,M89}, or in Large Eddy Simulation (LES) \\citep{PV05} where a sub-grid model is used both for the turbulence and for the scalar mixing. There are however major and still unresolved problems related to modelling of what happens on the very smallest scales with these methods. Several RANS codes with detailed chemistry are commercially available \\citep{c10,f10}, and there are a huge number of these codes found as in-house codes at different academic institutions and in many industrial departments around the world. There are also freely available open-source RANS codes with detailed chemistry \\citep{o10}. The reason for the popularity of RANS is its low demand on computational resources. Because of this RANS has, for decades, been the most used type of code for industrial purposes. Nevertheless, also LES has increased in popularity during the last years, and this has led, for example, to the inclusion of a LES module in \\cite{f10}. Most LES codes for combustion today are, however, in-house codes owned by different academic institutions. The most accurate way of simulating turbulent combustion is to use Direct Numerical Simulation (DNS)~\\citep{PV05} instead of RANS or LES. In DNS one resolves the full range of time and length scales of both the turbulence and the chemistry (using accurate high-order numerical methods for computational efficiency). The problem with DNS is however that it is very resource demanding, both on CPU-hours and memory. In this paper we present the implementation of a detailed chemistry module in a finite-difference code \\citep{PC} for compressible hydrodynamic flows. The code advances the equations in non-conservative form. The degree of conservation of mass, momentum and energy can then be used to assess the accuracy of the solution. The code uses six-order centered finite differences. For turbulence calculation we normally use the RK3-2N scheme of \\cite{Wil80} for the time advancement \\citep{BD02}. This scheme is of Runge-Kutta type, third order, and it uses only two chunks of memory for each dependent variable. For hydrodynamic calculations, the lengths of the time step is calculated based on a number of constraints involving maximum values of velocity, viscosity, and other quantities on the right-hand sides of the evolution equations. In some cases we use instead a fifth-oder Runge-Kutta-Fehlberg scheme with an automatic adaptive time step, subject to the aforementioned hydrodynamic constraints. However, in many cases we found it advantageous to use a fixed time step whose length is estimated based on earlier trial runs with an automatically calculated time step. On a typical processor, the cache memory between the CPU and the RAM is not big enough to hold full three-dimensional data arrays. Therefore, the {\\sc Pencil Code} has been designed to evaluate first all the terms on the right-hand sides of the evolution equations along a one-dimensional subset (pencil) before going to the next pencil. This implies that all derived quantities exist only along pencils. Only in exceptional cases do we allocate full three-dimensional arrays to keep derived quantities in memory. However, most of the time, multiple operations including the calculation of derivatives is performed without using intermediate storage. As far as we are aware, no open source high-order DNS code with detailed chemistry is currently available. The amount of man-hours for implementing a fully parallelized DNS code with detailed chemistry is enormous. It is therefore now timely to make such a code available in the public domain and to encourage further development by a wider range of scientists. Here we describe the implementation of such a scheme in the {\\sc Pencil Code}, which is currently maintained under the Google Code subversion repository, \\url{http://pencil-code.googlecode.com/}. The code is highly modular and comes with a large selection of physics modules. It is portable to all commonly used architectures using Unix or Linux operating systems. The code is well documented and independent of external libraries and any third party licenses. All parts of the code, including the current chemistry implementation, is therefore explicitly open source code. In particular, there are no pre-compiled binary files. Consequently there are no licenses required for running any part of the code. It is therefore straightforward to download the full source code from the original subversion repository on google-code. The Message Passing Interface libraries are needed when running on multiple processors, but all parts of the code can also run on a single processor without these libraries. The integrity of the code is monitored through the automatic execution of a selection of test cases on various platforms at different sites. The detailed history of the code with about 14,000 revisions is accessible. It should be emphasized that the use of high-order discretization is critical for optimizing the accuracy at a given resolution. Doubling the resolution of a 3D explicit code require 16 times more CPU time, but this increases the accuracy by a factor of 32. In fact, switching to a derivative module with a tenth order scheme is straightforward and not significantly more expensive. ", "conclusions": "In this paper we have presented a high-order public domain code for direct numerical simulation of compressible flows with detailed chemical reactions. The {\\sc Pencil Code} provides sixth-order spatial accuracy in the simple one-step reaction case, and fifth order accuracy in the case where upwinding for density advection is necessary. For validation purposes we compare our results with the Chemkin tool for 0D and 1D test problems, and show that they are in good agreement. Finally, we calculate the flame speed in 3D both in laminar and turbulent cases. The code is well suited for considering also more complicated reaction schemes such as methane combustion. Furthermore, it is straightforward to consider the interaction with additional chemicals such as nitrogen and to follow the production of NOx gases. In particular, it is important to consider combustion in the presence of steam. This is well known to lead to a reduction of NOx gases. Combustion in the presence of more complicated boundary conditions involving, for example, smaller inlet geometries has also been considered. Some of these cases, including those with a turbulent inlet, are available among the many sample cases that come with the code. For the benefit of the community, it is advantageous if prospective contributers to the code ask one of the code owners listed on \\url{http://pencil-code.googlecode.com/} to obtain permission as a committer." }, "1005/1005.2521_arXiv.txt": { "abstract": "{} { We identify a prominent absorption feature at 1115 GHz, detected in first HIFI spectra towards high-mass star-forming regions, and interpret its astrophysical origin. } { The characteristic hyperfine pattern of the \\hoplus{} ground-state rotational transition, and the lack of other known low-energy transitions in this frequency range, identifies the feature as \\hoplus{} absorption against the dust continuum background and allows us to derive the velocity profile of the absorbing gas. By comparing this velocity profile with velocity profiles of other tracers in the DR21 star-forming region, we constrain the frequency of the transition and the conditions for its formation. } { In DR21, the velocity distribution of \\hoplus{} matches that of the \\CII{} line at 158\\ ${\\mu}$m and of OH cm-wave absorption, both stemming from the hot and dense clump surfaces facing the \\HII{}-region and dynamically affected by the blister outflow. Diffuse foreground gas dominates the absorption towards Sgr~B2. The integrated intensity of the absorption line allows us to derive lower limits to the \\hoplus{} column density of $7.2\\times 10^{12}$~cm$^{-2}$ in NGC~6334, $2.3\\times 10^{13}$~cm$^{-2}$ in DR21, and $1.1\\times 10^{15}$~cm$^{-2}$ in Sgr~B2. } {} ", "introduction": "Oxidaniumyl or oxoniumyl \\citep[][]{IUPAC}, the reactive water cation, \\hoplus{}, plays a crucial role {\\fourthchanged in the chemical network describing the formation of oxygen-bearing molecules} in UV irradiated parts of molecular clouds \\citep{vDh_Black,Gerin}. It was identified at optical wavelengths in the tails of comets in the 1970's \\citep{Fehrenbach,H2O+_Kohoutek_1974a,H2O+_Kohoutek_1974b}, but its detection in the general interstellar medium has proven elusive. We report a detection of the ground-state rotational transition of \\hoplus{} in some of the first spectra taken with the HIFI instrument \\citep{HIFI} on board the {\\it Herschel} Space Observatory \\citep{Herschel} during the performance verification campaign and early science observations. Section 2 briefly introduces the properties of the sources where \\hoplus{} was detected. Section 3 summarises the spectroscopic data of the molecule. The observations and the line identification are described in Sect. 4 {\\fourthchanged and in Sect. 5 we discuss} the physical properties of the \\hoplus{} absorption layer. ", "conclusions": "That \\hoplus{} shows up in absorption against the dust continuum implies that the excitation of the molecule must be colder than the dust. As a reactive ion (see the discussion by \\citealt[][]{Black,Staeuber} for CO$^+$), \\hoplus{} is not expected to be in thermal equilibrium at the kinetic temperature of the gas. Its excitation reflects either the chemical formation process or the radiative coupling with the environment. From a single absorption line, {\\fourthchanged one can only} provide a lower limit to the \\hoplus{} column density, assuming a low excitation temperature where basically all \\hoplus{} resides in the ground state, which is applicable to temperatures well below the upper level energy of 53~K. Table~\\ref{line_parameter} provides the integral over the optical depth of the hyperfine components in the low temperature limit. For the overall $J=3/2-1/2$ fine structure transition, we obtain a line integrated optical depth of $\\int \\tau d\\nu/N\\sub{H_2O^+}=4.70\\,10^{-13}$ km s$^{-1}$ cm$^2$ per molecule, resulting in a lower limit to the \\hoplus{} ground-state column densities of $7.2\\times 10^{12}$~cm$^{-2}$ for NGC~6334, $2.3\\times 10^{13}$~cm$^{-2}$ for DR21, and $1.1\\times 10^{15}$ cm$^{-2}$ for Sgr~B2. These values are lower limits not only because of to the low-temperature approximation, but also because they assume that the absorption occurs in front of the continuum source and {\\fourthchanged not within the dusty cloud}, where the line absorption is partially compensated by dust emission. There may also be additional amounts of \\hoplus{} in the para species that would not contribute to the 1115~GHz line. Altogether, the total \\hoplus{} column density could be much higher than the lower limits given here. The excellent correlation between the \\hoplus{} profile and the OH absorption profile in DR21 indicates that both species occur in the same thin layer of hot gas \\citep{Jones} that directly faces the \\HII{} region at the blue-shifted blister outflow. There is no obvious correlation with the distributions of CO, H$_2$O, or HCO$^+$. For Sgr~B2, we can clearly identify absorption in multiple translucent foreground clouds. Their densities must be high enough to produce some molecular hydrogen, but low enough not to quickly destroy the \\hoplus{}. For NGC6334, the gas component producing the \\hoplus{} absorption remains unidentified. {\\changed With the identification of H$_2$O$^+$ in the interstellar medium, we provide a {\\thirdchanged first step to quantifying} an important intermediate node in the oxygen chemical network, connecting OH$^+$ in diffuse clouds and at cloud boundaries, through H$_3$O$^+$, with water in denser and cooler cloud parts.} To obtain {\\fourthchanged an estimate for the total \\hoplus{} abundance, we need to measure the excitation temperature of \\hoplus{}.} Observations of additional transitions of \\hoplus{}, such as those at 742~GHz, are therefore essential." }, "1005/1005.0148_arXiv.txt": { "abstract": "We constrain a stochastic background of primordial magnetic field (PMF) by its contribution to the angular power spectrum of cosmic microwave background anisotropies. We parametrize such stochastic background by a power-law spectrum with index $n_B$ and by its Gaussian smoothed amplitude $B_\\lambda$ on a comoving length $\\lambda$. We give an approximation for the spectra of the relevant correlators of the energy-momentum of the stochastic background of PMF for any $n_B$. By using the WMAP 7 year data in combination with ACBAR, BICEP and QUAD we obtain the constraint $B_{1 {\\rm Mpc}} < 5.0$ nG at $95\\%$ confidence level for a stochastic background of non-helical PMF. We discuss the relative importance of the scalar and vector contribution to CMB anisotropies in obtaining these constraints. We then forecast {\\sc Planck} capabilities in constraining $B_{1 {\\rm Mpc}}$. ", "introduction": "The origin of the large scale magnetic fields observed in galaxies and clusters of galaxies is an open issue of great importance in modern astrophysics (see \\cite{Widrow:2002ud} for a review). Primordial magnetic fields (PMF) generated in the early Universe could have been the seeds for large scale magnetic fields and have left an imprint in the anisotropy pattern of the cosmic microwave background (CMB). A primordial hypothesis for generating the seeds amplified afterwards by adiabatic compression and dynamo - cannot be discarded \\cite{Widrow:2002ud}, also in light of recent observations of strong magnetic fields in galaxies at high redshift \\cite{Bernet:2008qp,Wolfe:2008nk}. PMF with a comoving amplitude of several nG can leave interesting imprints on CMB anisotropies. A stochastic background of PMF is modelled as a fully inhomogeneous component and its energy momentum tensor (EMT) - quadratic in the magnetic fields - is considered at the same footing as linear inhomogeneities in the other components and linear metric fluctuations. A stochastic background of PMF generates independent modes for all kinds of linear perturbations: there has been several studies for scalar \\cite{KL,yamazaki,KR,GK,FPP,PFP,SL,BoCa}, vector \\cite{SB,MKK,lewis} and tensor \\cite{DFK,MKK,CDK} perturbations. See Refs. \\cite{PFP,SL} for studies which take into account all the types of contributions. A stochastic background of PMF affects also the statistics of CMB anisotropies, and not only its power spectra: being quadratic in the magnetic field amplitude, the EMT is non-Gaussian distributed \\cite{BC} and therefore the bispectrum of CMB anisotropies can also be a useful probe \\cite{CFPR,SS}. In our previous works \\cite{FPP,PFP} we have refined the computation of CMB anisotropies in presence of a stochastic background of PMF: we have computed the initial conditions for cosmological perturbations in the radiation era keeping into account only relativistic degrees of freedom and the correlators for the Fourier transforms of the EMT in presence of a sharp cut-off which mimics the damping scale due to viscosity \\cite{FPP,PFP} for few values of the spectral index $n_B$ of the stochastic background. In this paper we use and extend the above theoretical description to derive the CMB constraints on a stochastic background of PMF which can be obtained by current and future data. We therefore use a modified version of \\texttt{CosmoMC} \\cite{cosmomc} connected with a modified version of \\texttt{CAMB} \\cite{camb} containing all the above features to constrain $B_\\lambda$ - the amplitude of the magnetic field smoothed over a comoving scale $\\lambda$ - and $n_B$ with the most recent compilation of CMB anisotropies data, therefore updating previous investigations \\cite{yamazaki,yamazakilast,shawlewis2}. We also forecast the {\\sc Planck} \\cite{bluebook} capabilities in constraining such a background of PMF. Our paper is organized as follows. In Section II we present the theoretical set-up needed for deriving the CMB constraints. As new results we give an approximation for the PMF energy-momentum valid for any $n_B$, we discuss the correlator between the energy density and the Lorentz force and we extend the regular initial conditions for cosmological perturbations in the relativistic regime in presence of a stochastic background of PMF \\cite{PFP} including the correction due to matter, collecting the details in Appendix A, B and C, respectively. In Section III we discuss the constraints from WMAP 7 years data \\cite{jarosik,larson}, ACBAR \\cite{ACBAR}, BICEP \\cite{BICEP} and QUAD \\cite{QUAD} on a flat $\\Lambda$CDM model plus a stochastic background of PMF. We present the Planck capabilities in constraining $B_{\\lambda}$ and $n_B$ in Section IV and we summarize our results in Section V. ", "conclusions": "We studied the constraints on a stochastic background of PMF by current and forthcoming CMB data. In doing this, we have improved the theoretical CMB predictions in presence of a stochastic background of PMF. We gave approximations for the relevant components of the EMT for any spectral index $n_B$. We considered the correlation between $\\rho_B$ and $L_B$ and showed how the previous choice of total anti-correlation \\cite{FPP,PFP} was a rather good approximation, in particular for red values of $n_B$. We computed the initial conditions for magnetic scalar mode in presence of matter corrections showing how these corrections do not affect the results contrary to what claimed in \\cite{Kojima}. On the basis of previous works, we have considered only the scalar and vector contribution, and by using their regular initial condition, we constrain $B_{1 {\\rm Mpc}} < 5.0$ nG and $n_B < - 0.12$ at $95\\%$ confidence level. with the most updated combination of CMB anisotropies. {\\sc Planck} will be able to constrain the spectrum of a stochastic background of PMF even further at the level of $2.7$ nG. \\vspace{1cm} \\begin{center} {\\bf Acknowledgements} \\end{center} We wish to thank C. Caprini, R. Durrer, J. Hamann and L. Hollenstein for useful comments. This work has been done in the framework of the {\\sc Planck} LFI activities and is partially supported by ASI contract {\\sc Planck} LFI activity of Phase E2. We acknowledge the use of the Legacy Archive for Microwave Background Data Analysis (LAMBDA). Support for LAMBDA is provided by the NASA Office of Space Science. \\begin{widetext} \\begin{figure} \\includegraphics[scale=1]{n23scal.eps}\\includegraphics[scale=1]{nm27scal.eps}\\\\ \\includegraphics[scale=1]{n23L.eps}\\includegraphics[scale=1]{nm27L.eps}\\\\ \\includegraphics[scale=1]{n23V.eps}\\includegraphics[scale=1]{nm27V.eps} \\caption{Comparison of spectral fit (dashed line) and exact spectrum (solid line) for the rescaled magnetic energy density $|\\rho_B (k)|^2/U$ (top panel), scalar Lorentz force $|L_B (k)|^2/U$ (middle panel) and the vector part of the anisotropic stress $|\\Pi^{(V)} (k)|^2/(2 U)$ (bottom panel) as a function of $k/k_D$ with $U=(A^2 k_D^{2n_B+3})/(512 \\pi^4) $. In the left (right) column $n_B=2.3$ ($n_B=-2.7$) is displayed.} \\label{Comparison_fit_scal} \\end{figure} \\appendix" }, "1005/1005.5071_arXiv.txt": { "abstract": "Nucleosynthetic yield predictions for multi-dimensional simulations of thermonuclear supernovae generally rely on the tracer particle method to obtain isotopic information of the ejected material for a given supernova simulation. We investigate how many tracer particles are required to determine converged integrated total nucleosynthetic yields. For this purpose, we conduct a resolution study in the number of tracer particles for different hydrodynamical explosion models at fixed spatial resolution. We perform hydrodynamic simulations on a co-expanding Eulerian grid in two dimensions assuming rotational symmetry for both pure deflagration and delayed detonation Type Ia supernova explosions. Within a given explosion model, we vary the number of tracer particles to determine the minimum needed for the method to give a robust prediction of the integrated yields of the most abundant nuclides. For the first time, we relax the usual assumption of constant tracer particle mass and introduce a radially varying distribution of tracer particle masses. We find that the nucleosynthetic yields of the most abundant species (mass fraction $ > 10^{-5}$) are reasonably well predicted for a tracer number as small as 32 per axis and direction -- more or less independent of the explosion model. We conclude that the number of tracer particles that were used in extant published works appear to have been sufficient as far as integrated yields are concerned for the most copiously produced nuclides. Additionally we find that a suitably chosen tracer mass distribution can improve convergence for nuclei produced in the outer layer of the supernova where the constant tracer mass prescription suffers from poor spatial resolution. ", "introduction": "\\label{sec:int} The basic processes of Type~Ia supernovae (SNe~Ia) have been proposed almost 50 years ago \\citep[see e.g.][for a review]{hillebrandt2000a}: A thermonuclear explosion in electron-degenerate matter \\citep{hoyle1960a} produces radioactive \\nuc{56}{Ni} that by its decay delivers energy at exactly the rate observed in SN~Ia light curves \\citep{pankey1962a,truran1967a, colgate1969a, kuchner1994a}. Despite this long history, the questions of the progenitor system and the explosion scenario are not completely answered. Pure deflagrations in Chandrasekhar-mass white dwarfs, currently somewhat disfavoured for their apparent inability to produce bright explosions, have been one of the contending explosion scenarios for a long time \\citep[e.g.][]{nomoto1984a,gamezo2003a,roepke2007c}. If the initial deflagration can transition into a detonation \\citep[e.g.][]{khokhlov1997a,roepke2007d,woosley2007a,woosley2009a}, then better agreement of the models with observations can be obtained \\citep[e.g.][]{roepke2007b,bravo2008a,kasen2009a}. If such a deflagration to detonation transition does not occur \\citep[cf.\\ e.g.][]{niemeyer1999a}, the possibility that off-center ignition in a single spot leads to a detonation of a white dwarf out of hydrostatic equilibrium in the so-called GCD model is yet another proposed mechanism \\citep{plewa2004a,jordan2008a,meakin2009a}, although the robustness of the initiation of the detonation in this model is also far from certain \\citep[cf.][]{roepke2007a, seitenzahl2009c, seitenzahl2009b}. However, there are indications from recent stellar population synthesis studies \\citep[e.g.][]{ruiter2009a,mennekens2010a} and X-ray observations of elliptical galaxies and galaxy bulges \\citep{gilfanov2010a}, that the long favoured single degenerate Chandrasekhar-mass progenitor channel is unable to account for observational rates of SNe Ia \\citep{cappellaro1999a}. Supernovae resulting from the merger of two white dwarfs \\citep[e.g.][]{pakmor2010a} have more favourable statistics and remain a possible explosion channel. Last but not least, the double detonation sub-Chandrasekhar mass model, which has received renewed interest of late \\citep{fink2007a,fink2010a, sim2010a}, is currently again considered a serious alternative. The different nucleosynthesis occurring in all these explosion scenarios is an important test for the validity of the respective models. For one-dimensional simulations, the nucleosynthesis can be calculated during the simulation via the coupling of a nuclear reaction network to the hydrodynamics. Since today's most promising explosion models either explicitly break spherical symmetry (e.g.\\ mergers, double detonations, off-center ignitions) or, in the case of centrally ignited spherically symmetric explosions include buoyancy driven turbulent combustion, at least two-dimensional simulations are required to simulate the essential physics of the explosion. Unfortunately, in highly resolved two-dimensional simulations and especially in three-dimensional simulations use of an extended, full nuclear reaction network during the hydrodynamic evolution is computationally not feasible. Consequently, the nuclear energy release for multi-dimensional simulations is modeled with simplified and approximate schemes \\citep[e.g.][]{khokhlov1995a,reinecke2002b,vladimirova2006a,calder2007a,townsley2007a,seitenzahl2009a}. Although some works on multi-dimensional explosion simulations of SN Ia do not calculate any detailed nucleosynthesis at all \\citep[e.g.][]{reinecke2002d,plewa2004a,gamezo2005a,plewa2007a,jordan2008a,bravo2008a}, a detailed isotopic composition of the ejecta is often determined in a post-processing step \\citep[][]{travaglio2004a,travaglio2005a,brown2005a,roepke2006b,fink2010a,maeda2010a}. While alternative methods for obtaining isotopic nucleosynthetic yields for SNe Ia are currently under investigation \\citep{seitenzahl2008b,meakin2009a}, the tracer particle method still remains currently the only viable choice for the task. In Section~\\ref{sec:tpm} we briefly review the tracer particle method and introduce how we distribute tracer particles of variable mass in our initial models. In Section~\\ref{sec:sim} we give details about our hydrodynamic simulations. Section~\\ref{sec:conv} presents the results of a convergence study in the number of tracer particles and highlights the advantages of using a radially varying distribution of tracer masses. We conclude the paper with a summary and discussion in Section~\\ref{sec:conc}. ", "conclusions": "\\label{sec:conc} We have performed a resolution study in the number of tracer particles required to converge on the integrated nucleosynthetic yield calculations for three different SN Ia explosion models. We find that for all explosion models, total nucleosynthetic yields appear to have converged (with few exceptions, see Section~\\ref{sec:conv}) for abundant nuclides with mass fractions $> 10^{-5}$ to better than $2\\%$ at 128 tracer particles per axis and direction. Agreement at better than the $5\\%$ level is already achieved for many of the most abundant nuclides at particle numbers as low as 32 per axis and direction. Based on these results, we extrapolate that published yields from the literature \\citep[][]{travaglio2004a,travaglio2005a,brown2005a,roepke2006b,maeda2010a} are based on sufficiently high numbers of tracer particles per axis that an adequate to good prediction of the most abundant nuclei is given. Furthermore, we have shown that for isotopes whose origin is near the surface of the WD, such as e.g. \\nuc{20}{Ne}, \\nuc{26}{Mg} or \\nuc{27}{Al}, better convergence can be achieved if the constraint of constant tracer mass is relaxed and a choice for the tracer particle masses is made that better spatially resolves the outer layers. The question of how many particles are required to get converged spectra and light curves from subsequent radiative transfer calculations is related. It seems plausible that a convergence of the total integrated yields is a necessary but not sufficient constraint. For the radiative transfer not only the total mass of a given isotope is important but also its location in velocity space. It seems that the method of variable mass tracer particles is also promising in this context. Especially simulations where the nucleosynthesis is dominated by a detonation should greatly benefit from a variable tracer mass distribution, since the mass fractions in the inner regions vary smoothly radially and therefore a de-refinement in the inner part is not a high price to pay for increased resolution in the outer part. However, the reconstruction of an isotopic abundance for every computational cell of the radiative transfer calculation from the tracer particle distribution is not unique. The resultant spectra and light curves depend somewhat on particular choice of reconstruction algorithm used. These questions will likely be addressed in a future study." }, "1005/1005.5592_arXiv.txt": { "abstract": "We investigate the virialization of cosmic structures in the framework of flat FLRW cosmological models, in which the vacuum energy density evolves with time. In particular, our analysis focuses on the study of spherical matter perturbations, as they decouple from the background expansion, ``turn around'' and finally collapse. We generalize the spherical collapse model in the case when the vacuum energy is a running function of the Hubble rate, $\\CC=\\CC(H)$. A particularly well motivated model of this type is the so-called quantum field vacuum, in which $\\CC(H)$ is a quadratic function, $\\CC(H)=n_0+n_2\\,H^2$, with $n_0\\neq 0$. This model was previously studied by our team using the latest high quality cosmological data to constrain its free parameters, as well as the predicted cluster formation rate. It turns out that the corresponding Hubble expansion history resembles that of the traditional $\\Lambda$CDM cosmology. We use this $\\Lambda(t)$CDM framework to illustrate the fact that the properties of the spherical collapse model (virial density, collapse factor, etc.) depend on the choice of the considered vacuum energy (homogeneous or clustered). In particular, if the distribution of the vacuum energy is clustered, then, under specific conditions, we can produce more concentrated structures with respect to the homogeneous vacuum energy case. ", "introduction": "Several cosmological observations (supernovae type Ia, CMB, galaxy clustering, etc.) have converged to a paradigm of a cosmic expansion history that involves a spatially flat geometry and a recently initiated accelerated expansion of the universe (see \\cite{Teg04,Spergel07,essence,Kowal08,Hic09,komatsu08} and references therein). From a theoretical point of view, an easy way to explain such expansion is to consider an additional energy component, usually called dark energy (DE) with negative pressure, that dominates the universe at late times. The simplest DE candidate corresponds to the cosmological constant (see \\cite{Weinberg89,Peebles03,Pad03} for reviews). An elegant model that fits accurately the current observational data, is the spatially flat concordance $\\Lambda$CDM model, which includes cold dark matter (DM) and a cosmological constant, $\\Lambda$. However, the $\\Lambda$ model suffers, among other \\cite{Peri08}, form two fundamental problems: (a) {\\it The ``old'' cosmological constant problem} (or {\\it fine tuning problem}) i.e., the fact that the observed value of the vacuum energy density ($\\rho_{\\Lambda}=\\Lambda c^{2}/8\\pi G\\simeq 10^{-47}\\,GeV^4$) is many orders of magnitude below the value found using quantum field theory \\cite{Weinberg89}, and (b) {\\it the coincidence problem}\\,\\cite{coincidence} i.e., the fact that the matter energy density and the vacuum energy density are of the same order (just prior to the present epoch), despite the fact that the former is a rapidly decreasing function of time while the latter is stationary. Many authors have attempted to solve the above problems (see \\cite{Peebles03,Pad03,Egan08} and references therein), the key approach being the replacement of the constant vacuum energy either with a time evolving DE (quintessence and the like\\,\\cite{Peebles03}), or alternatively with a time varying vacuum energy density, $\\rho_{\\Lambda}(t)$ \\cite{Shap00,Shap02,Reuter00,Babic02,Grande06,Bas09b}. In the original scalar field models\\,\\cite{Dolgov82} and later in the quintessence context, one can ad-hoc introduce an adjusting or tracker scalar field $\\phi$\\,\\cite{Caldwell98} (different from the usual SM Higgs field), rolling down the potential energy $V(\\phi)$, which could mimic the DE \\cite{Peebles03,Pad03,Jassal,SR,Xin,SVJ}. However, it was realized that the idea of a scalar field rolling down some suitable potential does not really solve the problem because $\\phi$ has to be some high energy field of a Grand Unified Theory (GUT), and this leads to an unnaturally small value of its mass, namely one which is beyond all conceivable standards in Particle Physics. As an example, utilizing the simplest form for the potential of the scalar field, $V(\\phi)=m_{\\phi}^2\\,\\phi{^2}/2$, the present value of the associated vacuum energy density is $\\rho_{\\CC}=\\langle V(\\phi) \\rangle\\sim 10^{-11}\\,eV^4$, so for $\\langle \\phi\\rangle$ of order of a typical GUT scale near the Planck mass, $M_P\\sim 10^{19}$ GeV, the corresponding mass of $\\phi$ is expected in the ballpark of $m_{\\phi}\\sim\\,H_0\\sim 10^{-33}\\,eV$. Notice that the presence of such a tiny mass scale in scalar field models of DE is generally expected also on the basis of structure formation arguments\\,\\cite{Mota04,Nunes06}; namely from the fact that the DE perturbations % seem to play an insignificant role in structure formation for scales well below the sound horizon. The main reason for this homogeneity of the DE is the flatness of the potential, which is necessary to produce a cosmic acceleration. Being the mass associated to the scalar field fluctuation, proportional to the second derivative of the potential itself, it follows that $m_{\\phi}$ will be very small and one expects that the magnitude of DE fluctuations induced by $\\phi$ should be appreciable only on length scales of the order of the horizon. Thus, equating the spatial scale of these fluctuations to the Compton wavelength of $\\phi$ (hence to the inverse of its mass) it follows once more that $m_{\\phi}\\lesssim\\,H_0\\sim 10^{-33}\\,eV$. All in all, it appears that the problem that one is creating along with the introduction of $\\phi$ is far more worrisome than the problem one is intending to solve, for one is postulating a mass scale which is $30$ orders of magnitude below the mass scale associated to the value of the vacuum energy density ($m_{\\Lambda}\\equiv\\rho_{\\Lambda}^{1/4}\\sim 2.3\\times 10^{-3}\\,eV$). The analysis of the recent cosmological observations indicates that the DE equation of state (EOS) parameter $w (\\equiv P_{\\rm DE}/\\rho_{\\rm DE})$ is close to $-1$ to within $\\pm 10\\%$, if it is assumed to be constant \\cite{Teg04,Spergel07,essence,Kowal08,Hic09,komatsu08}, whilst it is much more poorly constrained if it varies with time\\,\\cite{essence}. More than two decades ago, Ozer \\& Taha \\cite{Oze87} proposed a time varying $\\Lambda$ as a possible candidate to solve the two fundamental cosmological puzzles, see also \\cite{Free187,Carvalho92} and references therein. In this cosmological paradigm, the dark energy EOS parameter $w$ is strictly equal to -1, but the vacuum energy density (or $\\Lambda$) does evolve with time. Of course, the weak point in this approach % is the unknown functional form of $\\Lambda(t)$, which is however also the case for the vast majority of the DE models. Indeed, in the aforementioned $\\Lambda(t)$ models, the evolution law is purely phenomenological\\,\\cite{Overduin98}, without a concrete % link to fundamental physics, such as the Quantum Field Theory (QFT) in Curved Space time\\,\\cite{ParkerToms09}. As emphasized in \\cite{ShapSol09}, a completely consistent formulation along these lines should eventually be developed, and such investigations could well be at the heart of one the most important endeavors of theoretical cosmology in the years to come. Therefore, the study of cosmic perturbations in these models is very important\\,\\,\\cite{Grande08,GSFS10} as they might reveal surprises not foreseen in the context of the scalar field models. The new effects may have impact both on the cosmological and the \\textit{astrophysical} domains. While we recently analyzed the potential implications on the former\\,\\cite{Bas09c}, here we focus on the latter domain. A pioneering QFT fundamental approach to variable $\\CC$ models was actually proposed long ago within the context of the renormalization group (hereafter RG) in\\, \\cite{NelPan82,Toms83}. Later on, the RG-running framework was further explored from different points of view in \\cite{Shap00,Shap02,Reuter00}, and a more systematic presentation from the viewpoint of QFT in curved space-time by employing the standard perturbation RG-techniques of Particle Physics appeared in\\,\\cite{Shap02,Babic02}. Subsequent elaborations, and comparison with the observational tests, confirmed the phenomenological viability of this approach \\,\\cite{RGTypeIa,SSS04,SS12,Fossil07,Bauer05,FSS09a,BFLWard}. In the class of RG models we shall focus on, the vacuum energy density is expected to vary with time according to the law\\,\\cite{Shap02,RGTypeIa,SSS04,SS12,Fossil07}: $\\CC=n_0+n_2\\,H^2$ (hereafter called the $\\Lambda_{RG}$ model or quantum field vacuum model). As already mentioned, in Ref.\\,\\cite{Bas09c} we have investigated thoroughly the global dynamics of this cosmological model (together with various alternative $\\Lambda(t)$ models), in the light of the most recent cosmological data. However, a serious issue here is how the main features of the largest collapsed cosmic structures, i.e., galaxy clusters, are affected by a running vacuum energy density. We have argued above that this problem can be addressed in scalar field models of the DE, but only at the expense of admitting extremely tiny mass scales which are uncommon in Particle Physics. In this paper, we wish to further explore the alternative option in which the DE component is a time evolving cosmological term $\\Lambda=\\Lambda(t)$, and in this way to assess if the clustering properties of the vacuum energy can shed some light on the fundamental issue of structure formation. The so called spherical collapse model\\,\\cite{Gunn72}, which has a long history in cosmology, is a simple but still a fundamental tool used to describe the growth of bound systems in the universe via gravitation instability\\,\\cite{Pee93}. In the last decade many authors have studied the small scale dynamics using this model and found that the main features of the cosmic structures (collapse factor, virial density, etc) can potentially be affected by the presence of dark energy \\cite{Lahav91,Wang98,Iliev01,Lokas,Battye03,Maini03,Bas03,Wein03, Mota04,manera,Horel05,Zeng05,Maor05,Perc05,Nunes06,Wang06,david,Fran08, Basi07,Shaef08,Lee09,Abramo}. The aim of the present work is to generalize the spherical collapse model within the variable $\\Lambda_{RG}$ cosmological model, in order to understand non-linear structure formation in such cosmologies and investigate the differences with the respect to the expectations of the concordance $\\Lambda$CDM cosmology. The structure of the paper is as follows. The basic theoretical elements of the problem are presented in section 2, where we introduce [for a spatially flat Friedmann-Lema\\^\\i tre-Robertson-Walker (FLRW) geometry] the basic cosmological equations. In section 3 we generalize the virial theorem in the case of the QFT $\\Lambda(t)$ cosmological model. Section 4 outlines the theoretical analysis of the spherical collapse model in which the vacuum energy density varies with the cosmic time, and in section 5 we compare the corresponding theoretical predictions of the different models and present a first attempt to use observational data to constrain the different models. We draw our conclusions in section 6. In the appendix A we remind the reader of some basic elements of the concordance $\\Lambda$CDM model in order to appreciate the fact that the $\\Lambda_{RG}$ cosmology is an interesting extension of the standard model. Finally, in appendix B we provide some basic mathematical formulae, while in appendix C we provide accurate fitting formulae for a few important parameters, ie., the density contrast at the turn around redshift and at the epoch of virialization, which do not have a simple fully analytical form. ", "conclusions": "In this paper we have studied analytically and numerically the spherical collapse model in the case of a time varying vacuum, with $\\CC(H)=n_0+n_2\\,H^2$, for a spatially flat FLRW geometry. We find that the amplitude and the shape of the virial density contrast is affected by the considered status of the vacuum energy model (homogeneous or clustered). We verify that in the case where the distribution of the vacuum energy is clustered the structures produced are more concentrated (under specific conditions) with respect to the homogeneous dark energy case. Finally, % by comparing the predicted virial density contrast at the present epoch with a preliminary analysis of a suitable subsample of the 2MASS High Density Contrast group catalog (at a mean redshift of $\\langle z \\rangle \\simeq 0.015$), we find that the inhomogeneous vacuum energy models can be accommodated, at a $2\\sigma$ level, if the vacuum clustering parameter is within the range: $-0.009\\mincir \\gamma_{s}\\mincir 0.012$. The latter result points to the direction that perhaps the $\\Delta_{vir}$ parameter, once estimated accurately from observations, could be used in order to determine the internal physical properties of the vacuum energy. \\vspace {0.5cm} {\\bf Acknowledgments.} JS has been supported in part by MEC and FEDER under project FPA2007-66665, by the Spanish Consolider-Ingenio 2010 program CPAN CSD2007-00042 and by DIUE/CUR Generalitat de Catalunya under project 2009SGR502. MP acknowledges funding by Mexican CONACyT grant 2005-49878. \\appendix" }, "1005/1005.0717.txt": { "abstract": "We undertake a comprehensive and rigorous analytic study of the evolution of radial profiles of covariant scalars in regular Lema\\^\\i tre--Tolman--Bondi dust models. We consider specifically the phenomenon of ``profile inversions'' in which an initial clump profile of density, spatial curvature or the expansion scalar, might evolve into a void profile (and vice versa). Previous work in the literature on models with density void profiles and/or allowing for density profile inversions is given full generalization, with some erroneous results corrected. We prove rigorously that if an evolution without shell crossings is assumed, then only the `clump to void' inversion can occur in density profiles, and only in hyperbolic models or regions with negative spatial curvature. The profiles of spatial curvature follow similar patterns as those of the density, with `clump to void' inversions only possible for hyperbolic models or regions. However, profiles of the expansion scalar are less restrictive, with profile inversions necessarily taking place in elliptic models. We also examine radial profiles in special LTB configurations: closed elliptic models, models with a simultaneous big bang singularity, as well as a locally collapsing elliptic region surrounded by an expanding hyperbolic background. The general analytic statements that we obtain allow for setting up the right initial conditions to construct fully regular LTB models with any specific qualitative requirements for the profiles of all scalars and their time evolution. The results presented can be very useful in guiding future numerical work on these models and in revising previous analytic work on all their applications. ", "introduction": "\\footnote{In order to motivate the reading of this long article, we have concentrated most of the background material in the Appendices and have written a proper summary of our results and their implications in the final section (section 11). Readers eager to know these results without going through the technical detail are advised to go directly to this section.} % The well known spherically symmetric LTB dust models \\cite{LTB} are among the most useful known exact solutions of Einstein's equations. There is an extensive literature (see \\cite{kras1,kras2} for comprehensive reviews) using these models, as they allow for probing non--linear effects in inhomogeneous sources by means of analytic and/or tractable numeric solutions. In most applications LTB dust solutions lead to simple but effective toy models of cosmological inhomogeneities, as for example in the series of papers \\cite{KH1,KH2,KH3,KH4,BKH}. The models are also useful in other theoretical contexts, such as censorship of singularities \\cite{sscoll,joshi} and even quantum gravity \\cite{quantum}. More recently, LTB models have been extensively used in the widespread effort to explain cosmic dynamics without resorting to an elusive source like dark energy, either by fitting observations \\cite{LTB1,LTB2,LTBkolb,num1,num2}, or in the context of scalar averaging of inhomogeneities \\cite{LTBfin,LTBchin,LTBave1,LTBave2,LTBave3} (see \\cite{celerier,ave_review} for comprehensive reviews on these applications). For a novel theoretical approach to the dynamics of these models see \\cite{wainwright}. Since large scale cosmic structure is dominated by large voids, the study and understanding of the spatial profiles of density and velocity in inhomogeneities is a relevant theoretical and practical issue that could provide important clues on the formation of these voids as part of the dynamics of cosmic sources. While LTB models provide an idealized description of cosmic inhomogeneities, a proper understanding (even qualitative) of radial profiles of scalars in these models still has a significant potential in astrophysical and cosmological implications in this context \\cite{kras1,kras2} (see \\cite{CBK} for an update). In particular, knowing how these radial profiles evolve is extremely useful when applying to these models a formalism of scalar averaging of inhomogeneities, such as Buchert's formalism \\cite{LTBave2,LTBave3}. An important theoretical issue in looking at density radial profiles is the possibility that a concavity change could occur under regular conditions (particularly around a symmetry center). This implies that a given initial density profile with a ``clump'' (over--density) form could evolve into a ``void'' (under--density) profile, or vice versa, a phenomenon that is usually denoted by a ``clump to void'' or ``void to clump'' density profile inversion. In a well known article Mustapha and Hellaby \\cite{mushel} claimed to have furnished proof for the existence of this profile inversion, and to have found the conditions in which it occurs. However, since the conditions provided by these authors are too complicated (involving second and third order radial derivatives), their proof (their section 5) is based on restrictive assumptions, and thus is not applicable to generic LTB models. In the end, they discuss the observational implications of these profile inversion and furnish basically an empiric proof of their existence by looking at specific models and numeric examples where it occurs. Following a completely different methodology from that of Mustapha and Hellaby, we correct, extend and fully generalize their work in this article. By considering fully generic and regular LTB models, we examine the evolution of the radial profiles of the basic covariant scalars: rest mass density, $\\rho$, spatial curvature, $\\RR$, and Hubble expansion scalar, $\\Theta$, aiming specifically at dealing in full generality with the following issues:\\, (i) to provide a precise characterization of radial profiles as ``clumps'' or ``voids'', and (ii) to study in full rigor the conditions that allow for the existence of profile inversions (assuming full regularity). As the reader can find out by going through the article (or by looking at the summary of results in section 11), we do find fully analytic and rigorous conditions for the existence of clump/void radial profiles and their possible inversions for regular LTB models in general. An interesting and similar approach to the study of the evolution of radial profiles is found in the articles by Krasinski and Hellaby \\cite{KH1,KH4} (see the review in \\cite{kras2}). These authors find the conditions for the existence of a unique LTB model, such that a given density or ``velocity'' profile at a given time constant hypersurface can be ``mapped'' to any other profile at a second hypersurface. While there is an obvious intersection and complementarity between the results of this article and those of \\cite{KH1,KH4}, our methodology is different and (in our opinion) less restrictive, as we only require generic initial data to be specified at a single time constant hypersurface that can be considered a Cauchy surface in the context of an initial value problem. A summary of articles that have examined LTB models with void profiles in a perturbative and non--perturbative context can be found in the extensive review of this literature (previous to 1994) given in section 3.1 of \\cite{kras1}, and in the update in section 18.5 of \\cite{kras2}. Some authors \\cite{occhio81,occhio83} have examined non--perturbative void profiles numerically in the context of structure formation scenarios (though, from our results in section 6, the elliptic models in \\cite{occhio81} cannot be free from shell crossings). In some suggested configurations the void is modeled as an under--dense FLRW region, which is connected to a cosmic ``background'' through a section of a closed elliptic model \\cite{cham,sato}. In other articles \\cite{boncham1,boncham2,meszaros}, the void is an LTB region containing a center, matched at a fixed comoving radius to another LTB region (the ``envelope''), which in turn is matched to a FLRW ``exterior''. Although these models are rather artificial, their results agree with those of this article. The contents of the article is summarized in the remaining of this section. Section 2 contains the bare basic background material needed to understand the article. The known analytic solutions of the field equations (given in Appendix A), which are determined by a preferred set of free functions ($M,\\,E,\\,\\tbb$), have become the standard formulation in practically all theoretical, empiric and in numeric work on these models. While this standard parametrization is adequate and works in practice, we utilize an alternative set of covariant quasi--local scalars that are more suitable for our purpose (and that can be very convenient for numerical work, as well as a better understanding of several theoretical and practical issues) \\cite{sussQL1,sussQL,suss08,suss09,suss10a,suss10b}. We also express these scalars and their fluctuations as quantities that are scaled with respect to their initial values in the framework of an initial value problem \\cite{suss10a}. Extra background material, which is important as well but can be distracting if it appears in the main body of the article, has been placed in Appendices B, C and D. For example, the parametrization of the known analytic solutions in terms of the new variables and its initial value formulation (Appendix B), the Hellaby--Lake conditions \\cite{suss10a,HLconds,ltbstuff} that guarantee an evolution free from shell crossings (Appendix C), and a discussion (Appendix D) of the relation between the radial coordinate and the proper radial length, which is the affine parameter along radial rays \\cite{suss10b}. This is important since we are looking at the behavior of covariant scalars in the radial direction. We provide in section 3 a precise and rigorous characterization of the ``clump'' or ``void'' nature of a radial profile of a scalar $A$ in terms of the sign of the gradient $A'$ (monotonicity of $A$), which leads to the definition of a ``turning value'' (to be denoted ``TV of $A$'') as the value $r=\\rtv$ where this gradient vanishes. In section 4 we examine the compatibility between initial clump/void profiles and the Hellaby--Lake conditions, while the concept of ``profile inversion'' is defined in section 5 in terms of the vanishing of the fluctuations defined in section 2, something which can occur with or without the existence of TV's. In section 6 we examine in detail density radial profiles and provide necessary and sufficient conditions for density profile inversions for parabolic, hyperbolic and open elliptic models or regions. We remark that the locus where the density fluctuation vanishes when there are density TV's corresponds to the ``density wave'' (see section 3.1 of \\cite{kras1}). The profiles and profile inversions for spatial curvature and the expansion scalar are respectively examined in sections 7 and 8. The case of closed elliptic models is considered in section 9, as for these models the spherical topology of the hypersurfaces of constant time introduces a TV of the area distance $R$. In section 10 we look at radial profiles and profile inversions in special LTB configurations: simultaneous big--bang singularity, LTB models or regions not containing symmetry centers and configurations constructed by glueing LTB regions with ``mixed'' kinematics (elliptic, parabolic and hyperbolic regions, see \\cite{ltbstuff}), and specially the mixed configuration made by an elliptic ``interior'' region surrounded by a hyperbolic ``exterior''. Section 11 provides a summary and final discussion. Appendix E describes an alternative parametrization of the models in terms of quantities that can be related to the Omega and a Hubble parameters of a FLRW cosmology in the homogeneous limit. It is straightforward to translate the results of this article in terms of these parameters, which have been used in various articles \\cite{num1,LTBfin,suss08} describing cosmological inhomogeneities in various contexts (see also \\cite{suss10a}). Appendix F provides the explicit expressions for the basic covariant scalars that characterize LTB models. ", "conclusions": "We have conducted a comprehensive and rigorous examination of the radial profiles of the main covariant scalars (density, spatial curvature and expansion, respectively denoted by $m,\\,k,\\,\\HH$, see definition (\\ref{mkHdefs})) for regular LTB models in full generality. The time evolution of these profiles has been considered, and in particular, we have addressed the issue of profile inversions in which an initial clump can evolve into a void (or vice versa). The necessary background material is the formalism of quasi--local variables within an initial value formulation (section 3), rephrasing in terms of this formalism the analytic solutions (Appendix B) and the Hellaby--Lake conditions (Appendix C), and the relation between the radial coordinate and proper radial length (Appendix D). The quasi--local density ($m_q$), spatial curvature ($k_q$) and expansion scalar ($\\HH_q$) are given by \\ref{mq})--(\\ref{Hq}), while the fluctuations $\\Da$ follow from (\\ref{Dadef}). \\subsection{Summary of results.} We have used thoughout the article the term ``turning value'' of any scalar $A$ (``TV of $A$'') to denote a value $r=\\rtv$ such that $A'$ vanishes at $r=\\rtv$ at a given hypersurface $\\T[t]$ marked by constant $t$ (this definition also applies to the $A_q$). In general, the value $\\rtv$ is different for each $\\T[t]$, but it can be the same for all $t$ (as for example a regular TV of $R$). Throughout sections 3--8 we have considered only LTB models or regions containing a symmetry center and with an open topology ($R'>0$ holds everywhere: no TV of $R$), leaving the case with closed topology (regular TV of $R$) and special configurations for sections 9 and 10. The main results contained in 10 lemmas that were proven in these sections are summarized below: \\begin{itemize} \\item {\\bf{Section 3}}. The clump/void character of radial profiles of local and quasi--local scalars, $A$ and $A_q$, in radial domains containing a center was given in terms of the existence of turning values, TV's. Assuming that $R'>0$ holds, we proved the following lemmas: % \\begin{itemize} \\item {\\bf{Lemma 1}}: The signs of the radial gradient $A'_q$ is related to the sign of the fluctuation $\\Da$. \\item {\\bf{Lemma 2}}: The necessary and sufficient condition for a TV of $A_q$ at some $r=\\rtv$ is given by $\\Da=0$ at $r=\\rtv$. In general, $\\rtv=\\rtv[t]$ is different at different times (see figure 1). \\item {\\bf{Lemma 3}}: The existence of a TV of $A_q$ in a radial domain is a sufficient condition for the existence of a TV of $A$ in the same domain. \\end{itemize} % These lemmas allow us to examine the monotonicity of scalars by looking at signs and the zeros of their fluctuations. As a consequence of Lemma 3, it is sufficient to examine the profiles of the scalars $A_q$ (which satisfy simpler scaling laws) to know the profiles of the scalars $A$. \\item {\\bf{Section 4}}. Character of initial density and spatial curvature profiles compatible with the Hellaby--Lake conditions. Initial density must have a clump profile in parabolic and elliptic models or regions, while initial clump and void profiles are possible in hyperbolic models or regions. Initial curvature must have a clump profile in elliptic models or regions, but hyperbolic models or regions admit initial clumps and voids. Explicit conditions are given for each case. \\item {\\bf{Section 5}}. Formal definition of profile inversion and its relation with TV's of scalars. Profile inversions can occur with and without TV's. The following lemmas were proven: % \\begin{itemize} \\item {\\bf{Lemma 4}}: A profile inversion of $A_q$ implies a profile inversion of $A$. \\item {\\bf{Lemma 5}}: Sufficient conditions for the existence of a profile inversion of $A_q$ follow from the existence of TVs of $A_q$ plus some extra requirements on the fluctuations $\\Da$. This type of profile inversion is illustrated by figure \\ref{fig1}. \\item {\\bf{Lemma 6}}: Sufficient conditions for the existence of a profile inversion of $A_q$ when there are no TV's of $A_q$. This type of profile inversion is illustrated by figure \\ref{fig2}. \\end{itemize} % Notice that Lemma 4 guarantees that the results of Lemmas 5 and 6 apply also to the local scalars $A$. The conditions in Lemmas 5 and 6 are independent of the Hellaby--Lake conditions, hence the latter place extra constraints when applying these lemmas to specific configurations. % \\item {\\bf{Section 6}}. Density radial profiles and profile inversions are examined in detail. The following lemmas were proven: % \\begin{itemize} \\item {\\bf{Lemma 7}}: Necessary and sufficient condition are given for the `clump to void' density profile inversion with density TV's in regular hyperbolic models or regions. Void to clump inversion is not possible. \\item {\\bf{Lemma 8}}: Necessary and sufficient condition for the `clump to void' inversion without TV's in regular hyperbolic models or regions. Void to clump inversion is not possible. \\item {\\bf{Lemma 9}}: Density profile inversions, with or without density TV's, cannot occur in regular open elliptic models or regions containing a center (no TV's of $R$). \\end{itemize} % The main result of this section (and possible of the whole article) follows as a consequence of these lemmas: \\begin{quote} \\noindent {\\it {the only density profile inversion (with and without TV's) compatible with absence of shell--crossings (the Hellaby--Lake conditions) is the `clump to void' inversion in hyperbolic models or regions.}} \\end{quote} \\noindent The specific conditions outlined by Lemmas 7 and 8 (equations (\\ref{dm32dk}) and (\\ref{CintoV2})), which must be satisfied by hyperbolic models and region, are simple restrictions on their initial value functions (density, spatial curvature and their fluctuations). It is evident that these are not weird or outlandish conditions, but reasonable and easy to prescribe. % \\item {\\bf{Section 7}}. Spatial curvature profiles and their inversions. The only inversion compatible with the Hellaby--Lake conditions is the `clump to void' inversion in hyperbolic models or regions. These inversions only occur in the particular parameter cases when there is no inversion of the density profile. % \\item {\\bf{Table 1}} provides the combination of initial conditions that allow for profile inversions of density and spatial curvature in hyperbolic models or regions. % \\item {\\bf{Section 8}}. Radial profiles and profile inversions of the expansion scalar. Hyperbolic models and regions allow for a regular evolution with and without profile inversions (conditions are provided for each case). However, profile inversions of the expansion scalar necessarily occur in elliptic models or regions containing a center. This is illustrated by figure \\ref{fig3}. \\item {\\bf{Section 9}}. Closed elliptic models. A TV of $R$ occurs under regular conditions at the same value $r=\\rtv$ for all times. We proved the following lemma: % \\begin{itemize} \\item {\\bf{Lemma 10}}: The TV of $R$ at $r=\\rtv$ implies that $\\rtv$ is a common TV of the local and quasi--local density, spatial curvature and expansion scalar. The converse is not true, as we proved in Lemmas 1--5 and 7--8 that TV's of these scalars can occur when there is no TV of $R$ ($R'>0$ holds everywere). \\end{itemize} % As an important corollary of this lemma, the fluctuations $\\Da$ do not change sign because of the TV of $R$. As a consequence, there are no profile inversions of density and spatial curvature, though profile inversions of the expansion scalar must occur as in open elliptic models. % \\item {\\bf{Section 10}}. Special configurations: % \\begin{itemize} \\item {\\underline{Regular hyperbolic and elliptic models with a simultaneous big bang}}. Profile inversions of density and spatial curvature (with or without TV's) cannot occur. Hyperbolic models allow for a void profile for all the time evolution. % \\item {\\underline{Regions not containing symmetry centers}}. The results of all proven lemmas hold, the only caveat being the fact that (in general) we have $\\Da\\ne 0$ at the boundaries of the radial range (as opposed to $\\Da= 0$ strictly holding at the symmetry center). % \\item {\\underline{LTB models with ``mixed'' kinematics}}. These configurations are made by glueing or matching combinations of parabolic, hyperbolic and elliptic regions (see \\cite{boncham1,boncham2,meszaros,ltbstuff}). The results of all proven lemmas hold. % \\item {\\underline{Elliptic region surrounded by a hyperbolic exterior}}. From Lemma 9 and assuming that Hellaby--Lake conditions hold, the elliptic region must have clump profiles for the density and spatial curvature, without inversions. However, void profiles and `clump to void' inversions are possible in the hyperbolic exterior in agreement with Lemma 7 and the combination of initial conditions specified in Table 1. \\end{itemize} \\end{itemize} \\subsection{Relaxation of regularity conditions.} In obtaining the results summarized above we have assumed that the Hellaby--Lake conditions \\cite{HLconds,ltbstuff,suss10a} (see Appendix C) hold for all the evolution times of the models (shell crossings are completely absent). Evidently, if we relax these conditions by demanding that they hold only for all $t>t_i$ for a given $t=t_i$, then some of the results of the Lemmas are no longer binding. The most important example is Lemma 9, which forbids density void profiles and `clump to void' density profile inversion in elliptic models. This result follows from the fact that a density void profile is incompatible with the joint fulfillment of two of the Hellaby--Lake conditions (\\ref{noshxGe}):\\, $\\tbb'\\leq 0$ and $\\tcoll'\\geq 0$, where $\\tbb$ and $\\tcoll$ denote the locus of the initial and collapsing singularity. However, if we demand $\\tcoll'\\geq 0$ to hold but not $\\tbb'\\leq 0$, then a shell crossing singularity necessarily emerges for $t\\approx \\tbb$ for all $r$, but not for $t\\approx \\tcoll$. Depending on the free parameters $m_{qi},\\,k_{qi}$, it may be possible to construct an elliptic model that is free from shell crossings for all $t>t_i$ for some $t_i>\\tbb$ \\cite{suss02}, and that admits a `clump to void' inversion and density void profiles in this time range. Since a dust source does not provide a good description of the physical conditions near an initial singularity in a cosmological model, it may be sufficient for a physically reasonable evolution to demand absence of shell crossings for all $t>t_i$, provided one can justify that the dust description beaks down for $t {\\rm Var}[\\hat\\kappa_{0}(\\boldsymbol{\\ell})]$ indicates the interval in $\\ell$ where the lensing potential can in theory be mapped. % Accordingly, this is also the region where the reconstructed convergence power spectra tracks more closely the input power spectrum. Also visible are the fluctuations in the recovered power spectrum which arise from the subtraction of the noise bias term estimated from a finite number of realizations of the unlensed CMB map. For the {\\planck} experiment, using the real space implementation we achieve a reasonable reconstruction of the input power spectrum in the interval where the input signal is larger than the variance of the estimator, although on smaller scales, $\\ell > 600,$ there is a decrease in the recovered power. Comparing the reconstructed power spectra for the {\\designer} and {\\planck} experiments we observe that the real space implementation is fairly insensitive to the detector noise and returns a reconstruction consistent with the input power spectrum. We can understand these results as follows. \\begin{figure}[t] \\setlength{\\unitlength}{1cm} \\vskip-1.5cm \\centerline{ \\hskip-0.5cm \\includegraphics[width=12cm]{% pikassoyla5aHSdesigner.pdf} \\hskip-3cm \\includegraphics[width=12cm]{% pikassoyla5bRSdesigner.pdf} } \\vskip-7.5cm \\caption{\\baselineskip=0.5cm{ {\\bf Mean square of the convergence map reconstructed with the harmonic space estimator (left panel) and the real space estimator (right panel).} The smooth black curve is the input power spectrum. The light grey and dark grey solid lines are the output power spectra before and after the removal of the bias respectively. The error bars measure the total standard deviation binned over logarithmically spaced intervals in $\\ell.$ Here $\\theta_{fwhm}=7.8',$ $FOV_{map}=66.6^{\\circ}$ and $m_{max}=4,$ $\\ell_{max}=4000$ for no detector noise. }} \\label{fig:cl_kk2_designer} \\end{figure} \\begin{figure}[h] \\setlength{\\unitlength}{1cm} \\vskip-1.5cm \\centerline{ \\hskip-0.5cm \\includegraphics[width=12cm] {% pikassoyla6aHSplanck.pdf} \\hskip-3cm \\includegraphics[width=12cm] {% pikassoyla6bRSplanck.pdf} } \\vskip-7.5cm \\caption{\\baselineskip=0.5cm{ {\\bf Mean square of the convergence map reconstructed with the harmonic space estimator (left panel) and the real space estimator (right panel).} The smooth black curve is the input power spectrum. The light grey and dark grey solid lines are the output power spectra before and after the removal of the bias respectively. The error bars measure the total standard deviation binned over logarithmically spaced intervals in $\\ell.$ Here $\\theta_{fwhm}=7.8',$ $FOV_{map}=66.6^{\\circ},$ and $m_{max}=4,$ $\\ell_{max}=4000$ for PLANCK noise at $\\nu=143~{\\rm GHz}.$ }} \\label{fig:cl_kk2_planck} \\end{figure} As attested by the power spectra of the maps reconstructed with the real space estimator, there is a loss of power at small scales beginning around the angular scale corresponding to the size of the kernel. This loss of power is a consequence of averaging modes smaller than the finite extent of the kernel and hence is an intrinsic constraint of the real space implementation. This intrinsic constraint, studied in more detail in Ref.~\\cite{bucher2010}, can be overcome by shrinking the extent of the kernel in real space, which can be achieved with a smaller beam and detector noise, thereby moving the support for the kernel to larger $\\ell.$ The $\\ell$ interval over which the power spectrum can be recovered by the real space estimator is determined by the characteristic lengths of the map and the kernel as follows. The lower $\\ell$ limit is determined by the size of the map, since it measures the longest wavelength mode that can be enclosed within the size of the map. The upper $\\ell$ limit is determined by the size of the kernel, since it measures the smallest wavelength mode that the kernel can probe, below which there is a loss of power due to the averaging of modes smaller than the size of the kernel. Within this $\\ell$ range, the real space estimator seems to perform fairly well both in the absence and in the presence of detector noise. Note that the $\\ell$ range for a reasonable reconstruction with the harmonic space estimator is wider, being limited at higher $\\ell$ by the angular scale corresponding to the beam size, since this is the scale which constrains the action of the kernel in harmonic space. Comparing the reconstructed power spectra in the absence and in the presence of detector noise, we observe that the real space estimator appears to be insensitive to the experimental noise. This is because the estimate of the convergence in each pixel is given by the sum of the product of pairs of neighbouring pixels weighted by the kernel. As a result, the noise, being independent in each pixel, is averaged out. To substantiate this result, we reconstructed the convergence map from a pure white noise input map using the real space estimator and the harmonic space estimator. In Fig.~\\ref{fig:cl_kk2_planck_pure_noise} we observe that the power spectrum recovered by the real space estimator is about seven orders of magnitude smaller than the input power spectrum, whereas the power spectrum recovered by the harmonic space estimator is comparable to the input power spectrum. Despite our attempt to remove the CMB and the detector noise bias from the reconstructed convergence power spectrum, there still remains a bias in the harmonic space implementation, as observed in Figs.~\\ref{fig:cl_kk2_designer} and \\ref{fig:cl_kk2_planck}. This bias presumably arises from a coupling between unlensed temperature modes due to the finite size of the map, providing an additional source of lensing which we have not treated here. This mode coupling is absent in the reconstruction over the full sky \\cite{inpaint}. The real space implementation appears to be insensitive to this bias. \\begin{figure}[t] \\setlength{\\unitlength}{1cm} \\vskip-1.5cm \\centerline{ \\hskip-0.5cm \\includegraphics[width=12cm] {% pikassoyla7noise.pdf} } \\vskip-7.5cm \\caption{\\baselineskip=0.5cm{ {\\bf Mean square of the convergence map reconstructed with the harmonic space estimator (light grey line) and the real space estimator (dark grey line) from a pure white noise map.} The grey solid lines are the output power spectra whereas the smooth black curve is the input power spectrum. Here $\\theta_{fwhm}=7.8',$ $FOV_{map}=66.6^{\\circ},$ and $m_{max}=4,$ $\\ell_{max}=4000$ for PLANCK noise at $\\nu=143~{\\rm GHz}.$ }} \\label{fig:cl_kk2_planck_pure_noise} \\end{figure} In this paper we implemented a new estimator of the convergence field for the extraction of the lensing potential from CMB maps. Our interest was in reconstructing the convergence field from CMB maps from which contaminants, such as point sources and the SZ effect, had previously been removed using multi-frequency information. The new estimator acts locally in real space, thus being able to treat the excision of pixels and nonuniform sky coverage in a flexible manner. We implemented the estimator on two experimental setups, one without and one with detector noise, based on the specifications of PLANCK for the $\\nu=143~{\\rm GHz}$ channel. For comparison of the performance of the new estimator, we also implemented the conventional estimator which acts in harmonic space. From the two experiments we learned that by increasing the range of the kernel in $\\ell$ space, and consequently reducing its range in $\\theta$ space, we improve the reconstruction of the power spectrum at small scales. The finite extent of the kernel is a desirable property of the proposed estimator, since in theory it allows the reconstruction of the lensing convergence, manifested essentially at very small scales, from a small map of the sky as long as the kernel probes sufficiently small angular scales. Even though there is a loss of power on scales smaller than the finite extent of the kernel, this effect could be studied using simulations to determine a form factor that could be applied to correct the loss of power \\cite{bucher2010}. We also observed that although the real space estimator is limited, in a statistical sense, to be as good as the harmonic space estimator, it has the advantage over the harmonic space estimator of being less sensitive to white noise. In practice, however, CMB experiments have to deal with correlated noise, which will pose as much of a challenge to the real space estimator as to the harmonic space estimator, most likely requiring the construction of a more carefully designed kernel in harmonic space that downweights the correlated modes before transforming to real space. On the basis of the results presented here, we envisage two follow-up studies. In the first study we intend to optimize the current implementation of the real space estimator to use a kernel capable of probing smaller angular scales % in the reconstruction of the lensing convergence, as we would expect for the ACT and SPT experiments. As we have demonstrated, the real space estimator can be applied to small patches of sky without incurring in the serious spectral leakage that affects the harmonic space estimator on rectangular, rather than torodial, domains in the presence of red spectra. In real space, the local filter acts up to a kernel length away from the edge of the map without leakage of power. We also intend to study the effect of a mask due to the excision of bad pixels, e.g. arising from the removal of bright point sources, that can cause aliasing of power from large scales to small scales. Inpainting has been proposed as a means of interpolating across masked regions that preserves the statistical properties of the map \\cite{inpaint}, so we plan to investigate how sensitive the real space estimator is to complex masks with and without the use of such techniques. A straightforward extension will be to develop the analogous estimator for the shear components of the convergence tensor and work out how to combine the different components for the reconstruction of the lensing potential. Since the shear components do not contain the $\\ell=0$ mode, the corresponding estimators will be less sensitive to a bias. The dilatation and shear components provide complementary information on the lensing potential but are not independent, being related by consistency conditions that are algebraic in harmonic space. Since these components probe the monopole and quadrupole sectors respectively, the noise associated with these components will be uncorrelated and the reconstruction can be harmonized, with the noise being reduced via inverse variance weighting of these components. This point is discussed in more detail in Ref.~\\cite{bucher2010}. In the second study we intend to extend the implementation of the real space estimator to CMB polarisation so as to optimize the reconstruction of the lensing potential from the PLANCK data. \\vspace{0.6cm} \\centerline{\\bf" }, "1005/1005.3527_arXiv.txt": { "abstract": "We report on the discovery and the timing analysis of the first eclipsing accretion-powered millisecond X-ray pulsar (AMXP): SWIFT~J1749.4--2807. The neutron star rotates at a frequency of $\\sim$517.9~Hz and is in a binary system with an orbital period of 8.8 hrs and a projected semi-major axis of $\\sim$1.90 lt-s. Assuming a neutron star between 0.8 and 2.2 $M_{\\odot}$ and using the mass function of the system and the eclipse half-angle, we constrain the mass of the companion and the inclination of the system to be in the $\\sim$0.46-0.81 M$_{\\odot}$ and $\\sim74.4^\\circ-77.3^\\circ$ range, respectively. To date, this is the tightest constraint on the orbital inclination of any AMXP. As in other AMXPs, the pulse profile shows harmonic content up to the 3rd overtone. However, this is the first AMXP to show a 1st overtone with rms amplitudes between $\\sim6$\\% and $\\sim23$\\%, which is the strongest ever seen, and which can be more than two times stronger than the fundamental. The fact that SWIFT~J1749.4--2807 is an eclipsing system which shows uncommonly strong harmonic content suggests that it might be the best source to date to set constraints on neutron star properties including compactness and geometry. ", "introduction": "\\label{sec:intro} The first accreting millisecond X-ray pulsar (hereafter AMXP) was discovered in 1998 \\citep[SAX~J1808.4--3658, see][]{Wijnands98} and since then, a total of 13 AMXPs have been found and studied in detail \\citep{Patruno10a}. Most AMXPs show near sinusoidal profiles during most of their outbursts. This is consistent with a picture in which only one of the hotspots (at the magnetic poles) is visible (see ref. below). Deviations from a sinusoidal profile (i.e, an increase in harmonic content) are generally interpreted as being caused by the antipodal spot becoming visible, perhaps as accretion rate falls and the disk retreats \\citep[see, e.g.][ and references therein]{Poutanen03,Ibragimov09}. Although the amplitude of the 1st overtone may reach that of the fundamental late in the outburst \\citep[see, e.g][]{Hartman08, Hartman09a}, no AMXP so far has shown pulse profiles where the 1st overtone is generally stronger than the fundamental throughout the outburst. The stability of the pulse profiles in some of the AMXPs means that pulse profile modeling can be used to set bounds on the compactness of the neutron star and hence the dense matter equation of state \\citep[see, e.g.][ and references therein]{Poutanen03,Poutanen09}. Unfortunately, there is often a large degeneracy between the parameters due to the number of free parameters needed to construct the model profile. One of these parameters is the inclination of the system, which to date has not been well-constrained for any AMXP. \\begin{figure}[http] \\center \\resizebox{1\\columnwidth}{!}{\\rotatebox{0}{\\includegraphics{./rmsvstime.eps}}} {\\footnotesize \\caption{\\textit{Top panel:} 2-10 keV flux as measured from RXTE/PCA and Swift/XRT observations. The flux of the last PCA observation (MJD 55307.5) is not shown; the spectrum of this observation was used as an estimate background emission (see text). Upper limits are quoted at 95\\% confidence level. We calculated the flux during the last PCA and last Swift/XRT detection using WebPIMMS (assuming a power law spectrum with index 1.8). \\textit{Middle panels:} Fractional rms amplitude and 95\\% confidence level upper limits of the fundamental and three overtones as a function of time. Detections ($>3\\sigma$ single trial) and upper limits are from $\\sim500$ and $\\sim3000$ sec datasets, respectively. \\textit{Bottom panel:} Ratio between the fractional rms amplitude of the 1st overtone and fundamental. Blue triangles represent points in which both harmonics are significantly detected in $\\approx500$ second datasets. Black circles represent the ratio between the fractional rms amplitude of the 1st overtone and the 95\\% confidence level upper limit to the amplitude of the fundamental. Grey circles represent the same ratio but when the fundamental is significantly detected and not the 1st overtone. This means that black circles represent lower limits while grey circles are upper limits. These ratios are independent of the background.}} \\label{fig:lc} \\end{figure} In this Letter we report on the discovery and timing of the accretion-powered millisecond X-ray pulsar SWIFT~J1749.4--2807. Thanks to the observed eclipses \\citep{Markwardt10}, we set the tightest constraint on system inclination for any AMXP. This, coupled with the fact that the amplitude of the first overtone is higher/comparable to that of the fundamental for much of the outburst and that the amplitude of the first overtone is unusually high, allows to put tight constraints on pulse profile models. We show that SWIFT~J1749.4--2807 has the potential to be one of the best sources for this approach to constraining the neutron star mass-radius relation and hence the EoS of dense matter. ", "conclusions": "" }, "1005/1005.0263_arXiv.txt": { "abstract": "It is shown that the Hubble constant can be derived from the standard luminosity function of galaxies as well as from a new luminosity function as deduced from the mass-luminosity relationship for galaxies. An analytical expression for the Hubble constant can be found from the maximum number of galaxies (in a given solid angle and flux) as a function of the redshift. A second analytical definition of the Hubble constant can be found from the redshift averaged over a given solid angle and flux. The analysis of two luminosity functions for galaxies brings to four the new definitions of the Hubble constant. The equation that regulates the Malmquist bias for galaxies is derived and as a consequence it is possible to extract a complete sample. The application of these new formulae to the data of the two-degree Field Galaxy Redshift Survey provides a Hubble constant of $( 65.26 \\pm 8.22 ) \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ for a redshift lower than 0.042. All the results are deduced in a Euclidean universe because the concept of space-time curvature is not necessary as well as in a static universe because two mechanisms for the redshift of galaxies alternative to the Doppler effect are invoked. ", "introduction": "The Hubble constant, in the following $H_0$, is defined as \\begin{equation} H_0 = \\frac{v}{D} [\\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}] \\quad , \\end{equation} where $v=cz$ is the recession velocity, $D$ is the distance in $Mpc$, $c$ is the velocity of light and $z$ is the redshift defined as \\begin{equation} z = \\frac { \\lambda_{obs} - \\lambda_{em} } { \\lambda_{em}} \\quad , \\end{equation} with $\\lambda_{obs}$ and $\\lambda_{em}$ denoting respectively the wavelengths of the observed and emitted lines as determined from the lab source. The first numerical values of the Hubble constant were : $H_{0}=625$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ as deduced by \\citet{Lemaitre1927}, $H_{0}=460$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ as deduced by \\citet{Robertson1928}, $H_{0}=500 $ $\\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ as deduced by \\citet{Hubble1929} and $H_{0}=290$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ as deduced by \\citet{Oort1931}. Figure~\\ref{hystory} reports the decrease of the numerical value of the Hubble constant from 1927 to 1980. \\begin{figure*} \\begin{center} \\includegraphics[width=10cm]{f01.eps} \\end {center} \\caption { Logarithmic values of the Hubble constant $H_0$ from 1927 to 1980. The error bar is evaluated according to the file http://www.cfa.harvard.edu/~huchra/hubble.plot.dat~. } \\label{hystory}% \\end{figure*} At the time of writing, two excellent reviews have been written, see \\citet{Tammann2006} $( H_{0}=(63.2\\pm1.3~ (random) ~\\pm5.3~ (systematic)) $ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1} ) $ and \\citet{Jackson2007} ($ H_0\\sim 70\\mbox{\\,--\\,}73$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1} $). We now report the methods that use the global properties of galaxies as indicators of distance: \\begin{enumerate} \\item Luminosity classes of spiral galaxies; $H_{0}=(55 \\pm 3) $ $\\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Sandage1999b} \\item 21 cm line widths; $H_{0}=(59.1 \\pm 2.5)$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Federspiel1999} \\item Brightest cluster galaxies; $H_{0}=(54.2 \\pm 5.4 )$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Sandage1973} \\item The D$_{n}$-$\\sigma$ or fundamental plane method; $H_{0}=(57 \\pm 4)$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Federspiel1999} \\item Surface brightness fluctuations; $H_{0}=71.8 $ $\\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Tammann2006} \\item Gravitational lens; $H_{0}=(72\\pm 12) $ $\\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Saha2006} \\item The Sunyaev--Zel'dovich effect; $H_{0}=(67\\pm 18)$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Udomprasert2004} \\item Ks-band Tully-Fisher Relation; $H_{0}=(84 \\pm 6 )$ $ \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$ , see \\citet{Russell2009}, where the Hubble constant was named Hubble parameter. \\end {enumerate} At the time of writing, the first important evaluation of the Hubble constant is through Cepheids (key programs with HST) and type Ia Supernovae, see \\citet{Sandage2006}, \\begin{equation} H_0 =(62.3 \\pm 5 ) \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1} \\quad . \\label{h0cefeidi} \\end {equation} A second important evaluation comes from the three years of observations with the Wilkinson Microwave Anisotropy Probe, see Table 2 in \\citet{Spergel2007}; \\begin{equation} H_{0}=(73.2 \\pm 3.2) \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1} \\quad . \\label{hzerowmap} \\end{equation} In the following, we will process galaxies having redshifts as given by the catalog of galaxies. The forthcoming analysis is based on two key assumptions: (i) the flux of radiation from galaxies in a given wavelength decreases with the square of the distance; (ii) the redshift is assumed to have a linear relationship with distance in $Mpc$. These two hypotheses allow some new physical mechanisms to be accepted which produce a linear relationship between redshift and distance, for redshifts lower than 1. In this framework, we can speak of a Euclidean universe because the distances are deduced from the Pythagorean theorem and a static universe because it is not expanding. The already listed approaches leave a series of questions unanswered or partially answered: \\begin {itemize} \\item Can the Hubble constant be deduced from the Schechter luminosity function of galaxies? \\item Can the Hubble constant be deduced from a new luminosity of galaxies alternative to the Schechter function? \\item Can the equation that regulates the Malmquist bias be derived in order to deal with a complete sample in apparent magnitude? \\item Can the reference magnitude of the sun be deduced from the luminosity function of galaxies? \\end{itemize} In order to answer these questions, Section~\\ref{secprelimaries} contains three introductory paragraphs on sample moments, the weighted mean and the determination of the so-called \"exact value\" of the Hubble constant. Section~\\ref{useful} reviews the basic system of magnitudes, a review of two alternative mechanisms for the redshift of galaxies, two analytical definitions of the Hubble constant in terms of the Schechter luminosity function of galaxies and two other definitions that can be found by adopting a new luminosity function for galaxies. Section~\\ref{secnumerical} contains a numerical evaluation of the four new formulae for the Hubble constant as deduced from the data of the two-degree Field Galaxy Redshift Survey. Section \\ref{secmsun} contains a numerical evaluation of the reference magnitude of the sun for a given catalog. ", "conclusions": "A careful study of the standard LF of galaxies allows the determination of the position of the maximum in the theoretical number of galaxies versus redshift and the theoretical averaged redshift. From the two previous analytical results, it is possible to extract two new formulae for the Hubble constant, equations~(\\ref{hzero1}) and (\\ref{hzero2}). The same procedure can be applied by analogy to a new LF as given by the mass-luminosity relationship, see equations~(\\ref{hzero3}) and (\\ref{hzero4}). The weighted mean of the four values of $H_0$ as deduced from Table~\\ref{hubblevalue} gives \\begin{equation} H_0 =( 65.26 \\pm 8.22 ) \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1} \\quad when~ z \\leq 0.042 \\nonumber \\quad . \\end {equation} This value lies between the value deduced from the Cepheids, see~\\citet{Sandage2006} and formula~(\\ref{h0cefeidi}) and the value deduced from WMAP, see~\\citet{Spergel2007} and formula~(\\ref{hzerowmap}). The developed framework also enables the deduction of the reference magnitude of the sun, see formula~(\\ref{msun}) and the application to the 2dFGRS gives \\begin{equation} M_{\\sun} = (5.5 \\pm 0.35) \\quad . \\end{equation} Assuming that the exact value is $ M_{\\sun}$ = 5.33, the efficiency in deriving the reference magnitude of the sun is $\\epsilon=96.63 ~\\% $ when $H_0=66.04 \\mathrm{\\ km\\ s}^{-1}\\mathrm{\\ Mpc}^{-1}$. We briefly review the basic cosmological assumptions adopted here to derive the Hubble constant: \\begin{itemize} \\item The mechanism that produces the redshift, here extracted from the catalog of galaxies, is not specified but we remember that the plasma redshift and DET (Dispersive Estinction Theory) do not produce a geocentric model for the universe as given by the Doppler shift, see \\citet{Wang2007}. \\item The number of galaxies as a function of redshift as well as the averaged redshift are evaluated in a Euclidean space or, in other words, the effects of space-curvature are ignored. \\item The spatial inhomogeneities present in the catalog of galaxies are partially neutralized by the operation of adding together the data of the south and north galactic pole of the 2dFGRS. The transition from a nonhomogeneous to a quasi-homogeneous universe is clear when Figure~\\ref{isto1} and Figure~\\ref{isto2} are carefully analyzed. \\item The initial assumptions of: (i) natural flux decreasing as given by equation~(\\ref{flr2}) ; (ii) linear relationship between redshift and distance which are present in the joint distribution in {\\it z} and {\\it f} for the number of galaxies are justified by the acceptable results obtained for the theoretical maximum in the number of galaxies, see Figure~\\ref{zeta_max_flux}. This fact allow us to speak of a Euclidean universe up to $z \\leq 0.042$. \\item The presence of the Malmquist bias does not allow to extrapolate the concept of a Euclidean, static universe for distances greater than $z\\,>\\,0.042$ when the 2dFGRS catalog is considered. \\end{itemize}" }, "1005/1005.4489_arXiv.txt": { "abstract": "By observing mergers of compact objects, future gravity wave experiments would measure the luminosity distance to a large number of sources to a high precision but not their redshifts. Given the directional sensitivity of an experiment, a fraction of such sources (gold plated -- GP) can be identified optically as single objects in the direction of the source. We show that if an approximate distance-redshift relation is known then it is possible to statistically resolve those sources that have multiple galaxies in the beam. We study the feasibility of using gold plated sources to iteratively resolve the unresolved sources, obtain the self-calibrated best possible distance-redshift relation and provide an analytical expression for the accuracy achievable. We derive lower limit on the total number of sources that is needed to achieve this accuracy through self-calibration. We show that this limit depends exponentially on the beam width and give estimates for various experimental parameters representative of future gravitational wave experiments DECIGO and BBO. ", "introduction": " ", "conclusions": "" }, "1005/1005.3056_arXiv.txt": { "abstract": "{Passive early-type galaxies (ETGs) provide an ideal laboratory for studying the interplay between dust formation around evolved stars and its subsequent destruction in a hot gas. Using {\\it Spitzer}-IRS and {\\it Herschel} data we compare the dust production rate in the envelopes of evolved AGB stars with a constraint on the total dust mass. Early-type galaxies which appear to be truly passively evolving are not detected by {\\it Herschel}. We thus derive a distance independent upper limit to the dust grain survival time in the hostile environment of ETGs of $< 46\\pm 25\\;\\rm Myr$ for amorphous silicate grains. This implies that ETGs which \\emph{are} detected at far-infrared wavelengths have acquired a cool dusty medium via interaction. Given likely time-scales for ram-pressure stripping, this also implies that only galaxies with dust in a cool (atomic) medium can release dust into the intra-cluster medium. } ", "introduction": "In this paper we compare the dust outflow rate from the evolved AGB stars in passive ETGs, with the total dust mass, to estimate dust grain lifetimes. Both quantities are derived from observations; the outflow rate from {\\it Spitzer}-IRS spectra, and the total dust mass from {\\it Herschel} (Pilbratt et al. 2010) maps of the {\\it Herschel} Virgo Cluster Survey, HeViCS (Davies et al. 2010, www.hevics.org). We assume that evolved stars are the only source of dust in these passive systems as there is very little evidence that type Ia supernovae produce dust. IRAS data have been used in the past in a similar way. Soifer et al. (1986) estimated the mass-loss from evolved stars in the bulge of M~31 from 12 and $25\\;\\rm \\mu m$ fluxes and the total dust mass from the 60 and $100\\;\\rm \\mu m$ fluxes. They concluded that given that the observed dust mass could be produced by stellar outflows in only $10^7\\;\\rm yr$, there must be some mechanism that depletes the bulge of inter-stellar matter. Jura et al. (1987) made similar calculations for a sample of elliptical galaxies. They estimated dust outflow rates by assuming that half of the $12\\;\\rm \\mu m$ flux came from dusty stellar envelopes. They also found that the stellar mass-loss rate would be sufficient to produce the observed cool ISM in much less than a Hubble time. Dust mass production rate and total dust mass can only reasonably be compared in objects that are truly passively evolving. Temi et al. (2007) find, that for a given blue luminosity, the 70 and $160\\;\\rm \\mu m$ luminosities vary by 2 orders of magnitude, indicating that many ETGs have a significant dust component not directly attributable to mass-loss from evolved stars. For our purposes, these ETGs are not `passive' because other processes (mergers?) have probably contributed to the dust mass. Recent studies of ETGs with {\\it Spitzer} (Bressan et al. 2006; Clemens et al. 2009) have shown that even in samples of ETGs selected to be the most passive objects (ie. lying on the colour-magnitude relation) a significant fraction show evidence of either on-going or recent past star formation. These results indicate that ETGs in which dust features are seen in the optical (e.g. Sadler \\& Gerhard, 1985) and many that have been detected by IRAS at 60 and $100\\;\\rm \\mu m$ (Knapp et al. 1989) and {\\it Spitzer} at 70 and $160\\;\\rm \\mu m$ (Kaneda et al. 2007) may also host low levels of star formation. Intriguigly, of the 7 ellipticals detected at 70 and $160\\;\\rm \\mu m$ by Kaneda et al. (2007) the only object not showing PAH emission features is actually a radio source, so that the far-infrared emission is probably synchrotron! Tsai and Mathews (1995) studied dust destruction in ETGs via thermal sputtering in the hot ($10^6 - 10^7$\\,K), low density ($n_H \\sim 0.1$\\,cm$^{-3}$) gas, and found that the destruction time-scale is short compared to any cooling flow or dust transport time-scale. They therefore concluded that dust is destroyed `on the spot' before it has time to migrate within a galaxy. In a following work, Tsai and Mathews (1996) found dust-to-gas mass ratios that are orders of magnitude less than that in typical stellar ejecta or in the ISM of the Milky Way. Their interpretation of the IRAS data was, at that time, hampered by the resolution of the existing observations; they were therefore not able to tie down the origin and distribution of the observed dust emission. As part of the {\\it Herschel} Science Demonstration Phase (SDP), a $4\\times4\\;\\rm deg^2$ field, centred approximately on M~87, has been observed as part of the HeViCS with both the PACS (Poglitsch et al. 2010) and SPIRE (Griffin et al. 2010) instruments. See Davies et al. (2010) for details. ", "conclusions": "We interpret the ratio of dust mass to dust production rate as a measure of grain lifetime. In principle, however, dust could be removed from the galaxy rather than destroyed. A wind of hot gas produced by type Ia supernovae probably cannot drive dust (or gas) out of the potential well of massive ETGs (Mathews \\& Loewenstein 1986). However, good evidence of ram-pressure stripping of the hot ISM in elliptical galaxies has been found in the form of X-ray tails (Randall et al. 2008; Machacek et al. 2006). If dust is found in this hot gas then it too may be removed. However, the timescale for such stripping is estimated to be of the order a few $10^8\\;\\rm yr$ (Takeda et al. 1984; Murakami \\& Babel 1999, eq. 12) which is an order of magnitude longer than our upper limit to the grain lifetime. The observation that the X-ray emission of most ETGs is \\emph{not} displaced from the stellar distribution, in fact, suggestes that either the stripping time-scale is longer than the hot gas production time-scale or that the ram-pressure is typically insufficient to remove the hot medium. It is now clear that if dust is isolated from the hot medium in denser, cooler clouds, than ram-pressure does have time to strip dust (Cortese et al. 2010) but the passive ETGs in our study do not show any evidence of accumulating dust in such clouds. \\emph{Passive} ETGs do not directly pollute the intra-cluster medium with dust. Grain lifetimes of $<46\\;\\rm Myr$ have immediate implications for ETGs that \\emph{are} detected in the FIR. Either they produce dust at rates much higher than those given in Table~\\ref{table:1} or their dust is shielded from the hot gas in cool clouds. For example, the dust mass of NGC~4435 of $1.2 \\times 10^{6}\\;\\rm M_{\\odot}$ would require a total dust production rate of $0.03\\;\\rm M_{\\odot}\\,yr^{-1}$, more than an order of magnitude greater than any value in Table~\\ref{table:1}. This object is known to host star formation (Panuzzo et al., 2007) and so dust is likely to be in cool clouds more typical of late-type galaxies; grain lifetimes in this case may be much longer. Dust may have been acquired from an interaction, but in any case it probably arrived as part of a cooler medium. In a passive ETG, the ISM, and the dust within it, come mainly from mass-loss from evolved stars. Unless the dust-to-gas ratio in this released material is very variable, one would expect dust and gas production to follow one another. If the X-ray luminosity, $L_{x}\\propto n_{e}^2$ one would naively expect a relation $L_{x}\\propto M_{\\rm d}^2$. However, if the grain lifetime, $t_{\\rm g}\\propto 1/n_{\\rm H}$, then we would expect only a linear correlation between $L_{x}$ and $M_{\\rm d}$. Kaneda et al. (2007) find evidence of an \\emph{anti-correlation} for a small sample of elliptical galaxies detected by {\\it Spitzer}-MIPS, but this may be expected if the dust were of external origin. Only NGC~4371 is detected by MIPS at $70\\;\\rm \\mu m$ (marginally at $160\\;\\rm \\mu m$). The detection is consistent with our {\\it Herschel} detection limits. This object actually shows extremely weak PAH features in its {\\it Spitzer}-IRS spectrum, and is therefore probably not totally passive. The object is included for comparison, and its IRS spectrum included in Fig.~\\ref{fig:1}. We have assumed that AGB stars are the only source of dust. Were type Ia supernovae able to produce dust from the metals they produce, we estimate that the contribution to the dust production rate would be, at most, similar to that of AGB stars. Our limit on grain lifetime would be a factor of $\\sim$2 lower in this case. Although dust grains are rather short-lived in ETGs, the prospects for survival in the intra-cluster medium are rather better because the gas densities are typically at least 2 orders of magnitude lower than in ETGs (B{\\\"o}hringer et al., 1994). Therefore, if dust can be removed from cluster galaxies, or form in the intra-cluster medium, there is a good chance that such dust will be detected by the HeViCS, especially at larger cluster radii where gas densities are lower." }, "1005/1005.1924_arXiv.txt": { "abstract": "Intergalactic magnetic fields (IGMF) can cause the appearance of halos around the gamma-ray images of distant objects because an electromagnetic cascade initiated by a high-energy gamma-ray interaction with the photon background is broadened by magnetic deflections. We report evidence of such gamma-ray halos in the stacked images of the 170 brightest active galactic nuclei (AGN) in the 11-month source catalog of the \\textit{Fermi} Gamma-Ray Space Telescope. Excess over point spread function in the surface brightness profile is statistically significant at $3.5\\sigma$ (99.95\\% confidence level), for the nearby, hard population of AGN. The halo size and brightness are consistent with IGMF, $B_{\\rm IGMF}\\approx 10^{-15} ~ {\\rm G}$. The knowledge of IGMF will facilitate the future gamma-ray and charged-particle astronomy. Furthermore, since IGMF are likely to originate from the primordial seed fields created shortly after the Big Bang, this potentially opens a new window on the origin of cosmological magnetic fields, inflation, and the phase transitions in the early Universe. ", "introduction": "Intergalactic magnetic fields (IGMF) had not been measured until now, despite their importance for gamma-ray and cosmic-ray astronomy and their likely connection to the primordial fields that could have seeded the stronger magnetic fields observed in galaxies, Sun, and Earth. This is because IGMF are too small for conventional astronomical probes, such as Zeeman splitting or Faraday rotation. Unlike the fields in galaxies, which are believed to have been amplified by the dynamo action of the large-scale convective motions of gas, the fields in voids remain low, close to their primordial values modified only by the relatively small contribution of the fields leaking out of galaxies \\citep{Kronberg1994,Grasso2001,Widrow2002,Kulsrud2007}. The observational and theoretical upper bounds on IGMF constrain their magnitudes to be below $10^{-9}$~G \\citep*{1997PhRvL..78.3610B}, whereas any value above $\\sim$10$^{-30}$~G is sufficient to explain the $\\sim \\mu$G Galactic magnetic fields generation by the dynamo mechanism \\citep*{1999PhRvD..60b1301D}. One can detect such extremely weak fields using high-energy gamma rays \\citep*{Aharonian1994,Plaga1995}. Very energetic photons emitted from active galactic nuclei (AGN) or other strong sources produce pairs of electrons and positrons in their interactions with the extragalactic background light (EBL). These pairs up-scatter the cosmic microwave background photons to high energies, giving rise to an electromagnetic cascade, and the photons from the cascade are detected by gamma-ray telescopes, such as {\\it Fermi}. Since the trajectories of electrons and positrons in the cascade are affected by magnetic fields, a gamma-ray image of AGN is expected to exhibit a halo of secondary photons around a bright central point-like source \\citep{Aharonian1994,Dolag2009,Neronov2009}. The central image is expected to be composed of photons emitted directly from the source with energies below the pair production threshold. In addition, delays in arrival times of the secondary photons can be used to probe IGMF \\citep{Plaga1995,Ando2004,Murase2008}. Finally, at TeV energies, the secondary photons produced in interactions of cosmic rays with EBL may have already been observed by the air Cherenkov telescopes \\citep{Essey2010a,Essey2010b}. Thus far, in TeV range, HEGRA \\citep{HEGRA} and MAGIC \\citep{MAGIC} did not detect any halo component of two bright blazars, Mrk~501 and Mrk~421, and they set upper limits on the flux. In particular, the analysis of MAGIC using gamma rays above 300~GeV excludes some range of IGMF between $4\\times 10^{-15}$ and $10^{-14}$~G. Very recently, IGMF above $3\\times 10^{-16}$~G were proposed as an explanation of non-observation by {\\it Fermi} of several AGN known to be bright TeV sources (\\citealt{Neronov2010}; see also \\citealt{2010MNRAS.tmpL..82T}). In this {\\it Letter}, we present evidence of extended images and of IGMF at $3.5\\sigma$ level, based on gamma-ray data collected by the Large Area Telescope (LAT) onboard {\\it Fermi}, in the energy range between 1~GeV and 100~GeV. It is consistent with pair-halo scenario with IGMF, $B_{\\rm IGMF} \\approx 10^{-15} ~ {\\rm G}$. The knowledge on IGMF will facilitate the future gamma-ray and cosmic ray astronomy, and it will open a new window on the origin of cosmological magnetic fields \\citep{Cornwall:1997ms}, inflation \\citep{Turner:1987bw,DiazGil:2007dy}, and the phase transitions in the early Universe \\citep{Vachaspati:1991nm,Baym:1995fk,Vachaspati:2001nb}. ", "conclusions": "" }, "1005/1005.3921_arXiv.txt": { "abstract": "We use a suite of cosmological, hydrodynamical simulations to investigate the chemical enrichment history of the Universe. Specifically, we trace the origin of the metals back in time to investigate when various gas phases were enriched and by what halo masses. We find that the age of the metals decreases strongly with the density of the gas in which they end up. At least half of the metals that reside in the diffuse intergalactic medium (IGM) at redshift zero (two) were ejected from galaxies above redshift two (three). The mass of the haloes that last contained the metals increases rapidly with the gas density. More than half of the mass in intergalactic metals was ejected by haloes with total masses less than $10^{11}\\,\\Msun$ and stellar masses less than $10^9\\,\\Msun$. The range of halo masses that contributes to the enrichment is wider for the hotter part of the IGM. By combining the `when' and `by what' aspects of the enrichment history, we show that metals residing in lower density gas were typically ejected earlier and by lower mass haloes. ", "introduction": "Despite the fact that `metals' - heavy elements produced by stars - make up a very tiny portion of the cosmic matter budget, they are of critical importance to our understanding of the Universe. From a diagnostic point of view, they are very useful in constraining star formation and tracing feedback from massive stars and supernovae. More importantly, however, metals impact the rate at which gas cools, affecting structure formation on scales of galaxy clusters down to dust grains. This effect should not go understated since, depending on the temperature, metals can reduce the cooling time by over an order of magnitude for solar metallicity \\cite[e.g.][]{Sutherland1993, Gnat2007, Wiersma2009a}. Observations clearly show that a substantial fraction of the diffuse intergalactic medium (IGM) has been enriched with heavy elements \\citep[e.g.][]{Cowie1998,Schaye2000a,Ellison2000,Schaye2003,Simcoe2004,Aracil2004,Schaye2007,Aguirre2008,Danforth2008,Cooksey2010,Pieri2010}. However, the physical mechanism, the timing and the sources of the enrichment all remain unknown. The IGM may have been enriched by the first generations of stars and galaxies at very high redshifts, or it may have been polluted by more massive galaxies at intermediate redshifts. Metals may be carried out into intergalactic space by galactic winds driven by supernovae, radiation pressure from star light, or by active galactic nuclei (AGN) \\citep[e.g.][]{Cen1999a,Aguirre2001,Aguirre2001z3,Madau2001,Theuns2002,Tornatore2004,Aguirre2005,Oppenheimer2006,Dave2007,Oppenheimer2008,Scannapieco2006b,Kawata2007,Kobayashi2007,Sommer-Larsen2008,Wiersma2009b,Shen2009,Tornatore2010,Choi2010}. In dense environments such as groups and clusters of galaxies metals may also be efficiently dispersed via processes such as ram pressure stripping or tidal stripping (see e.g.\\ \\citealt{Borgani2008} for a review). The most straightforward way to investigate when the IGM received its metals, is to plot the metallicity or metal mass as a function of redshift, as we did for a series of cosmological simulations in \\citet{Wiersma2009b}. By comparing models with different physical implementations, one can obtain information about the processes that are important for IGM enrichment. However, the metals that reside in the IGM at high redshift may not be the same metals that are in the IGM at the present day. What is hot IGM gas at high redshift, may have cooled by low redshift. Intergalactic metals may also accrete onto galaxies and end up in the ISM or in stars. Here we also use cosmological simulations to study the enrichment of the cosmos - noting that the IGM is of particular interest - but we take a novel approach in that we classify the gas based on its state at the redshift of interest (we choose $z=0$ and $z=2$) and trace its enrichment history back in time with the aim of understanding how the observed metals were put into place. Specifically, we will address the following two questions: When did gas that resides in the IGM at $z=0$ and $z=2$ receive its metals? And what are the masses of the haloes that last contained these metals? This paper is organised as follows. In Section~\\ref{sec-method} we describe the simulations used, emphasising the details that are particularly pertinent here. Section~\\ref{sec-when} discusses \\textit{when} the gas was enriched, taking each sub-phase in turn. In Section~\\ref{sec-what} we investigate \\textit{what} enriched the gas, using a halo-finder to determine what kinds of objects are mainly responsible for the enrichment of various gas phases. We connect the two ideas in Section~\\ref{sec-conn}, relating enrichment time with the last halo that contained the metals that end up in the IGM. Finally, we summarise our findings in Section~\\ref{sec-summary}. ", "conclusions": "\\label{sec-summary} We have used cosmological, hydrodynamical simulations that include radiative cooling, star formation, stellar evolution and supernova feedback, and, in some cases, AGN feedback, to investigate the metal enrichment history of the Universe. Specifically, we took advantage of the Lagrangian nature of our simulations to investigate when gas that ends up in a particular gas phase at redshift $z=0$ or $z=2$ received its metals and what the masses were of the haloes that last contained these metals. We considered different physical models, all taken from the OWLS project, including simulations with a different implementation of supernova feedback and a model that includes AGN feedback. While there are some small differences, such as slightly higher enrichment redshifts for models with more efficient feedback in massive galaxies, the main trends are strikingly similar for all models. We cannot rule out that some of our conclusions may depend on our prescriptions for feedback or that they are specific to SPH. However, the fact that three drastically different feedback models give such similar results suggests that our conclusions are robust. The time since the enrichment varies most strongly with the gas density. Metals in lower density gas are typically much older than metals in high-density gas, a trend that extends over ten orders of magnitude in gas density. At least half of the metals that reside in the diffuse IGM at $z=0$ ($z=2$) were ejected above redshift two (three). The enrichment redshift also varies with the temperature of the gas, but this mostly reflects the fact that the typical gas density of enriched gas varies with its temperature. For gas that is dense enough to be able to cool, the time since enrichment correlates well with the cooling time. Gas with shorter cooling times received its metals more recently. The typical mass of the haloes from which the metals residing in the IGM were ejected, increases rapidly with the gas density. At least half of the metal mass was ejected by haloes with total masses less than $10^{11}\\,\\Msun$, which corresponds to stellar masses smaller than $10^9\\,\\Msun$. For the low-redshift IGM the mass of the dominant IGM polluters may be substantially smaller than this because our predictions for present day low-density gas are not close to converged with respect to the numerical resolution. The age of the metals ending up in the diffuse IGM is strongly anti-correlated with the mass of the haloes from which they were ejected. In other words, older metals were typically ejected by lower mass haloes. This anti-correlation is less strong for the hotter part of the IGM (the WHIM), for which recent enrichment by low-mass haloes is more important. While massive haloes ($> 10^{12}\\,\\Msun$) are unimportant for the enrichment of the diffuse IGM, they do contribute to the pollution of the warm-hot IGM at low redshift. Our main and strongest conclusion is that metals residing in lower density gas were typically ejected earlier and by lower mass haloes. This suggests that travel-time is a limiting factor for the enrichment of the IGM, as proposed by \\citet{Aguirre2001z3}. Metals that have been ejected recently simply have not had sufficient time to reach the low-density gas far from galaxies. The anti-correlation between metal age and halo mass would then follow because typical galaxy masses are lower at higher redshifts. Another mechanism that could explain the trends is fall back onto galaxies. If metals ejected by lower mass galaxies are less likely to fall back \\cite[e.g.][]{Oppenheimer2010} or if they typically leave the haloes with larger velocities, then we would expect lower mass galaxies to be more important for gas that is further away. Finally, low-mass galaxies are more weakly clustered than high-mass galaxies, thus low-density gas is more accessible to low-mass galaxies. It is not immediately obvious how one could compare this result with observations. One way would be to consider abundance ratios of elements in various phases as a function of density and redshift in an attempt to find a signature that could be searched for in high-redshift observations. However, uncertainties in the nucleosynthetic yields and the type Ia supernova rates \\citep[see][]{Wiersma2009b}, as well as in the ionization corrections, may make such a comparison troublesome. Correlations between absorbers and the distances to different types of galaxies may also help shed light on the enrichment mechanism \\citep[e.g.][]{Steidel2010,Chen2009,Wilman2007,Stocke2006}, although is is important to keep in mind that the metals that are observed to be near a galaxy may have been injected by its lower-mass progenitors \\citep{Porciani2005,Scannapieco2005}. In fact, our results would imply that this is likely to be the case for metals in low-density gas. We leave an analysis of abundance ratios and other observational comparisons to a future work. The importance of low-mass galaxies is a challenge for simulations of the enrichment of the IGM, as it implies the need for high-resolution in large volumes. On the other hand, our results suggest that the low-density IGM provides us with an exciting opportunity to study the consequences of feedback in low-mass galaxies, a key but poorly understood ingredient of models of galaxy formation, and that it provides us with a fossil record of galaxy formation in the high-redshift Universe." }, "1005/1005.0958_arXiv.txt": { "abstract": "Quasi-equilibrium sequences of binary neutron stars are constructed for a variety of equations of state in general relativity. Einstein's constraint equations in the Isenberg-Wilson-Mathews approximation are solved together with the relativistic equations of hydrostationary equilibrium under the assumption of irrotational flow. We focus on unequal-mass sequences as well as equal-mass sequences, and compare those results. We investigate the behavior of the binding energy and total angular momentum along a quasi-equilibrium sequence, the endpoint of sequences, and the orbital angular velocity as a function of time, changing the mass ratio, the total mass of the binary system, and the equation of state of a neutron star. It is found that the orbital angular velocity at the mass-shedding limit can be determined by an empirical formula derived from an analytic estimation. We also provide tables for 160 sequences which will be useful as a guideline of numerical simulations for the inspiral and merger performed in the near future. ", "introduction": "Coalescing binary neutron stars are among the most promising sources of gravitational waves for ground-based laser-interferometric gravitational-wave detectors such as LIGO \\citep{bro04}, GEO600 \\citep{luc06}, TAMA300 \\citep{and05}, and VIRGO \\citep{ace07}. Merger of binary neutron stars, together with that of black hole-neutron star binaries, is also considered to be one of the candidates for the central engines of short-hard gamma-ray bursts \\citep{nar92}. These facts motivate us to study coalescing binary neutron stars. Binary neutron stars evolve as a result of gravitational radiation reaction and eventually merge. This evolutionary sequence is usually divided into three stages, depending on the characteristic timescales associated with orbital period and gravitational radiation reaction, as well as on the tidal effects for each neutron star. The first stage is the adiabatic, inspiral phase. In this phase, the timescale of orbital shrink due to the emission of gravitational waves is much longer than the orbital period, and thus, the binary system evolves adiabatically. In addition, each neutron star can be treated as a point mass, because the orbital separation is much larger than the neutron star radius and hence the tidal deformation of the neutron star is negligible. In this phase, a post-Newtonian approximation together with the point particle approximation is a robust tool for determining the orbital evolution and for computing gravitational waveforms (see e.g. \\citet{bla06} and references therein). The second stage is called the intermediate phase or the quasi-equilibrium phase. In this phase, the binary system is considered to be still in the adiabatic, inspiral phase, but we need to take into account tidal effects on each neutron star, i.e., hydrodynamic effects in neutron stars, as well as full effects of general relativity, because the orbital separation between two neutron stars is only a few times of the neutron-star radius and thus they are in a strong two-body gravitational field. One of the aims of the present paper is to contribute to the understanding of this phase. We will explain more details about the purpose of the present paper later. The last stage is the merger phase, for which the timescale of orbital shrink becomes shorter than the orbital period and thus the evolution of the system proceeds in a dynamical manner. Furthermore, the system becomes highly general relativistic, because the compactness of the system, defined by the ratio of the gravitational radius to the radius of the system, becomes larger than $\\sim 0.2$. To clarify the merger phase, numerical relativity is the unique approach. Since the first fully general relativistic merger simulation was performed by \\citet{shi99}, huge effort has been devoted in this research field \\citep{shi00,shi02,shi03,shi05,shi06a,due03,mil04,and08a,and08b,yam08,liu08,bai08,gia09,kiu09,kiu10}. (See e.g. \\citet{oec07a} and \\citet{oec07b} for works focusing on the micro-physics in a neutron star but not in fully general relativistic framework.) Now we return to the intermediate phase and explain the purposes for studying the quasi-equilibrium phase of binary neutron stars in general relativity in detail. There are two roles for this study. One is to clarify the physical conditions in this phase, for example, how large the tidal deformation of a neutron star is, when the mass-shedding from the neutron star occurs, and what the orbital angular velocity at the mass-shedding limit is. The other is to provide initial data for studying the merger phase by numerical relativity simulations. Numerical relativity, in which Einstein's evolution equations are solved, requires initial data that satisfy Einstein's constraint equations and also that should be as physical as possible. Obviously, it is important to derive accurate initial data for a scientific study. Constructing such initial data is just obtaining relativistic binary neutron stars in quasi-equilibrium. The first effort on this issue was devoted to constructing corotating binary neutron stars in general relativity because implementing a numerical code for computing such solutions is relatively easy \\citep{bau97,bau98,mar98,usu00,usu02,tan02b,tan03,tic09}. \\citet{koc92} and \\citet{bil92}, however, found that the timescale of coalescence driven by the gravitational-radiation reaction is much shorter than that of synchronization due to the viscosity in a neutron star. This implies that if the spin angular velocity of neutron stars is much smaller than the orbital angular velocity in a late inspiral phase, we can regard the rotation state of a neutron star to be approximately irrotational for the subsequent phase until the merger sets in. Additionally, any neutron star spins down due to electromagnetic radiation during its life from birth to the coalescence. The spin down timescale of a neutron star in a known binary is at longest as short as the coalescing timescale \\citep[larger than $\\sim 10^8$ yrs;][]{lor08}. Moreover, the spin period of neutron stars in a known binary is always larger than 20 ms which is $\\sim 10$ times larger than the orbital period in the late inspiral phase just prior to merger, 2--3 ms. Therefore, we can conclude that the irrotational flow is physically more realistic.\\footnote[2]{If a first born neutron star in a binary was strongly recycled during the evolution of the companion, it may result in fast rotation and weak magnetic field. In such a case, the neutron star may rotate on the order of milliseconds even just before the merger. Some effort to approximately construct quasi-equilibrium, non-corotating, and non-irrotational binary systems is reported in \\citet{mar03} and \\citet{bau09}.} Under the assumption of irrotation, formulation for solving relativistic hydrostatic equations was derived \\citep{bon97,asa98,shi98,teu98}. Soon after the formulation was derived, quasi-equilibrium sequences of irrotational binary neutron stars were calculated \\citep{bon99,mar99,ury00a,ury00b,gou01,tan02b,tan03,bej05}, based on the Isenberg-Wilson-Mathews (IWM) approximation to general relativity \\citep{ise78,ise08,wil89}. (See \\citet{shi04} for an advanced formulation and \\citet{ury06,ury09} for the results.) Even though a lot of sequences have been, so far, calculated, systematic survey has not yet been done. In particular, {\\it unequal-mass}, irrotational binary neutron stars with an {\\it equation of state other than single polytrope} has not been studied in detail. In \\citet{tan02b,tan03}, quasi-equilibrium sequences of unequal-mass binaries were calculated, but a polytropic equation of state (EOS) was used. In \\citet{bej05} and \\citet{ury09}, non-polytropic EOSs were used, but quasi-equilibrium sequences of non-equal-mass binaries were not computed. Actually, unequal-mass, irrotational binary neutron stars with realistic EOSs in quasi-circular orbits were constructed and used initial data for merger simulations in \\citet{shi05}, \\citet{shi06a}, and \\citet{kiu09,kiu10}.\\footnote[3]{We use the term ``realistic equations of state'' for the EOSs derived from nuclear physics, although no one really knows a realistic one.} We, however, have not constructed sequences, and rather computed only some initial data sets for each neutron star mass. The purpose of the present paper is to complete the issue and to provide a database of the sequences. To compute unequal-mass binary systems of arbitrary mass ratio, we develop a new code for the present research, because the numerical code that we developed for the previous works \\citep{tan02b,tan03} had a problem with calculating binary systems composed of significantly different-mass neutron stars in general relativity even though the problem was not in Newtonian computation \\citep{tan02a}. As we will explain in Section 2, the method to determine the center of mass of unequal-mass binary systems, i.e., the position of the rotation axis, caused the problem in the previous code, but we have overcome it by employing a new method. In addition, we employ a wide variety of EOSs; piecewise polytropic EOSs \\citep{rea09a,rea09b}, tabulated realistic EOSs derived from nuclear physics, and fitted EOSs to the tabulated realistic EOSs. Some of the first and second EOSs were, respectively, used in \\citet{ury09} and in \\citet{bej05}, but we adopt a wider set of EOSs in this paper. Furthermore, we systematically study the unequal-mass binaries, whereas \\citet{bej05} and \\citet{ury09} focused only on the equal-mass case. This paper is organized as follows. We briefly review the basic equations and explain the improvement of the numerical code in Section 2. In Section 3, the results for the code test are shown. In Section 4, we show numerical results and discuss the effects of EOS on each sequence. Section 5 is devoted to a summary. Throughout this paper we adopt geometrized units with $G=c=1$, where $G$ denotes the gravitational constant and $c$ the speed of light. Latin and Greek indices denote purely spatial and spacetime components, respectively. ", "conclusions": "In the present paper, quasi-equilibrium sequences of binary neutron stars are constructed for 18 EOSs except for the $\\Gamma=2$ polytrope, in the IWM framework in general relativity. The EOSs we choose are nine piecewise polytropes, six tabulated realistic EOSs derived by using various theories of dense nuclear matter and different solution methods of the many-body problem in nuclear physics, and three EOSs expressed by a fitting formula. We employ three total masses, $M_0=2.4 M_{\\odot}$, $2.7 M_{\\odot}$, and $3.0 M_{\\odot}$ for each EOS, and compute sequences of three mass ratios for each total mass: i.e., $M_{\\rm ADM}^{\\rm NS1}$ versus $M_{\\rm ADM}^{\\rm NS2}$ $=1.20 M_{\\odot}$ versus $1.20 M_{\\odot}$, $1.10 M_{\\odot}$ versus $1.30 M_{\\odot}$, and $1.00 M_{\\odot}$ versus $1.40 M_{\\odot}$ for $M_0=2.4 M_{\\odot}$; $1.35 M_{\\odot}$ versus $1.35 M_{\\odot}$, $1.25 M_{\\odot}$ versus $1.45 M_{\\odot}$, and $1.15 M_{\\odot}$ versus $1.55 M_{\\odot}$ for $M_0=2.7 M_{\\odot}$; and $1.50 M_{\\odot}$ versus $1.50 M_{\\odot}$, $1.40 M_{\\odot}$ versus $1.60 M_{\\odot}$, and $1.30 M_{\\odot}$ versus $1.70 M_{\\odot}$ for $M_0=3.0 M_{\\odot}$. We focus on the unequal-mass sequences and compare their results with those of the equal-mass case. Changing the mass ratio, the total mass, and the EOSs, we investigate the behavior of the binding energy and total angular momentum along a sequence, the endpoint of sequences, and the orbital angular velocity as a function of time. For example, it is found for the piecewise polytropic EOSs that the change in stellar radius fixing the stiffness of the core EOS makes the orbital angular velocity at the mass-shedding limit vary widely, while the change in stiffness of the core EOS fixing the stellar radius does not change the orbital angular velocity at the mass-shedding limit significantly. It is also found that the orbital angular velocity at the closest separation decreases as we decrease the mass ratio, $M_{\\rm ADM}^{\\rm NS1}/M_{\\rm ADM}^{\\rm NS2} \\le 1$. The reason is that the less massive star in an unequal-mass binary is tidally deformed by the companion more massive star and starts shedding mass at larger separation than that for the equal-mass case. It is found that the orbital angular velocity at the mass-shedding limit increases as the neutron-star mass increases. This is because a more massive star becomes more compact and more difficult to be tidally disrupted for the same EOS. This implies that the binary neutron stars with massive stars need to come closer than those with less massive stars for reaching the mass-shedding limit. The orbital angular velocity at the mass-shedding limit is analyzed by using a Newtonian argument, and an empirical formula is found as \\begin{equation} M_0 \\Omega_{\\rm ms} =0.270~ {\\cal C}_{\\rm NS1}^{3/2} \\Bigl( 1+\\frac{1}{q} \\Bigr)^{3/2} q^{1/2}. \\end{equation} We have provided tables for 160 sequences as shown in Appendix \\ref{app:sequence}. Those tables may be useful as one of the database for future works on binary neutron stars in quasi-equilibrium and as a guideline of numerical simulations for the inspiral and merger." }, "1005/1005.1390_arXiv.txt": { "abstract": "We present a new theoretical calculation of the contribution to the extragalactic gamma-ray background radiation (EGRB) from star-forming galaxies, based on a state-of-the-art model of hierarchical galaxy formation that is in quantitative agreement with a variety of observations of local and high-redshift galaxies. Gamma-ray luminosity ($L_\\gamma$) and spectrum of galaxies are related to star formation rate ($\\psi$), gas mass ($M_{\\rm gas}$), and star formation mode (quiescent or starburst) of model galaxies using latest observed data of nearby galaxies. We try the two limiting cases about gamma-ray production: the escape limit ($L_\\gamma \\propto \\psi M_{\\rm gas}$) and the calorimetric limit ($L_\\gamma \\propto \\psi$), and our standard model predicts 7 and 4\\% contribution from star-forming galaxies to the total EGRB flux (including bright resolved sources) recently reported by the {\\it Fermi} Gamma-Ray Space Telescope. Systematic uncertainties do not allow us to determine the EGRB flux better than by a factor of $\\sim$ 2. The predicted number of nearby galaxies detectable by {\\it Fermi} is consistent with the observation. Intergalactic absorption by pair-production attenuates the EGRB flux only by a modest factor of $\\sim$1.3 at the highest {\\it Fermi} energy band, and the reprocessed cascade emission does not significantly alter EGRB at lower photon energies. The sum of the known contributions from AGNs and star-forming galaxies can explain a large part of EGRB, with a remarkable agreement between the predicted model spectrum and observation. ", "introduction": "\\label{sec:INTRODUCTION} The existence of the extragalactic diffuse gamma-ray background (EGRB) has been revealed first by the SAS-2 satellite (\\citealt{1977ApJ...217L...9F}; \\citealt*{1978ApJ...222..833F}). Better determinations of the flux and spectrum of EGRB were achieved by the EGRET detector on board the Compton Gamma-Ray Observatory (\\citealt{1998ApJ...494..523S}; \\citealt*{2004ApJ...613..956S}). The most reliable measurement of EGRB has very recently been reported based on the data of the {\\it Fermi} Gamma-Ray Space Telescope (\\citealt{fermi}), and the EGRB spectrum is well described by a single power-law with a photon index of $2.41 \\pm 0.05$ and the photon flux of about $(1.03 \\pm 0.17) \\times 10^{-5} \\;{\\rm photons\\;cm^{-2}\\;s^{-1}\\;sr^{-1}}$ above 100 MeV (\\citealt{PhysRevLett.104.101101}). The origin of EGRB has been discussed for a long time and various sources have been discussed as possible contributors to EGRB, such as active galactic nuclei (AGNs, especially blazars), galaxy clusters, intergalactic shocks produced by structure formation, and dark matter annihilation (see, e.g., \\citealt{2007AIPC..921..122D} for a review). Almost all of the known extragalactic gamma-ray sources are blazars, and their contribution to the EGRB has been intensively studied (e.g., Inoue \\& Totani 2009; Venters 2010; Abdo et al. 2010d, and references therein). However, star-forming galaxies should also be gamma-ray emitters by cosmic-ray interactions with interstellar medium (ISM) and interstellar radiation field (ISRF) (\\citealt{2000ApJ...537..763S}; \\citealt{2004ApJ...613..962S}), and there must be non-zero contribution to EGRB. This is obvious because we know that the Galactic disk is a strong source of diffuse gamma-rays, and gamma-rays from Large Magellanic Cloud (LMC) have been detected by EGRET (\\citealt{1992ApJ...400L..67S}). Furthermore, gamma-ray emission in GeV--TeV from Small Magellanic Cloud (SMC, \\citealt{2010arXiv1008.2127T}) and two nearby starburst galaxies, M82 and NGC 253, have recently been discovered by H.E.S.S. (\\citealt{2009Sci...326.1080A}), VERITAS (\\citealt{2009Natur.462..770V}), and {\\it Fermi} (\\citealt{2010ApJ...709L.152A}). The purpose of this paper is to present a new estimate of the contribution from star-forming galaxies to EGRB. There are a number of previous studies on this issue (Strong et al. 1976; Lichti et al. 1978; Dar \\& Shaviv 1995; Pavlidou \\& Fields 2002; Thompson et al. 2007; Bhattacharya \\& Sreekumar 2009; Ando \\& Pavlidou 2009; Lacki et al. 2010; Fields et al. 2010), and the estimates of the contribution to EGRB ranges $\\sim$10--50\\%\\footnote{ Here, the contribution is against the total (physical) extragalactic background photon flux ($>$100 MeV) including bright resolved sources. It is often calculated against the unresolved component of EGRB, but it depends on the flux sensitivity of a particular detector. Throughout the paper, ``the contribution to EGRB'' is defined against the total EGRB flux, which we estimated from the resolved and unresolved components of the {\\it Fermi} data (Abdo et al. 2010d). We used a photon index of 2.4 for resolved blazars (Abdo et al. 2010d) to extrapolate the resolved component from 200 to 100 MeV.}. Most of these studies calculated gamma-ray luminosity ($L_\\gamma$) from star formation rate (SFR) of galaxies, because the cosmic-ray energy input is expected to be proportional to SFR. However, if escape of cosmic-rays from galaxies is significant, SFR cannot be used as a reliable indicator of $L_\\gamma$ (see \\S \\ref{sec:ModelingLg} in more details). Several studies used infrared luminosity of galaxies as SFR indicators, but IR luminosity (especially at far-IR in early-type galaxies) is not a perfect SFR estimator because a significant amount of dust can be heated by ISRF from relatively old stars (e.g., Salim et al. 2009 and references therein). Amount of interstellar gas should also be important to determine $L_\\gamma$, because the degree of cosmic-ray escape is affected by the target amount. In fact, recent observations indicate that gamma-ray luminosity is nicely correlated with the product of SFR and gas mass in galaxies(Abdo et al. 2010a). Pavlidou \\& Fields (2002) included gas mass in the theoretical prediction of EGRB using the data of the cosmic star formation history (CSFH). Galaxies in the universe is considered as a closed box containing stellar and gas mass in present-day galaxies, and CSFH is used to solve the time evolution of relative fractions of stars and gas. However this approach likely overestimates the gas mass contributing to gamma-ray production at high redshifts, because most of the present-day stellar mass is in the form of gas and assumed to contribute to gamma-ray emission. In reality, only the gas in collapsed dark halos can contribute to cosmic-ray interactions in galaxies, which is a small fraction at high redshifts according to the structure formation theory. Recently, Fields et al. (2010) incorporated gas mass in galaxies using the Schmidt-Kennicutt relation in their calculation of EGRB. In this case one must know galaxy size to relate SFR and $M_{\\rm gas}$, and a single value of galaxy size at each redshift was assumed as a simple model. Here we present a new calculation of the contribution from star-forming galaxies in EGRB using a state-of-the-art theoretical model of galaxy formation in the framework of hierarchical galaxy formation, which is in quantitative agreement with a variety of observations at high redshifts as well as the local universe. An advantage of our approach compared with previous studies is that we can calculate both SFR and gas mass of individual galaxies at various redshits. Furthermore, we utilize the information of the gamma-ray spectra recently observed for nearby starbursts, in addition to the standard spectrum of the Galactic diffuse emission, to predict the EGRB spectrum based on the galaxy formation model including both quiescent and starburst galaxy populations. The paper is organized as follows. In the next section we describe our model calculations. We present the results on EGRB and statistics about the number of nearby galaxies detectable by {\\it Fermi} in \\S \\ref{sec:RESULTS}. After discussion on the uncertainties in our calculation (\\S \\ref{sec:Uncertainties}) and implications for the origin of EGRB (\\S \\ref{sec:EGRB-origin}), conclusions will be presented in \\S \\ref{sec:Conclusions}. In this work, cosmological parameters of $\\Omega_{0}=0.3$, $\\Omega_{\\Lambda}=0.7$, and $H_{0}=70~{\\rm Mpc^{-1}~km~s^{-1}}$ are adopted. ", "conclusions": "\\label{sec:Conclusions} We have presented a new calculation of EGRB from cosmic-ray interactions in star-forming galaxies, based on a state-of-the-art galaxy formation model in the framework of hierarchical structure formation. The galaxy formation model is quantitatively consistent with various observations at high redshifts as well as the local universe. Gamma-ray luminosities of galaxies are calculated by star formation rate and gas mass in model galaxies, with the relation $L_\\gamma \\propto (\\psi M_{\\rm gas})^{0.86}$ (the escape limit) or $L_\\gamma \\propto (\\psi)^{1.2}$ (the calorimetric limit) which are calibrated by the recent observational data of nearby galaxies. The predicted number of nearby galaxies that are detectable by {\\it Fermi} is consistent with the actual number observed so far. We found that star-forming galaxies make 4--7\\% contribution to the total EGRB flux reported by {\\it Fermi} in our standard model. Combined with the contribution from blazars as estimated by the {\\it Fermi} data, more than $\\sim$50\\% of EGRB can be accounted for. If the soft power-law tail of CXB is extending from MeV to GeV region as expected from the MeV background data and theoretical considerations of AGN accretion disks (Inoue et al. 2008), additional $\\sim$20 \\% can be explained. The combined spectrum by AGNs and star-forming galaxies is remarkably similar to the observed EGRB spectrum. It should be noted that there is no free parameters that can be tuned to fit the obeserved EGRB spectrum in the present model; the blazar EGRB spectrum of IT09 is determined by the spectral templates of the blazar SED (spectral energy distribution) sequence, and that of the star-forming galaxy component in the present work by templates constructed by observed nearby galaxies. Therefore we conclude that a large part ($\\gtrsim$70\\%) of EGRB can be explained by reasonable sources of AGNs and star-forming galaxies. Further examination is required to see whether the apparent $\\lesssim$30\\% residual of EGRB is mainly a result of modeling uncertainties, experimental/observational uncertainties in deriving the EGRB data, or significant contributions from completely different source populations. Given the good spectral agreement of the {\\it Fermi} data and AGNs/star-forming galaxies, the rest of EGRB is also likely dominated by astrophysical sources accelerating particles, even if a completely diffrent population is responsible for it. It is in contrast that a large part of the blazar component of EGRB will be resolved into discrete sources by the ultimate {\\it Fermi} sensitivity in the near future (IT09), while almost all of the star-forming galaxy component will remain unresolved because of the faintness of individual sources. Therefore, any exotic contribution like dark matter annihilation cannot be probed directly under the level of the star-forming galaxies, i.e., $\\sim$ 10\\% of the total EGRB flux. Another approach such as utilizing anisotropy would be required to search for the signal under that level (see, e.g. Ando et al. 2007; \\citealt{2009PhRvL.102x1301S}, \\citealt{2009arXiv0912.1854H}, \\citealt{2010arXiv1005.0843C}) We assumed a simple empirical relation between gamma-ray luminosity, SFR, and gas mass of galaxies. Only two spectral templates were used in the calculation. A next step would be to construct a more physical model of gamma-ray luminosity and spectrum based on a larger number of physical quantities (e.g., size and stellar radiation field in addition to SFR and gas mass), taking into account propagation of cosmic-rays and production processes of gamma-rays." }, "1005/1005.3489_arXiv.txt": { "abstract": "Iron, the Universe's most abundant refractory element, is highly depleted in both circumstellar and interstellar environments, meaning it exists in solid form. The nature of this solid is unknown. In this Letter, we provide evidence that metallic iron grains are present around oxygen-rich AGB stars, where it is observationally manifest as a featureless mid-infrared excess. This identification is made using \\emph{Spitzer Space Telescope} observations of evolved globular cluster stars, where iron dust production appears ubiquitous and in some cases can be modelled as the only observed dust product. In this context, FeO is examined as the likely carrier for the 20-$\\mu$m feature observed in some of these stars. Metallic iron appears to be an important part of the dust condensation sequence at low metallicity, and subsequently plays an influential r\\^{o}le in the interstellar medium. We explore the stellar metallicities and luminosities at which iron formation is observed, and how the presence of iron affects the outflow and its chemistry. The conditions under which iron can provide sufficient opacity to drive a wind remain unclear. ", "introduction": "Iron is the sixth most abundant element in the Universe, and the most abundant refractory element. It is observed to be highly depleted in both interstellar and circumstellar environments \\citep{SB79,MvWLE05,DIRMV09}, and must therefore predominantly exist in an unknown solid form. Iron may be incorporated in other dust grains, primarily silicates, though these are usually iron-poor \\citep{GS99,KdKW+02}. Alternatively, iron may form a metallic condensate \\citep{KdKW+02,VvdZH+09}. Dust grains form around evolved stars by condensation from the gas phase, either directly, or onto molecular `seeds'. Individual dust species can usually be identified by distinct infrared emission bands. In oxygen-rich environments, amorphous or crystalline forms of silicates, spinels and corundum are observed to form, whilst carbon-rich environments give rise to amorphous carbon, graphite and SiC. Metallic iron grains, however, produce featureless infrared emission which can be difficult to differentiate from other sources, particularly amorphous carbon dust. Iron has hitherto only been inferred as a likely component of oxygen-rich dusty winds \\citep{KdKW+02,VvdZH+09}, but never positively identified. Previous observational studies \\citep{MvLD+09,BMvL+09} of giant stars in globular clusters have found a featureless contribution in addition to the flux emanating from the star's photosphere, which has been attributed to an unidentified circumstellar dust species. We herein show that this dust species is metallic iron. ", "conclusions": "Using mid-infrared spectra and photometry, we have shown that there is a considerable amount of unattributed infrared flux in a large selection of globular cluster giants. Through radiative transfer modelling, we deduce that this flux is most likely due to metallic iron grains forming in a truncated stellar wind, corroborated by a potential identification of FeO with the 20-$\\mu$m emission feature. The production of metallic iron seems to become more dominant at lower metallicity, suggesting that the dust condensation sequence is fundamentally different in metal-poor objects. Large-scale production of iron grains and iron oxide in AGB stars can explain iron depletion in the gas and solid phases of the post-AGB stars and planetary nebulae; as well as isotopic ratios in FeO grains in meteorites. While iron increases opacity in oxygen-rich winds, it remains unclear whether it can dominate driving of metal-poor winds. Future observations to determine outflow velocities in these stars should help determine the driver of these stellar winds. \\vspace{5 mm} This paper uses observations made using the \\emph{Spitzer Space Telescope}, operated by JPL, California Institute of Technology under NASA contract 1407 and supported by NASA through JPL (contract number 1257184). We thank Martha Boyer for her helpful comments." }, "1005/1005.4867_arXiv.txt": { "abstract": "{} {Aims: To provide a significantly improved probability distribution for the H-test for periodicity in X-ray and $\\gamma$-ray arrival times, which is already extensively used by the $\\gamma$-ray pulsar community. Also, to obtain an analytical probability distribution for stacked test statistics in the case of a search for pulsed emission from an ensemble of pulsars where the significance per pulsar is relatively low, making individual detections insignificant on their own. This information is timely given the recent rapid discovery of new pulsars with the Fermi-LAT t $\\gamma$-ray telescope. } {Methods: Approximately $10^{14}$ realisations of the H-statistic ($H$) for random (white) noise is calculated from a random number generator for which the repitition cycle is $\\gg 10^{14}$. From these numbers the probability distribution $P(>H)$ is calculated. } {Results: The distribution of $H$ is is found to be exponential with parameter $\\lambda=0.4$ so that the cumulative probability distribution $P(>H)=\\exp{(-\\lambda H)}$. If we stack independent values for $H$, the sum of $K$ such values would follow the Erlang-K distribution with parameter $\\lambda$ for which the cumulative probability distribution is also a simple analytical expression. }", "introduction": "When searching for a periodic signal in X-ray or $\\gamma$-ray arrival times dominated by noise, we may either perform a blind search for $\\gamma$-ray pulsars as demonstrated by the Fermi-LAT Collaboration (Abdo et al. \\cite{fermi1}), or, search for such a signal where the frequency parameters have been prescribed by contemporary radio data (Weltevrede \\cite{fermi2}). Following the folding of say $N$ arrival times $t_1, ...,t_N$ modulo the pulsar spin parameters, we arrive at a set of phases $\\theta_i$, $i=1, ...N$. However, in the case of blind searches, Atwood et al. \\cite{atwood} introduced a time differencing technique in which case the number of trial periods is significantly reduced. de Jager, Swanepoel \\& Raubenheimer (\\cite{dsr}, hereafter DSR) reviewed the general class of {\\it Beran} statistics (Beran \\cite{beran}), from which the most general test statistics such as Pearson's $\\chi^2$, Rayleigh and $Z^2_m$ statistics are derived, and from within this class they derived the well known {\\it H-test} for X-ray and $\\gamma$-ray Astronomy. The probability distribution of the {\\it H-test} statistic as given by DSR was derived from Monte Carlo simulations employing $\\sim 10^8$ simulations. The computational power and random number simulators on typical IBM machines during the 1980's had limited ranges of applicability and the {\\it H-test} suffered accordingly. For values of $H<23$ we found that the probability distribution was exponential with parameter $\\lambda=0.398$ (or 0.4), whereas a hard tail developed for $H>23$, which resulted in a significant compromise in sensitivity. The old version of the {\\it H-test} probability distribution is already extensively used by e.g. the Fermi-LAT Collaboration for pulsar searches (e.g. Abdo et al. \\cite{fermi1} and Weltevrede et al. \\cite{fermi2}), and from this paper it will become clear that the significances assigned to pulsar detections (or non-detections) may be too conservative, so that some pulsars may be missed, especially when many trial periods are involved, so that large values of the $H$-statistic are required for a significant detection . In this Letter we notify the community that all previous published significances from the {\\it H-test} should be reassessed, based on the new improved distribution presented below. Before we do so, we first briefly review the origin and properties of the {\\it H-test}. ", "conclusions": "For $M=1$ we retrieve the well-known {\\it Rayleigh test}, with the exception that the parameter for the exponential distribution has been reduced from $\\lambda=0.5$ (for the {\\it Rayleigh test}) to $\\lambda=0.4$ for the $H-test$. This slight loss of sensitivity is the effect of the number of trials taken implicitly into account as a result of the search through $m$ within the $H-test$. Finally, it is clear that the corrected distribution of the {\\it H}-statistic follows a simple exponential with parameter $\\lambda=0.4$ and evaluation of results for $H>23$ (i.e. the $10^{-4}$ significance level) would yield more significant results compared to the old distribution presented by DSR. For example, for $H=50$ a probability of $4\\times 10^{-8}$ is typically quoted in the literature, whereas the true probability for uniformity is actually $2\\times 10^{-9}$ - already a factor 20 smaller. A Fermi-LAT example of the Vela pulsar ($>500$ MeV) shows clearly values for $M$ up to 20 for ``skymap'' significances $X=p\\sqrt{N} > 20$, whereas $H$ scales with $H\\propto X^2$ as expected for Beran-type tests. The scaling $H=1.9X^2$ can be used to predict $H$-test statistics for Vela-like pulsars above 500 MeV if we assume the excess on the skymap is all pulsed. The hard tail of the distribution beyond $H>23$ presented by DSR probably arose from the repitition cycle of the random number generator used in those days, so that the same fluctuations at large $H$ values were repeated given the finite cycle length of the generator used. In this case we however used a generator with a cycle time much longer than $10^{14}$, in which case we did not see the repitition of outliers as a result of a finite cycle length. Confirmation of the possible break (i.e. downturn) in the probability distribution at $H>70$ requires extensive simulations beyond $4\\times 10^{14}$ and is beyond the scope of this paper. \\begin{figure} \\centering \\includegraphics[width=8 cm]{h-statistic_2.EPS} \\caption{The distribution of the {\\it H-statistic} derived from $4\\times 10^{14}$ Monte Carlo simulations with the best fit model for the cumulative probability $P(>H)=\\exp{(-0.4H)}$ (solid line) and the old version of the probability distribution given by DSR shown as a dashed line.} \\label{fig2} \\end{figure}" }, "1005/1005.1445_arXiv.txt": { "abstract": "\\vspace{1cm} \\centerline{\\bf ABSTRACT}\\vspace{2mm} Very recently, Verlinde considered a theory in which space is emergent through a holographic scenario, and proposed that gravity can be explained as an entropic force caused by changes in the information associated with the positions of material bodies. Then, motivated by the Debye model in thermodynamics which is very successful in very low temperatures, Gao modified the entropic force scenario. The modified entropic force (MEF) model is in fact a modified gravity model, and the universe can be accelerated without dark energy. In the present work, we consider the cosmological constraints on the MEF model, and successfully constrain the model parameters to a narrow range. We also discuss many other issues of the MEF model. In particular, we clearly reveal the implicit root to accelerate the universe in the MEF model. ", "introduction": "\\label{sec1} Very recently, Verlinde~\\cite{r1} considered a theory in which space is emergent through a holographic scenario, and proposed that gravity can be explained as an entropic force caused by changes in the information associated with the positions of material bodies. In this scenario, Verlinde has successfully derived the Newton's law of gravitation, the Einstein equations, and the law of inertia, from the entropic point of view. In fact, the entropic force scenario is similar to the old idea of Jacobson~\\cite{r2}, but also beyond it in some sense. Similar entropic insight into gravity has also been made by Padmanabhan~\\cite{r3} independently and simultaneously. Here we briefly mention some key points of the entropic force scenario following~\\cite{r1}. Motivated by Bekenstein's argument~\\cite{r4}, Verlinde postulated that the change in entropy near the holographic screen is linear in the displacement $\\Delta x$, namely, \\be{eq1} \\Delta S=2\\pi k_B\\frac{mc}{\\hbar}\\Delta x\\,, \\ee where $m$ is the mass of test particle, whereas $k_B$, $c$ and $\\hbar$ are Boltzmann constant, speed of light and the reduced Planck constant, respectively. The effective entropic force acting on the test particle due to the change in entropy obeys the first law of thermodynamics \\be{eq2} F\\Delta x=T\\Delta S\\,, \\ee where $T$ is the temperature. If one takes the Unruh temperature $T$ experienced by an observer in an accelerated frame whose acceleration is $a$, i.e., \\be{eq3} k_B T=\\frac{1}{2\\pi}\\frac{\\hbar a}{c}\\,, \\ee to be the temperature associated with the bits on the holographic screen, from Eqs.~(\\ref{eq1})---(\\ref{eq3}), it is easy to recover the second law of Newton \\be{eq4} F=ma\\,. \\ee Considering a sphere as the holographic screen, Verlinde assumed that the number of used bits on the holographic screen $N$ is proportional to the area $A=4\\pi r^2$, i.e., \\be{eq5} N=\\frac{Ac^3}{G\\hbar}\\,. \\ee According to the equipartition law of energy, the total energy inside the screen is \\be{eq6} E=\\frac{1}{2}N k_B T\\,. \\ee Of course, one can identifies $E$ with the mass $M$ inside the screen through \\be{eq7} E=Mc^2. \\ee From Eqs.~(\\ref{eq1}), (\\ref{eq2}), and (\\ref{eq5})---(\\ref{eq7}), one can recover the Newton's law of gravitation \\be{eq8} F=G\\frac{Mm}{r^2}\\,, \\ee where $G$ can be identified with the Newton constant now. From Eqs.~(\\ref{eq3}), (\\ref{eq4}) and (\\ref{eq8}), it is easy to find the gravitational acceleration \\be{eq9} g=\\frac{GM}{r^2}\\,, \\ee and the temperature \\be{eq10} T=\\frac{\\hbar}{k_B c}\\frac{g}{2\\pi}\\,. \\ee As shown in~\\cite{r1}, a relativistic generalization of the presented arguments directly leads to the Einstein equations. We strongly refer to the original paper~\\cite{r1} for great details. Soon after Verlinde's proposal of entropic force, many relevant works appeared. For examples, Cai, Cao and Ohta~\\cite{r5}, Shu and Gong~\\cite{r6} derived the Friedmann equations from entropic force simultaneously. Smolin~\\cite{r7} derived the Newtonian gravity in loop quantum gravity. Li and Wang~\\cite{r8} showed that the holographic dark energy can arise in the entropic force scenario. Easson, Frampton and Smoot~\\cite{r9} considered the entropic accelerating universe and the entropic inflation. Tian and Wu~\\cite{r10}, Myung~\\cite{r11} discussed the thermodynamics of black holes in the entropic force scenario. Vancea and Santos~\\cite{r12} considered the uncertainty principle from the point of view of entropic force. Zhang, Gong and Zhu~\\cite{r13}, Sheykhi~\\cite{r14} derived the modified Friedmann equation from the corrected entropy. Also, Modesto and Randono~\\cite{r15} discussed the corrections to Newton's law from the corrected entropy. Cai, Liu and Li~\\cite{r16} considered a unified model of inflation and late-time acceleration in the entropic force scenario. For other relevant works to entropic force, we refer to e.g.~\\cite{r17,r18,r19,r36} and references therein. The works mentioned above are in fact closely following Verlinde's proposal of entropic force~\\cite{r1}. To be honest, here we should also mention the other works which are strongly criticizing the entropic force scenario. For instance, the author of~\\cite{r38} argued that there are some possible flaws in Verlinde's idea. In~\\cite{r39}, Culetu argued that the relativistic Unruh temperature cannot be associated with the bits on the screen in the form considered by Verlinde. In~\\cite{r40}, Hossenfelder argued that some additional assumptions made by Verlinde are unnecessary and there are some gaps in Verlinde's arguments. In~\\cite{r41}, Myung found that entropic force does not always imply the Newtonian force law, and the connection between Newtonian cosmology and entropic force cannot be confirmed. In~\\cite{r42}, Li and Pang found that inflation is inconsistent with the entropic force scenario. In~\\cite{r43}, Lee argued that there are some inconsistencies in Verlinde's arguments from a classical point of view. So far, we have briefly surveyed the current status of the works relevant to the entropic force scenario. It is fair to say that the entropic force scenario is still in controversy. On the other hand, there is no breakthrough on entropic force after Verlinde's proposal~\\cite{r1}. A deep insight is needed to understand the nature of gravity. In addition, further discussions on the entropic force scenario are also desirable. Only when more and more results on entropic force are available, one can say something conclusively at that time. To this end, we would like to contribute our effort and try to learn more about the entropic force scenario. In this work, we will consider a modified entropic force scenario proposed by Gao~\\cite{r20}, which has some interesting features. And then, we will constrain the modified entropic force scenario with the latest observational data. In~\\cite{r20}, Gao noted that statistical thermodynamics reveals the equipartition law of energy does not hold in the very low temperatures. Instead, as is well known, the Debye model~\\cite{r21,r22} is very successful in explaining the experimental results when the temperatures are very low. Since the equipartition law of energy plays an important role in the derivation of entropic force, the entropic force should be modified for the very weak gravitational fields which correspond to very low temperatures. Especially, the large-scale universe is in such an extreme weak gravitational field, and hence the modified entropic force (MEF) makes sense in cosmology. Following~\\cite{r20}, we briefly mention the key points of MEF model. Similar to the Debye model~\\cite{r21,r22} in thermodynamics, one can modify the equipartition law of energy in Eq.~(\\ref{eq6}) to \\be{eq11} E=\\frac{1}{2}N k_B T D(x)\\,, \\ee where $D(x)$ is the Debye function which is defined by \\be{eq12} D(x)=\\frac{3}{x^3}\\int_0^x\\frac{y^3}{e^y-1}\\,dy\\,, \\ee and $x$ is related to the temperature $T$ as \\be{eq13} x\\equiv\\frac{T_D}{T}\\,, \\ee in which $T_D$ is the Debye temperature. By definition, $x$ is positive. With the modified equipartition law of energy, namely Eq.~(\\ref{eq11}), similar to the original entropic force, one can easily to obtain~\\cite{r20} \\be{eq14} g=\\frac{GM}{r^2}\\frac{1}{D(x)}\\,, \\ee in which (nb. Eq.~(\\ref{eq10})) \\be{eq15} x=\\frac{T_D}{T}=\\frac{g_D}{g}\\,, \\ee where $g_D\\equiv(2\\pi\\,k_B\\,c/\\hbar)\\,T_D$ is the Debye acceleration. Actually, Eq.~(\\ref{eq14}) corresponds to the modified Newtonian law of gravity. In the limit of strong gravitational field, $g\\gg g_D$ and hence $x\\ll 1$, from Eq.~(\\ref{eq12}) it is easy to find that $D(x)\\to 1$ and the Newtonian gravity is recovered. On the other hand, in the limit of weak gravitational field, $g\\ll g_D$ and hence $x\\gg 1$, one can see that $D(x)\\to\\pi^4/(5x^3)$ and then $g\\propto 1/\\sqrt{r}$, which significantly deviates from the familiar inverse square law~\\cite{r20}. However, as argued in~\\cite{r20}, one need not to worry about the possibility of MEF against the experimental results of the inverse square law. Since these experiments testing the inverse square law were done on the Earth or in the solar system, which are actually in the strong gravitational fields, we have $x\\ll 1$ and $D(x)\\to 1$, therefore the deviation from the inverse square law are extremely tiny. The significant deviation from the inverse square law can only occur in the very large scale in the universe where the gravitational fields are very weak, and hence it can escape the detection of these experiments testing the inverse square law. Of course, this argument relies on a small $g_D$. We will justify it later in the present work. Next, we turn to the cosmological issues in the MEF Model. For convenience, we set the units $k_B=c=\\hbar=1$ hereafter. Using the derivation method in~\\cite{r5,r6}, one can find that the modified Raychaudhuri equation is given by~\\cite{r20} \\be{eq16} 4\\pi G\\left(\\rho+p\\right)=-\\left(\\dot{H}-\\frac{K}{a^2}\\right) \\left[-2D(x)+\\frac{3x}{e^x-1}\\right]\\,, \\ee where $\\rho$ and $p$ are the total energy density and total pressure of cosmic fluids, respectively; $K$ is the spatial curvature of the universe; $H\\equiv\\dot{a}/a$ is the Hubble parameter; $a=(1+z)^{-1}$ is the scale factor (we have set $a_0=1$); $z$ is the redshift; a dot denotes the derivatives with respect to cosmic time $t$; the subscript ``0'' indicates the present value of the corresponding quantity. Taking into account the Hawking temperature $T$ for the universe~\\cite{r23} \\be{eq17} T=\\frac{H}{\\,2\\pi}\\,, \\ee from Eqs.~(\\ref{eq13}) and (\\ref{eq15}), it is easy to see that \\be{eq18} x=\\frac{H_D}{H}\\,, \\ee where $H_D=g_D$. On the other hand, the energy conservation equation still holds in the MEF model, namely \\be{eq19} \\dot{\\rho}+3H\\left(\\rho+p\\right)=0\\,. \\ee From Eqs.~(\\ref{eq16}) and (\\ref{eq19}), one can derive the corresponding Friedmann equation. It is anticipated that Friedmann equation is also modified, $H^2\\not=(8\\pi G\\rho)/3$, due to the correction term $-2D(x)+3x/(e^x-1)$ in Eq.~(\\ref{eq16}). The MEF model is in fact a modified gravity model. Gao showed that the MEF model can describe the accelerating universe without dark energy. We refer to the original paper~\\cite{r20} for details. Since Gao has not considered the constraints on the MEF model in~\\cite{r20}, we will try to obtain the cosmological constraints with the latest observational data in the next section. Further, we will discuss some relevant issues of the MEF model in Sec.~\\ref{sec3}. Finally, we give the brief conclusion and some meaningful remarks in Sec.~\\ref{sec4}. \\begin{center} \\begin{figure}[tbhp] \\centering \\includegraphics[width=0.5\\textwidth]{sn.eps} \\caption{\\label{fig1} The $68.3\\%$ and $95.4\\%$ confidence level contours in the $\\Omega_{m0}-\\zeta$ parameter space. The best-fit parameters are also indicated by a black solid point. This result is obtained by using the data of 557 Union2 SNIa alone.} \\end{figure} \\end{center} \\vspace{-10mm} % ", "conclusions": "\\label{sec4} In summary, we considered the cosmological constraints on the MEF model in this work, by using the observational data of 557 Union2 SNIa, the distance parameter $A$ from LSS, the shift parameter $R$ from CMB, and 59 Hymnium GRBs. We found that the key parameter $\\zeta$ in MEF model has been limited in a narrow range $0\\leq\\zeta\\,\\lsim\\, 0.2$ within $95.4\\%$ confidence level. By using the important result given in Eq.~(\\ref{eq28}), we have clearly shown the key point to understand the reason for accelerating the universe without dark energy in the MEF model. We showed the MEF model reduces to $\\Lambda$CDM model when $\\zeta\\ll 1$. However, the best-fit $\\zeta$ for the observations SNIa+LSS+CMB and SNIa+LSS+CMB+GRBs significantly deviates from zero. This indicates the new feature of MEF model different from $\\Lambda$CDM model. We have justified the approximate solution $E(z)$ given in Eq.~(\\ref{eq28}). We plotted $E(z)$ and $q(z)$ as functions of redshift $z$, and clearly showed that the universe can be accelerated in late time without dark energy. Finally, we have shown that MEF can avoid the conflict with the experiments testing the inverse square law. After all, some remarks are in order. In the MEF model, as shown in this work, the universe can be accelerated without dark energy. The only component is dust matter. The MEF model is in fact a modified gravity model, similar to the $f(R)$ models and the braneworld models. The MEF model can be degenerate to $\\Lambda$CDM model, but it has not an explicit cosmological constant in the model. This is a big advantage in fact, beyond some $f(R)$ and braneworld models in this sense. Secondly, we point out the possibility to extend the original MEF model. In principle, it is not necessary to restrict the energy component in the universe to be dust matter only. The universe can contain other components, such as, dark energy. For example, we can consider a universe containing both dust matter and dark energy whose equation-of-state parameter (EoS) $w_X$ is a constant. In this case, the total energy density $\\rho=\\rho_m+\\rho_X=\\rho_{m0}\\,a^{-3}+\\rho_{X0}\\,a^{-3(1+w_X)}$. Substituting into Eq.~(\\ref{eq20}), we have \\be{eq37} \\left[-2D\\left(\\frac{\\zeta}{E}\\right)+\\frac{3\\,\\zeta/E} {e^{\\,\\zeta/E}-1}\\right]\\cdot 2E\\,\\frac{dE}{dz}= 3\\Omega_{m0}\\,(1+z)^2+ 3\\left(1+w_X\\right)\\Omega_{X0}(1+z)^{3w_X+2}\\,, \\ee where $\\Omega_{X0}\\equiv(8\\pi G\\rho_{X0})/(3H_0^2)$. Note that $\\Omega_{m0}+\\Omega_{X0}\\not=1$, since Friedmann equation has been modified in the MEF model. In principle, one can numerically solve Eq.~(\\ref{eq37}) to get $E(z)$. Similar to the original MEF model, we find that an approximation of the exact differential equation~(\\ref{eq37}) is given by \\be{eq38} \\left(1-\\frac{3}{4}\\frac{\\zeta}{E}\\right)\\cdot 2E\\,dE= \\Omega_{m0}\\,da^{-3}+\\Omega_{X0}\\,da^{-3(1+w_X)}\\,. \\ee Integrating Eq.~(\\ref{eq38}), we have \\be{eq39} E^2-\\frac{3}{2}\\,\\zeta E=\\Omega_{m0}\\,a^{-3} +\\Omega_{X0}\\,a^{-3(1+w_X)}+const., \\ee where $const.$ is the integral constant, which can be determined by requiring $E(z=0)=1$. Finally, we find that \\be{eq40} E^2-\\frac{3}{2}\\,\\zeta E=\\Omega_{m0}\\,(1+z)^3+ \\Omega_{X0}\\,(1+z)^{3(1+w_X)} +\\left(1-\\frac{3}{2}\\,\\zeta-\\Omega_{m0}-\\Omega_{X0}\\right), \\ee which is a quadratic equation of $E$ in fact. Noting that $E$ is positive, we solve Eq.~(\\ref{eq40}) to get \\be{eq41} E(z)=\\frac{3}{4}\\,\\zeta+\\frac{1}{2}\\left\\{\\frac{9}{4}\\,\\zeta^2 +4\\left[\\Omega_{m0}\\,(1+z)^3+\\Omega_{X0}\\,(1+z)^{3(1+w_X)} +\\left(1-\\frac{3}{2}\\,\\zeta-\\Omega_{m0}-\\Omega_{X0}\\right) \\right]\\right\\}^{1/2}. \\ee In fact, this is just a simple example. One can include any type of dark energy, for instance, the CPL dark energy whose EoS is given by $w_{de}=w_0+w_a(1-a)$, quintessence, phantom, k-essence, hessence, (generalized) Chaplygin gas, holographic/agegraphic dark energy, vector-like dark energy, spinor dark energy, and so on. Therefore, we would like to give the more general formulae. In this case, the total energy density $\\rho=\\rho_m+\\rho_{de}=\\rho_{m0}\\,a^{-3}+\\rho_{de,0}\\,f(a)$, where $f(a)$ can be any function of $a$ which satisfies $f(a=1)=1$. Substituting into Eq.~(\\ref{eq20}), we obtain \\be{eq42} \\left[-2D\\left(\\frac{\\zeta}{E}\\right)+\\frac{3\\,\\zeta/E} {e^{\\,\\zeta/E}-1}\\right]\\cdot 2E\\,\\frac{dE}{dz}= 3\\Omega_{m0}\\,(1+z)^2- \\Omega_{de,0}\\,(1+z)^{-2}f^\\prime\\,, \\ee where $f^\\prime\\equiv df/da$, and $\\Omega_{de,0}\\equiv(8\\pi G\\rho_{de,0})/(3H_0^2)$. Note again that $\\Omega_{m0}+\\Omega_{de,0}\\not=1$, since Friedmann equation has been modified in the MEF model. Eq.~(\\ref{eq42}) is the exact differential equation, which can be used to find the exact $E(z)$ numerically. Also, we give the corresponding approximate solution as \\be{eq43} E(z)=\\frac{3}{4}\\,\\zeta+\\frac{1}{2}\\left\\{\\frac{9}{4}\\,\\zeta^2 +4\\left[\\Omega_{m0}\\,(1+z)^3+ \\Omega_{de,0}\\,f\\left(\\frac{1}{1+z}\\right) +\\left(1-\\frac{3}{2}\\,\\zeta-\\Omega_{m0}-\\Omega_{de,0}\\right) \\right]\\right\\}^{1/2}. \\ee Similarly, if one need to add other component, such as radiation, it is not a hard work. In fact, we can give the most general formulae. In this case, the total energy density $\\rho=\\rho_0\\,f(a)$, where $f(a)$ can be any function of $a$ which satisfies $f(a=1)=1$. Substituting into Eq.~(\\ref{eq20}), we have \\be{eq44} \\left[-2D\\left(\\frac{\\zeta}{E}\\right)+\\frac{3\\,\\zeta/E} {e^{\\,\\zeta/E}-1}\\right]\\cdot 2E\\,\\frac{dE}{dz}= -\\Omega_0\\,(1+z)^{-2}f^\\prime\\,, \\ee where $\\Omega_0\\equiv(8\\pi G\\rho_0)/(3H_0^2)$. Note again that $\\Omega_0\\not=1$, since Friedmann equation has been modified in the MEF model, namely $H^2\\not=(8\\pi G\\rho)/3$. Eq.~(\\ref{eq44}) is the exact differential equation, which can be used to find the exact $E(z)$ numerically. Also, we give the corresponding approximate solution as \\be{eq45} E(z)=\\frac{3}{4}\\,\\zeta+\\frac{1}{2}\\left\\{\\frac{9}{4}\\,\\zeta^2 +4\\left[\\,\\Omega_0\\,f\\left(\\frac{1}{1+z}\\right) +\\left(1-\\frac{3}{2}\\,\\zeta-\\Omega_0\\right)\\right]\\right\\}^{1/2}. \\ee In fact, noting that $E=H/H_0$, Eq.~(\\ref{eq45}) can be regarded as the approximate modified Friedmann equation in the MEF model. If $\\zeta\\ll 1$, Eq.~(\\ref{eq45}) reduces to \\be{eq46} H^2=\\frac{8\\pi G}{3}\\,\\rho+\\Lambda_{\\rm eff}\\,, \\ee where $\\Lambda_{\\rm eff}=\\left(1-\\Omega_0\\right)H_0^2=const.$ is actually an effective cosmological constant. Therefore, in the most general case, we can clearly reveal the implicit root to accelerate the universe in the MEF model, regardless of the energy components in the universe. An effective cosmological constant is the intrinsic feature (we refer to e.g.~\\cite{r23} for a previous insight). The other exotic features of the MEF model could emerge only when $\\zeta$ significantly deviates from zero. Thirdly, we would like to say some words on the understanding of the original entropic force model~\\cite{r1} and the modified entropic force model~\\cite{r20}. In fact, entropic force is just a new perspective to gravity, from the thermodynamical point of view. So, the original entropic force can only recover all results of the usual (Newton and Einstein) gravity. The only new thing is the reversed logic which might reveal the nature of gravity. In the original entropic force model~\\cite{r1}, using the fundamental assumptions Eqs.~(\\ref{eq1}), (\\ref{eq3}), (\\ref{eq5}) and~(\\ref{eq6}), Verlinde derived the Newton's law of gravitation Eq.~(\\ref{eq8}) for the (non-relativistic) Euclidean spacetime in section~3 of~\\cite{r1}, and also derived the Einstein gravitational equations for any (relativistic) curved spacetime in section~5 of~\\cite{r1}. On the other hand, the Friedmann equations were derived in~\\cite{r5,r6} for the Friedmann-Robertson-Walker (FRW) universe. There is {\\em no} any mixing here. We should mention that both the original entropic force~\\cite{r1} and the modified entropic force~\\cite{r20} cannot be understood in {\\em only} non-relativistic or relativistic cases. In fact, they are equivalent to gravity itself in all cases. As the usual understanding, the Newton's law of gravitation is just the approximation of Einstein gravitational equations in the (non-relativistic) small scale limit, whereas the Friedmann equations are just the special case of Einstein gravitational equations in the cosmic scale (homogeneous and isotropic spacetime). The situation is similar in the modified entropic force model. Using the fundamental assumptions Eqs.~(\\ref{eq1}), (\\ref{eq3}), (\\ref{eq5}) and~(\\ref{eq11}), in~\\cite{r20} Gao derived the Newton's law of gravitation Eq.~(\\ref{eq14}) for the (non-relativistic) Euclidean spacetime, and also derived the second Friedmann equation~(\\ref{eq16}) for the FRW universe. Note that the first Friedmann equation can be derived from the second Friedmann equation~(\\ref{eq16}) and the energy conservation equation~(\\ref{eq19}). On the other hand, following Verlinde's derivations in section~5 of~\\cite{r1}, one can derive the corresponding (modified) Einstein gravitational equations for any (relativistic) curved spacetime. In fact, this is just the lacked sector in the modified entropic force model. However, it is available in principle, although it has not been given in the literature. In the modified entropic force model, there is {\\em no} any mixing too. The modified Newton's law of gravitation Eq.~(\\ref{eq14}) is just the approximation of the (lacked but available in principle) modified Einstein gravitational equations in the (non-relativistic) small scale limit, whereas the modified second Friedmann equation~(\\ref{eq16}) is just the special case of the (lacked but available in principle) modified Einstein gravitational equations in the cosmic scale (homogeneous and isotropic spacetime). Fourthly, we said that the MEF model is similar to $f(R)$-gravity or braneworld scenario. Notice that they are similar {\\em only} in the sense that the gravity has been modified in these models. Of course, both $f(R)$-gravity and braneworld scenario were derived from the known actions, whereas the action for MEF is still lacked in the literature. However, as mentioned above, following Verlinde's derivations in section~5 of~\\cite{r1}, in principle one can derive the corresponding (modified) Einstein gravitational equations for any (relativistic) curved spacetime. Once this lacked sector has been done, the explicit action is ready. Since the present work focuses on cosmology in the MEF model, we leave this task to future works. Fifthly, as mentioned in this work, in the MEF model, there is {\\em no} dark energy in fact. The universe is matter-dominated always. The expansion of our universe is accelerated due to the fact that gravity has been modified. In the non-relativistic case, there is no dark energy too, but gravity is also modified. However, as mentioned in this work, this modification to Newtonian gravity is negligible on the Earth or in the solar system. In the larger scale, the modified gravity is described by Eq.~(\\ref{eq14}). As shown in~\\cite{r37}, the Debye entropic force can be an alternative to the modified Newtonian dynamics (MOND) to explain the rotational velocity curves of spiral galaxies. In fact, the MEF model~\\cite{r20} and the Debye entropic force model~\\cite{r37} are very similar. So, it is anticipated that the ``non-relativistic cosmology'' of the MEF model could be an alternative to dark matter, which is usually invoked to explain the rotational velocity curves of spiral galaxies. Finally, we admit that the entropic force proposed by Verlinde is based on several unproved hypotheses, and it is still controversial in the physical community. On the other hand, the Debye model in the thermodynamics has not been used in the gravity theory previously. However, in the history, many great theories also appeared controversially in their beginning. Therefore, we consider that it is better to keep an open mind to these speculative attempts." }, "1005/1005.4326_arXiv.txt": { "abstract": "We study the energy released from phase-transition induced collapse of neutron stars, which results in large amplitude stellar oscillations. To model this process we use a Newtonian hydrodynamic code, with a high resolution shock-capturing scheme. The physical process considered is a sudden phase transition from normal nuclear matter to a mixed phase of quark and nuclear matter. We show that both the temperature and the density at the neutrinosphere oscillate with time. However, they are nearly 180$^{\\circ}$ out of phase. Consequently, extremely intense, pulsating neutrino/antineutrino and leptonic pair fluxes will be emitted. During this stage several mass ejecta can be ejected from the stellar surface by the neutrinos and antineutrinos. These ejecta can be further accelerated to relativistic speeds by the electron/positron pairs, created by the neutrino and antineutrino annihilation outside the stellar surface. We suggest that this process may be a possible mechanism for short Gamma-Ray Bursts. ", "introduction": "Recently, by using simulations performed with the Newtonian numerical code introduced in \\cite{Lin06}, it was shown that the resulting quark star, produced by the phase transition induced collapse of a neutron star, will undergo a series of oscillations. The collapse process with a conformally flat approximation to general relativity was also simulated in \\cite{Ab08}. The works of \\cite{Lin06,Ab08} focus on the gravitational wave signals emitted by the collapse process. It is the purpose of the present paper to consider another important implication of this result, namely, the effect of the oscillations of the newly formed quark star on the neutrino emission. The oscillations can enhance the neutrino emission rate in a pulsating manner, and the neutrinos are emitted in a much shorter time scale. Moreover, through the process of neutrino-antineutrino annihilation, a large amount of electron-positron pairs is also produced. Therefore the neutron-quark phase transition in compact objects may be the energy source of GRBs \\cite{Ch09}. Such a model can also explain the lack of detection of a neutron star or pulsar formed in the SN 1987A \\cite{Cha09}, by assuming that the newly formed neutron star at the center of SN 1987A underwent a phase transition after the neutrino trapping timescale ($\\sim$ 10 s). Consequently, the compact remnant of SN 1987A may be a strange quark star, which has a softer equation of state than that of neutron star matter. Such a phase transition can induce stellar collapse and result in large amplitude stellar oscillations. Extremely intense pulsating neutrino fluxes, with submillisecond period and with neutrino energy (greater than 30 MeV), can be emitted because the oscillations of the temperature and density are out of phase almost 180 $^{\\circ}$. If this is true, the current X-ray emission from the compact remnant of SN 1987A will be lower than $10^{34}$ erg s$^{-1}$, and it should be a thermal bremsstrahlung spectrum for a bare strange star with a surface temperature of around $\\sim $ 107 K. ", "conclusions": "In this paper we have studied the possible consequences of the phase-induced collapse of neutron stars to strange stars. We have found that both the density and the temperature inside the star will oscillate with the same period, but almost 180$^{\\circ}$ out of phase, which will result in the emission of intense pulsating neutrinos and pairs. We want to point out that the intense pulse neutrino luminosity can be maintained due to the oscillatory fluid motion, which can carry thermal energy directly from the stellar core to the surface. This process can replenish the energy loss of neutrino emission much quicker than the neutrino diffusion process. A large fraction of the neutrino energy, roughly (1-1/$e$), will be absorbed by the matter very near the stellar surface. When this amount of energy exceeds the gravitational binding energy, some mass near the stellar surface will be ejected, and this mass will be further accelerated by absorbing pairs created from the neutrino and antineutrino annihilation processes outside the star. Although matter will be ejected periodically, each ejecta can have different masses and Lorentz factors, and therefore the intrinsic period could not be observed. We suggest that the collisions among these ejecta may produce short GRBs. Our numerical simulations are simulating a spherically symmetric and non-rotating collapsing stellar object, and they also do not contain magnetic fields. Therefore the radiation emission produced in this model is isotropic. However, a realistic neutron star should have finite angular momentum and strong magnetic field, and hence these two factors could produce asymmetric mass ejection. This effect will be considered in future work. The phase-transition from a neutron star to a strange star was simulated in \\cite{FrWo98}, with the conclusion that this process is most likely not a gamma-ray burst mechanism. They mimic the phase-transition by the arbitrary motion of a piston deep within the star, and they have found that the mechanic wave will eject $\\sim 10^{-2}M_{\\odot}$ baryons, which causes the baryon contamination for the gamma-ray bursts. In our simulations, we assume a sudden change of equation of state to mimic the phase-transition, and we use the Newtonian hydrodynamic code to study the response of the stellar interior after such a sudden change of the EOS. In our simulations we find that the mass ejection by the motion of the fluid is very small. We estimate that the major mass ejection would result from the heating of neutrinos and pairs on the stellar crust, which is not modeled in the simulations. Our total energy output and total mass in ejecta are close to that of \\cite{FrWo98}. However, the neutrino energy injection is pulsating, and hence the mass ejection is also pulsating. The mass of individual ejecta range from $\\sim 10^{-9}$ to $\\sim 10^{-4}M_{\\odot}$, with output energy in the range of $10^{48}$ to $10^{50}$ ergs. Therefore, some ejecta cannot be relativistic, and they cannot contribute to GRBs. However, there are still many relativistic ejecta in each simulation model, which can have Lorentz factors $>$100, and with a total energy of $\\sim 10^{50}$ --- $10^{51}$ ergs. This could be a possible mechanism for short GRBs. \\bigskip KSC and TH are supported by the GRF Grants of the Government of the Hong Kong SAR under HKU7013/06P and HKU7025/07P respectively." }, "1005/1005.4332_arXiv.txt": { "abstract": "{ {The stable propagation of jets in FRII sources is remarkable if one takes into account that large-scale jets are subjected to potentially highly disruptive three-dimensional (3D) Kelvin-Helmholtz instabilities.} {Numerical simulations can address this problem and help clarify the causes of this remarkable stability. Following previous studies of the stability of relativistic flows in two dimensions (2D), it is our aim to test and extend the conclusions of such works to three dimensions.} {We present numerical simulations for the study of the stability properties of 3D, sheared, relativistic flows. This work uses a fully parallelized code (\\textit{Ratpenat}) that solves equations of relativistic hydrodynamics in 3D.} {The results of the present simulations confirm those in 2D. We conclude that the growth of resonant modes in sheared relativistic flows could be important in explaining the long-term collimation of extragalactic jets.} {} ", "introduction": "\\label{intro} Extragalactic jets from AGN present a morphological dichotomy between FRI and FRII type jets \\citep{fr74}. The former \\citep[e.g., 3C~31,][]{lb02} show a disrupted structure at kiloparsec scales, whereas the latter \\citep[e.g., Cyg A,][]{cb96} are highly collimated. Although the origin of this dichotomy is a complex combination of several intrinsic (e.g., jet power) and external (i.e., environmental) factors, the stable propagation of jets in FRII sources is remarkable considering that large-scale jets are subjected to potentially highly disruptive three-dimensional (3D) modes of the Kelvin-Helmholtz (KH) instability. Any perturbation may couple to an instability and grow in amplitude, becoming potentially disruptive \\citep{pe+05}, hence the importance of studying their development and growth from the linear to the nonlinear regime in terms of the flow parameters. Jet stability in the relativistic regime has been thoroughly studied in different scenarios and parameter ranges \\citep[see][for a recent review on the state-of-art]{ha06}, from both the analytical and numerical perspectives. The linear analysis of KH instability in relativistic flows started with the work of \\cite{ts76} and \\cite{bp76}, who derived and solved a dispersion relation for a single plane boundary between a relativistic flow and the ambient medium. Next, \\cite{ftz78} and \\cite{ha79} examined properties of the KH instability in relativistic cylindrical jets. The effects of a shear layer have been examined by \\cite{ur02}. More recently, \\cite{ha08} has studied the stability of magnetized relativistic jets with poloidal magnetic fields. An extension of the linear stability analysis in the relativistic case to the weakly nonlinear regime has been performed by \\cite{hz95,hz97} and led to the conclusion that KH instability saturates at finite amplitudes because of various nonlinear effects. The most significant effect results from the relativistic character of the jet flow, namely from the velocity perturbation not exceeding the speed of light in the reference frame of the jet. This result was confirmed by numerical simulations \\citep{pe+04a,pe+04b}. In the purely nonlinear regime, \\cite{ma97}, \\cite{ha98} and \\cite{ro99} demonstrate that jets with high Lorentz factors and high internal energy are influenced very weakly by the Kelvin-Helmholtz instability, contrary to the cases with lower Lorentz factors and lower internal energies. The former do not develop modes of KH instability predicted by the linear theory, which is interpreted as the result of a lack of appropriate perturbations (triggered by the backflow in this kind of simulations) generating the instability in the system, because in the limit of high internal energies of the jet matter the Kelvin-Helmholtz instability is expected to develop with the highest growth rate. A combination of linear and nonlinear studies of jet stability in the relativistic case was applied for the first time by \\cite{ha98} in the case of axisymmetric, cylindrical jets and then extended to the 3D case by \\cite{ha01} in the spatial approach and by \\cite{pe+04a,pe+04b} in the temporal approach, for 2D slab and cilyndrical jets, including the case of sheared flows in \\cite{pe+05}. Recently, \\cite{mi07} have extended this study to 3D relativistic magnetized jets, which may be stabilized (with respect to current-driven modes) by a sheath layer surrounding the jet core. Focusing on the combination of linear and nonlinear studies, \\cite{pe+04a,pe+04b} studied the growth of single symmetric modes in relativistic jets, sweeping different parameters, such as jet temperature, Lorentz factor, and density ratio, with the external medium. The linear work confirmed that KH modes grow faster in slower and hotter flows than in cold and fast ones. In Perucho, Mart\\'{\\i} \\& Hanasz (2005, Paper I from now on), this work was extended to sheared jets to study the growth of a mixture of symmetric and antisymmetric modes in this case. In \\cite{pe+04b} and Paper~I, the authors determined the stability properties of relativistic flows for a wide range of parameters in density contrast ($10^{-4}-0.1$), Lorentz factor ($5-20$) and specific internal energy ($0.08-76.5\\,\\rm{c^2}$). It was shown that the inertia of the jet, in terms of the density contrast with the external medium, the relativistic Mach number and the Lorentz factor are the parameters that determine the stability of jets in the nonlinear regime, as a relativistic counterpart to the work of \\cite{bm94} in the classical case. Moreover, in Paper I and \\cite{pe+07} (Paper II, from now on), it was shown that the growth of high-order Kelvin-Helmholtz modes developing in the shearing layer, hereafter referred to as resonant modes, could dramatically change the nonlinear evolution of the flow in the case of cold and fast jets. The results could be summarized as follows. a) Cold and slow jets (small relativistic Mach number and Lorentz factor) are unstable with the growth of the instability and are disrupted after the generation of a shock front crossing the boundary between the jet and the ambient medium, which results from the development of long wavelength KH instability modes (UST1 jets). b) Hot and slow jets (intermediate values of relativistic Mach numbers and Lorentz factor) are also unstable, but disrupted and mixed in a continuous way by the growth of the mixing layer down to the jet axis (UST2 jets). c) Faster (high values of Mach number and/or Lorentz factor) jets develop short-wavelength, high-order modes that grow in the shear layer and saturate the growth of the instability without loss of either collimation or mixing, and generate a hot shear layer around the core of the jet (ST jets). d) Lighter jets are more unstable than denser ones (UST1). In this work, we extend the previous studies to 3D and show the existence of a new mechanism to prevent jet decollimation based on the development of resonant KH perturbations. With the aforementioned experience in mind, and taking into account that 3D simulations are highly demanding in terms of computational resources, we present here three numerical simulations using parameters that are representative of the different stability regions found in Paper I, but now including helical and elliptic modes. It is our aim to test whether the same conclusions are valid and, thus, if resonant modes provide a general stabilising mechanism in the absence of strong magnetic fields. The paper is organized as follows. Section~\\ref{code} is devoted to presenting of a new 3D-RHD code for these simulations and the numerical setup used. Section~\\ref{res} contains the results of the linear analysis for the parameters considered and the simulations. In Section~\\ref{disc}, the results are discussed and the conclusions of this work are presented in Section~\\ref{conc}. ", "conclusions": "\\label{conc} In this work, we presented the first solutions of the linear problem for 3D sheared relativistic flows. Following this result, we presented numerical simulations of the growth of KH instability modes from the linear to the nonlinear regimes, for three selected relativistic jet models. These numerical simulations were performed with the new high-resolution shock-capturing 3D RHD code \\textit{Ratpenat}, which was parallelized using a hybrid scheme with OMP and MPI processes. The simulations of jets in the typical stability regions of the space of parameters (Paper~I) have confirmed previous results obtained in the 2D approximation. We proved that under certain conditions, the appearance of the shear-layer resonant modes of KH instability has important consequences for the stability of relativistic jets. In particular, the implications for the collimation of FRII jets were discussed. We also show, as a by-product of our simulations, that the development of different modes with a variety of wavelengths may induce rich transversal structure in jets, with shorter modes showing up in the central regions of the jet and longer ones dominating the structure in the mixing layer. In this respect, we confirm the results obtained in \\cite{pe+06}. Future work along this line will consider the extension of our study in both the linear and nonlinear regime by means of simulations with higher resolution to a wider sample of jet parameters, including magnetic fields and a more realistic equation of state." }, "1005/1005.4618_arXiv.txt": { "abstract": "We present additional evidence that dust is really forming along the red giant branch (RGB) of 47~Tuc at luminosities ranging from above the horizontal branch to the RGB-tip \\citep{ori07}. The presence of dust had been inferred from an infrared excess in the $(K-8)$ color, with $K$ measured from high spatial resolution ground based near-IR photometry and ``8'' referring to Spitzer-IRAC 8\\micron\\ photometry. We show how $(K-8)$ is a far more sensitive diagnostic for detecting tiny circumstellar envelopes around warm giants than colors using only the Spitzer-IRAC bands, for example the $(3.6-8)$ color used by \\citet{boy10}. In addition, we also show high resolution HST-ACS $I$ band images of the giant stars which have $(K-8)$ color excess. These images clearly demonstrate that \\citet{boy10} statement that our detections of color excess associated with stars below the RGB-tip arise from blends and artefacts is simply not valid. ", "introduction": "\\label{intro} In \\citet{ori07} we presented estimated mass loss rates for first ascent red giant branch (RGB) stars in the globular cluster 47~Tuc. These were based on near and mid-infrared photometry obtained from a Spitzer-IRAC (3.6, 4.5, 5.6 \\& 8\\,\\micron) survey and ground-based high resolution $JHK$ observations. We found about 100 giants with ($K-8$) color excess that we attributed to the presence of dusty circumstellar envelopes. These candidate dusty stars were mainly found in the inner 2\\arcmin\\ (in radius) and had luminosities ranging from above the level of the Horizontal Branch (HB) to the Red Giant Branch (RGB)-tip. For a given luminosity only a fraction of stars exhibited this color excess, and this fraction increases toward the tip of the RGB. From the color excess we derived a mass loss rate and found a shallower dependence on luminosity than that expected from Reimers \\citep{rei75a,rei75b} formula. After correcting the observed frequency of dusty envelopes for incompleteness, we derived an average duty cycle which could be used along with the observed mass loss rates and evolutionary times to compute the total amount of mass lost on the RGB. Recently, \\citet{boy10}, using the $(3.6-8)$ Spitzer-IRAC color as their main diagnostic tool, found evidence for a dust excess only in asymptotic giant branch stars and possibly a few giants near the RGB tip. On the basis of this finding, they concluded \\citep[see also ][]{boy08,boy09} that our candidate dusty giants below $M_{\\rm bol} = -2.5$ were spurious, mainly blends and/or artefacts. We emphasize that {\\it i}) this conclusion was not based on a direct star-to-star check, {\\it ii}) their analysis employed a different diagnostic, and it was not optimized to cover the innermost region of the cluster sampled by our work. The AKARI analysis of 47~Tuc presented by \\citet{ita07} also showed dust excess only near the tip, but it had not the spatial resolution to properly investigate the stellar population in central regions of globular clusters. We ourselves using ISOCAM found circumstellar dust excess only near the RGB tip \\citep{ori02}, again because of the lower spatial resolution of ISOCAM compared to IRAC. In this Letter we respond to the \\citet{boy10} criticism. We demonstrate the importance of an optimum sampling of the innermost region of the cluster where most of the red giants, along with the rest of the stellar population, are found. We show that $(K-8)$ is a much better diagnostic of dusty giants than $(3.6-8)$. Finally, using a high resolution HST-ACS $I$ band image, we clearly demonstrate that our detected $(K-8)$ color excess in giant stars below the RGB tip is not an effect of blending and/or an artefact. ", "conclusions": "The $(K-8)$ color excess detected by \\citet{ori07} in a fraction of RGB stars down to $\\mbol \\sim 0$ within the central 2\\arcmin\\ of 47~Tuc is real. It is not an artefact, since blends and other possible spurious effects have been properly accounted for by cross-correlating the Spitzer sources with high spatial resolution optical and near infrared catalogs of stellar counterparts. The fact that \\citet{boy10} did not find as many dusty stars in the central 2\\arcmin\\ is mainly due to the use of a different diagnostic, namely the $(3.6-8)$ color, which is only effective for detecting dust excess in cooler (hence more luminous) giants. Further, their photometric analysis is not optimized in the central, densest region. \\citet{boy10} find only few dusty giants in the outer regions, fully consistent with \\citet{ori07} and the additional CMDs shown in Figure 2. As discussed in Section 4 of \\citet{ori07}, mass loss rates and duty cycles decrease with decreasing stellar luminosity. For a given luminosity, the estimated rates are higher (between a factor of 2 near the tip up to 2 orders of magnitude down to $\\mbol \\sim 0$ than those predicted by the Reimers law, as also noted by \\citet{boy10}. The shallower slope we found has been also suggested by \\citet{mes09}, from chromospheric line wind diagnostics of giants in metal poor clusters. The Reimers result was obtained from Population~I objects, and one might anticipate some differences. Our higher mass loss rate is not a major problem, given that both the Reimers law and our results contain free parameters, {\\it i}i) the Reimers efficiency $\\eta$, and {\\it ii}) the gas to dust ratio and expansion velocity. In either case the free parameters must be set by indirect observational constraints on the total mass lost during the RGB evolution like the HB morphology. The total mass loss occurs predominantly in the upper $\\sim2$ magnitudes near the RGB tip. The contribution of the low luminosity giants ($\\mbol > -1$) is small ($<20$\\%) to negligible, and within the estimated uncertainty. The importance of our observation of mass loss in less luminous stars is not its impact on total mass lost, but rather the clue it gives us to the physics of mass loss. Likewise, the episodic nature of mass loss in non-variable stars tells us something. In both cases, it seems quite clear that the underlying driver of the mass loss is {\\em not} radiation pressure on dust. Another important outcome of our larger project will be an investigation of the differential mass loss among clusters with different metallicities and HB morphologies. This has been the goal of our Spitzer survey and the complementary ground based observations, and the results will appear in forthcoming papers." }, "1005/1005.1451_arXiv.txt": { "abstract": "Using the state-of-the-art cosmological hydrodynamic simulations of the standard cold dark matter model with star formation feedback strength normalized to match the observed star formation history of the universe at $z=0-6$, we compute the metal enrichment history of the intergalactic medium (IGM). Overall we show that galactic superwind (GSW) feedback from star formation can transport metals to the IGM and that the properties of simulated metal absorbers match current observations. The distance of influence of GSW from galaxies is typically limited to about $\\le 0.5$Mpc and within regions of overdensity $\\delta \\ge 10$. Most \\civ and \\ovi absorbers are located within shocked regions of elevated temperature ($T\\ge 2\\times 10^4$K), overdensity ($\\delta \\ge 10$), and metallicity ($[Z/\\zsun]=[-2.5,-0.5]$), enclosed by double shocks propagating outward. \\ovi absorbers have typically higher metallicity, lower density and higher temperature than \\civ absorbers. For \\ovi absorbers collisional ionization dominates over the entire redshift range $z=0-6$, whereas for \\civ absorbers the transition occurs at moderate redshift $z\\sim 3$ from collisionally dominated to photoionization dominated. We find that the observed column density distributions for \\civ and \\ovi in the range $\\log N {\\rm cm}^2=12-15$ are reasonably reproduced by the simulations. The evolution of mass densities contained in \\civ and \\ovi lines, $\\omegaciv$ and $\\omegaovi$, is also in good agreement with observations, which shows a near constancy at low redshifts and an exponential drop beyond redshift $z=3-4$. For both \\civ and \\ovi\\, most absorbers are transient and the amount of metals probed by \\civ and \\ovi lines of column $\\log N {\\rm cm}^2=12-15$ is only $\\sim 2\\%$ of total metal density at any epoch. While gravitational shocks from large-scale structure formation dominate the energy budget ($80-90\\%$) for turning about 50\\% of IGM to the warm-hot intergalactic medium (WHIM) by $z=0$, GSW feedback shocks are energetically dominant over gravitational shocks at $z\\ge 1-2$. Most of the so-called ``missing metals\" at $z=2-3$ are hidden in a warm-hot ($T=10^{4.5-7}$K) gaseous phase, heated up by GSW feedback shocks. Their mass distribution is broadly peaked at $\\delta=1-10$ in the IGM, outside virialized halos. Approximately $(37,46,10,7)\\%$ of the total metals at $z=0$ are in (stars, WHIM, X-ray gas, cold gas); the distribution stands at $(23,57,2,18)\\%$ and $(14,51,4,31)\\%$ at $z=2$ and $z=4$, respectively. ", "introduction": "One of the pillars of the Big Bang theory is its successful prediction of a primordial baryonic matter composition, made up of nearly one hundred percent hydrogen and helium with a trace amount of a few other light elements \\citep[e.g.,][]{1998Schramm,2001Burles}. The metals, nucleosynthesized in stars later, are found almost everywhere in the observable IGM, ranging from the metal-rich intracluster medium \\citep[e.g.,][]{1997Mushotzky} to moderately enriched damped Lyman systems \\citep[e.g.,][]{1997Pettini, 2003Prochaska} to low metallicity Lyman alpha clouds \\citep[e.g.,][]{2003Schaye}. When and where were the metals made and why are they distributed as observed? We address this fundamental question in the context of the standard cold dark matter cosmological model \\citep[][]{2009Komatsu} using latest simulations. Our previous simulations \\citep[][]{1999bCen, 2005Cen} provided some of the earlier attempts to address this question with measured successes. In this investigation we use substantially better simulations to provide significantly more constrained treatment of the feedback processes from star formation (SF) that drive energy and metals from supernovae into the IGM through galactic winds \\citep[e.g.,][]{1999bCen, 2001Aguirre, 2002bTheuns, 2003Adelberger, 2003Springel}. Metal-line absorption systems in QSO spectra are the primary probes of the metal enrichment of the IGM as well as in the vicinities of galaxies \\citep[e.g.,][]{1969Bahcall}. The most widely used metal lines include \\ion{Mg}{2} $\\lambda \\lambda$2796, 2803 doublet \\citep[e.g.,][]{1992Steidel}, \\ion{C}{4} $\\lambda \\lambda$1548, 1550 doublet \\citep[e.g.,][]{1982Young}, and \\ion{O}{6} $\\lambda \\lambda$1032, 1038 doublet \\citep[e.g.,][]{2002Simcoe}. We here focus on the \\civ and \\ovi absorption lines and the global evolution of metals in the IGM. We will limit our current investigation to the observationally accessible redshift range of $z=0-6$, which in part is theoretically motivated simply because the theoretical uncertainties involving still earlier star formation are much larger. At $z=0$ the \\ovi line (together with \\ion{C}{7} and \\ion{O}{8} lines) provide vital information on the missing baryons \\citep[e.g.,][]{2003Mathur, 2008Tripp, 2008Danforth, 2009Nicastro}, predicted to exist in a Warm-Hot Intergalactic Medium (WHIM) \\citep[][]{1999bCen, 2001Dave}. For a well understood sample of QSO absorption lines, one could derive the cosmological density contained in them \\citep[e.g.,][]{2009Cooksey}. Early investigations indicate that $\\omegaciv$ remains approximately constant in the redshift interval $z \\sim 1.5 - 4$ \\citep[][]{2001Songaila,2005Songaila,2003Boksenberg}. There have been recent efforts to extend the measurements of $\\omegaciv$ to $z<1.5$ \\citep[][]{2009Cooksey} and to $z>5$ \\citep[][]{2006Simcoe,2006Ryanweber, 2009Ryanweber,2009Dodorico, 2009Becker}. Observations in these redshift ranges have been difficult to carry out because \\civ transition moves to the UV at low redshift and to the IR band at high redshift. \\citet[][]{2009Dodorico} find evidence of a rise in the \\civ mass density for $z<2.5$. \\citet[][]{2006Simcoe} and \\citet[][]{2006Ryanweber} found evidence of \\civ density at $z\\sim 6$ being consistent with estimations at $z \\sim 2-4.5$. More recently, however, \\citet[][]{2009Becker} set upper limits for $\\omegaciv$ at $z\\sim 5.3$ and \\citet[][]{2009Ryanweber} observe a decline in intergalactic \\civ approaching $z=6$, which we will show are in good agreement with our simulations. The ionization potential of \\ovi and the relatively high oxygen abundance are very favorable for production of \\ovi absorbers in the IGM \\citep[e.g.,][]{1983Norris, 1986Chaffee}. The rest wavelength of OVI ($1032,1037$\\AA) places it within the Ly-$\\alpha$ forest, which makes the identifications of these lines more complicated, although being a doublet helps significantly. At $z\\ge 2 $, however, \\ovi absorption can probe the metal content of the IGM in ways complementary to what is provided by \\civ lines. For example, the \\ovi lines can probe IGM that is hotter than that probed by the \\civ lines and can reach lower densities thank to higher abundance. There are now several observational studies at redshifts $z=2-3$ that describe the properties of \\ovi absorbers and attempt to estimate the \\ovi mass density, $\\omegaovi$ \\citep[][]{2002Carswell,2002Bergeron, 2004Simcoe, 2006Simcoe, 2008Frank, 2008Danforth, 2008Tripp, 2008Thomb}. At $z\\sim 2-3$ there is a missing metals problem: only 10-20\\% of the metals produced by all stars formed earlier have been identified in stars of Lyman break galaxies (LBG), in damped Lyman alpha systems (DLAs) and $\\lya$ forest. The vast majority of the produced metals appear to be missing \\citep[e.g.,][]{1999bPettini}. The missing metals could be in hot gaseous halos of star-forming galaxies \\citep[][]{1999bPettini, 2005Ferrara}. We will show that most of the missing metals are in a warm-hot ($T=10^{4.5-7}$K) but diffuse IGM at $z=2-3$ of overdensities of $\\sim 10$ that are outside of halos. The outline of this paper is as follows. In \\S 2 we detail our simulations and the procedure of normalizing the uncertain feedback processes from star formation. Results on the metal enrichment of the IGM are presented in \\S 3. In \\S 3.1 we give a full description of the properties of the \\civ and \\ovi lines at $z=0-6$, followed \\S 3.2 discussing \\civ and \\ovi absorbers as metals reservoirs. We devote \\S 3.3 to a general discussion of global distribution of metals, addressing several specific topics, including the metallicity of the moderate overdense regions at moderate redshift, the missing metals at $z\\sim 3$. Conclusions are given in \\S 4. ", "conclusions": "We have carried out the state-of-the-art cosmological hydrodynamic simulations of the standard cold dark matter model to investigate the process of metal enrichment of the intergalactic medium. Our simulations have substantially higher resolution than our previous simulations to address this problem. More importantly, we can now constrain the strength of the feedback process by matching the star formation history in our current simulation to the observed one in the range $z=0-6$. We find that our model reproduces the observed mean flux of the $\\lya$ forest and the mass density of \\civ and \\ovi absorbers. It is also in general consistent with observed physical properties of absorbers. This indicates that we can explain the metal enrichment of the IGM by considering star formation to be the main feedback mechanism, with no apparent need of significant contribution from AGN in terms of additional energy. We conclude from our results that: (1) The overall star formation history depends rather sensitively on the feedback strength. This is likely due to GSW significantly reducing the concentration of cold gas around halos. Nevertheless, GSW do not significantly alter the overall large-scale filamentary baryonic structure that follows the cosmic web of dark matter distribution. While GSW could travel far into low density regions sometimes, the amount of energy and metals that are deposited in underdense regions is very small. Most of the GSW energy and metals remain in regions of overdensity $\\delta \\ge 10$, with the distance of influence of GSW from galaxies limited to about $\\le 0.5$Mpc. Metal bubbles blown by GSW coincide with temperatures bubbles, suggesting a tight coupling of energy and metal deposition, and they are terminated by shock fronts. (2) Both \\civ and \\ovi absorbers are located in regions that have been swept by feedback shocks, of elevated temperature ($T\\ge 2\\times 10^4$K), density ($\\delta \\ge 10$) and metallicity ($[Z/\\zsun]=[-2.5,-0.5]$), demarcated by a double shock propagating outwards, with \\ovi absorbers typically having a higher metallicity than \\civ absorbers. Within these shocked regions, most of \\civ absorbers tend to arise from moderate density peaks that are troughs in temperature and are thus relatively quiescent. The \\ovi absorbers are from regions that are dynamically hotter near shock fronts. There is a trend for the population of \\civ and \\ovi absorbers to be more collisionally ionized at higher redshift; for \\ovi collisional ionization dominates over the entire redshift range $z=0-6$, whereas for \\civ the transition occurs at moderate redshift $z\\sim 3$ from collisionally dominated to photoionization dominated. (3) The evolution of the mass density contained in \\civ and \\ovi lines, $\\omegaciv$ is in good agreement with observations, with both the latest observations and simulations of $\\omegaciv$ exhibiting an exponential drop beyond redshift $z=4$; $\\omegaciv$ drop exponentially beyond redshift $z=3$; the near constancy of $\\omegaciv$ at redshift $z=1-3$ does not reflect the evolution of the overall metal content in the IGM. In the case of $\\omegaovi$, we find a less good agreement between observations and out results. This might be in part due to cosmic variance. (4) Most of \\civ and \\ovi absorbers, while clustered around galaxies, are transient and intergalactic in origin, produced by galactic superwinds in the process of transporting both energy and metals from galaxies into the IGM; the metal mass densities contained in \\civ and \\ovi lines in the range $\\log N {\\rm cm}^2=12-15$ each constitutes $\\sim 0.1\\%$ of total metal density at all redshifts; the amount of metals probed by \\civ and \\ovi lines in the range $\\log N {\\rm cm}^2=12-15$ is $\\sim 1\\%$ of the total metal density at all redshifts. (5) While gravitational shocks from large-scale structure formation dominate the energy budget ($80-90\\%$) for turning about $50\\%$ of IGM to the warm-hot intergalactic medium (WHIM) by $z=0$, galactic superwind feedback shocks are energetically dominant over gravitational shocks at $z\\ge 1-2$. (6) Most of the so-called ``missing metals\" at $z=2-3$ are hidden in a warm-hot gaseous phase ($T>3\\times 10^4$K) that is heated up by star formation feedback shocks. Their mass distribution is broadly peaked at overdensity $1-10$ in the IGM, outside virialized halos. Approximately $(37,46,10,7)\\%$ of the total metals at $z=0$ are in (stars, WHIM, X-ray gas, cold gas); the distribution stands at $(23,57,2,18)\\%$ and $(14,51,4,31)\\%$ at $z=2$ and $z=4$, respectively. (7) The metallicity of the IGM with moderate overdensities ($1-10$) that are probed by the $\\lya$ forest shows a rapid increase with decreasing redshift. We show that velocity ``diffusion'' effect that arises from the peculiar velocities could enhance the ``apparent\" metallicity of the $\\lya$ forest clouds, as supported by our cross-correlation analysis. Tentatively, we suggest that this may reconcile, at least in part, the discrepancy between our simulations and observations at $z=2-4$ based on pixel optical depth (POD) method." }, "1005/1005.0382_arXiv.txt": { "abstract": "{ Past and current X-ray mission allow us to observe only a fraction of the volume occupied by the ICM. After reviewing the state of the art of cluster outskirts observations we discuss some important constraints that should be met when designing an experiment to measure X-ray emission out to the virial radius. From what we can surmise \\wfxt ~ is already designed to meet most of the requirements and should have no major difficulty in accommodating the remaining few. ", "introduction": "Galaxy clusters form through the hierarchical accretion of cosmic matter. The end products of this process are virialized structures that feature, in the X-ray band, similar radial profiles of the surface brightness $S_{\\rm b}$ (e.g. Vikhlinin et al. 1999, Neumann 2005, Ettori \\& Balestra 2009) and of the plasma temperature $T_{\\rm gas}$ (e.g. Allen et al. 2001, Vikhlinin et al. 2005, Leccardi \\& Molendi 2008). Such measurements have definitely improved in recent years thanks to the arcsec resolution and large collecting area of the present X-ray satellites, like \\chandra\\ and \\xmm, but still remain difficult, in particular in the outskirts, because of the low surface brightness associated to these regions. Present observations provide routinely reasonable estimates of the gas density, $n_{\\rm gas}$, and temperature, $T_{\\rm gas}$, up to about $R_{2500}$ ($\\approx 0.3 R_{200}$; $R_{\\Delta}$ is defined as the radius of the sphere that encloses a mean mass density of $\\Delta$ times the critical density at the cluster's redshift; $R_{200}$ defines approximately the virialized region in galaxy clusters). Only few cases provide meaningful measurements at $R_{500}$ ($\\approx 0.7 R_{200}$) and beyond (e.g. Vikhlinin et al. 2005, Leccardi \\& Molendi 2008, Neumann 2005, Ettori \\& Balestra 2009). Consequently, more than two-thirds of the typical cluster volume, just where primordial gas is accreting and dark matter halo is forming, is still unknown for what concerns both its mass distribution and its thermodynamical properties. This poses a significant limitation in our ability to characterize the physical processes presiding over the formation and evolution of clusters and to use clusters as cosmological tools, as also outlined in the Scientific Justification for the \\wfxt\\ (Giacconi et al. 2009). Indeed the characterization of thermodynamic properties at large radii would allow us to provide constraints on the virialization process, while measures of the metal abundance would allow us to gain insight on the enrichment processes occurring in clusters (e.g. Fabjan et al. 2010). Morever the X-ray emission at large radii could also be used to improve significantly measures of the gas and total gravitating masses thereby opening the way to a more accurate use of galaxy clusters as cosmological probes (e.g. Voit 2005). In these proceedings, we take stock of the situation on cluster outskirts and suggest how to make progress. In Sect.~\\ref{sec: obs}, we provide an observational overview of currently available measures of cluster outer regions, while in Sect.~\\ref{sec: map} we discuss some important constraints that should be met when designing an experiment to measure X-ray emission out to the virial radius. In Sect.~\\ref{sec: future}, we present an overview of future missions which have cluster outskirts observations as one of their goals, our main results are recapitulated in Sect.~\\ref{sec: summary}. \\begin{figure*} \\begin{center} \\includegraphics[height=4.5cm]{eb09_fig2.ps} \\end{center} \\caption{\\small {\\bf From left to right}: Example of a surface brightness profile with the fitted background ({\\it horizontal dotted line}) and the radius $R_{200}$ ({\\it vertical dashed line}); the signal-to-noise profile evaluated as $S2N = (S_b - B)/ \\epsilon$, where the error $\\epsilon$ is the sum in quadrature of the Poissonian error in the radial counts and the uncertainties in the fitted background, $B$; the best-fit values of the slope of the surface brightness profile as a function of $r / R_{200}$. These values are estimated over 6 radial bins (thick horizontal solid line: the slope evaluated between $0.4 \\times R_{200}$ and $R_{S2N}$ with a minimum of 3 radial bins; dashed line: best-fit of $d \\ln (S_b) / d \\ln (r/R_{200})$ with the functional form $s_0 +s_1 \\ln (r/R_{200})$ over the radial range $0.1 \\times R_{200} - R_{S2N}$, with the best-fit parameters quoted in Table~3 of Ettori \\& Balestra 2009). } \\label{fig:s2n} \\end{figure*} A Hubble constant of 70 $h_{70}$ km s$^{-1}$ Mpc$^{-1}$ in a flat universe with $\\Omega_{\\rm m}$ equals to 0.3 is assumed throughout this manuscript. ", "conclusions": "\\label{sec: summary} Past and current X-ray mission allow us to observe only a fraction of the volume occupied by the ICM. Indeed, typical measures of the surface brightness, temperature and metal abundance extend out to a fraction of the virial radius. The coming into operation of the second generation of medium energy X-ray telescopes at the turn of the millennium, has resulted in relatively modest improvements in our ability to characterize cluster outskirts. Even though recent results from \\suzaku\\ show some improvement, the most sensitive instrument to low surface brightness to have flown thus far is quite possibly the \\swift\\ XRT which, ironically, never had cluster outer regions as one of its top scientific objectives. The construction of an experiment capable of making measures out to $R_{200}$ is well within the reach of currently available technology. What is required is an experiment design that will minimize the background, both instrumental and cosmic, and maximizes the grasp, i.e. the product of effective area and FOV. Since cluster emission in the outskirts will be background dominated, instrument design and observational strategy should also allow for a meticulous characterization of the background. Detailed simulations based on realistic estimates of the different spectral components and of the precision with which the may be determined shows that an experiments such as the one we envisage will allow a solid characterization of cluster outskirts. From what we can surmise \\wfxt\\ is already designed to meet most of the requirements which are necessary to characterize cluster outskirts, and should have no major difficulty in accommodating the remaining few." }, "1005/1005.0985_arXiv.txt": { "abstract": "{% We present results of our 5-years-long program of ground-based spectroscopic and photometric observations of individual Kepler asteroseismic targets and the open clusters NGC\\,6866 and NGC\\,6811 from the Kepler field of view. We determined the effective temperature, surface gravity, metallicity, the projected rotational velocity and the radial velocity of 119 Kepler asteroseismic targets for which we acquired high-resolution spectra. For many of these stars the derived atmospheric parameters agree with $T_{\\rm eff}$, $\\log g$, and [Fe/H] from the Kepler Input Catalog (KIC) to within their error bars. Only for stars hotter than 7000\\,K we notice significant differences between the effective temperature derived from spectroscopy and $T_{\\rm eff}$ given in the KIC. For 19 stars which we observed photoelectrically, we measured the interstellar reddening and we found it to be negligible. Finally, our discovery of the $\\delta$ Sct and $\\gamma$ Dor pulsating stars in the open cluster NGC\\,6866 allowed us to discuss the frequency of the occurrence of $\\gamma$ Dor stars in the open clusters of different age and metallicity and show that there are no correlations between these parameters. } ", "introduction": "Our program of ground-based spectroscopic and photometric observations of stars selected for the asteroseismic targets for the Kepler space telescope by the Kepler Asteroseismic Science Consortium KASC\\footnote{Kepler Asteroseismic Science Consortium (KASC) is a group of collaborating scientists and/or institutions established to accomplish the activities of the Kepler Asteroseismic Investigation (KAI), represented by Ronald Gilliland (see http://astro.phys.au.dk/KASC).} was started in 2005 at the Osservatorio Astrofisico di Catania, OACt, (the {\\it M.G. Fracastoro\\/} station, Mt.\\ Etna, Italy) and is continued since then. Apart from the OACt, we perform spectroscopic and photometric observations of Kepler asteroseismic targets at the F.\\ L.\\ Whipple Observatory, FLWO, (Mount Hopkins, Arizona, USA), and the Astrophysical Observatory of the University of Wroc\\l{}aw in Bia\\l{}k\\'ow (Poland). At the OACt, we use a 91-cm telescope, at FLWO, a 1.5-m telescope, and at the Bia\\l{}k\\'ow Observatory, a 60-cm telescope. We make use also of the archival data collected at the 1.5-m telescope at the Oak Ridge Observatory, ORO, (Harvard, Massachusetts, USA) and at the Multiple Mirror Telescope, MMT, before it was converted to the monolithic 6.5-m mirror. Using the spectroscopic data acquired at OACt, FLWO, ORO, and MMT, we aim at the determination of the atmospheric parameters of the program stars, i.e., the effective temperature, $T_{\\rm eff}$, surface gravity, $\\log g$, and metallicity, $\\rm [Fe/H]$, and measuring the projected rotational velocity, $v\\sin i$, and the radial velocity, $v_r$, of the stars. At the Bia\\l{}k\\'ow Observatory, we perform photometric time-series observations of the open clusters NGC\\,6866 and NGC\\,6811 in the Kepler field of view, aiming at the discovery of new pulsating stars, and the determination of the degree of the modes of their pulsations. Both sites, the Bia\\l{}k\\'ow Observatory and the OACt, took part in the international multi-site photometric campaign on NGC\\,6866 launched in 2009, and will take part in a similar campaign on NGC\\,6811 which will be launched in 2010 (for the details, see Uytterhoeven et al. 2010b.) \\begin{figure} \\includegraphics[width=60mm,height=80mm,angle=270]{mol_fig2.ps} \\caption{The Kepler asteroseismic targets observed at the Osservatorio Astrofisico di Catania, the F.\\ L.\\ Whipple Observatory, and the Oak Ridge Observatory. Indicated are the borders of the 42 Kepler CCDs and the borders of the constellations.} \\label{targets} \\end{figure} ", "conclusions": "" }, "1005/1005.2722_arXiv.txt": { "abstract": "Solar flares are currently understood as the explosive release of energy stored in the form of stressed magnetic fields. In many cases, the released energy seems to take the form of large numbers of electrons accelerated to high energies (the nonthermal electron ``thick target'' model), or alternatively plasma heated to very high temperatures behind a rapidly moving conduction front (the ``thermal'' model). The transport of this energy into the remaining portion of the atmosphere results in violent mass motion and strong emission across the electromagnetic spectrum. Radiation processes play a crucial role in determining the ensuing plasma motion. One important phenomenon observed during flares is the appearance in coronal magnetic loops of large amounts of upflowing, soft X-ray emitting plasma at temperatures of $1-2\\ \\times 10^7$ [K]. It is believed that this is due to chromospheric evaporation, the process of heating cool ($T \\sim 10^4$ [K]) chromospheric material beyond its ability to radiate. Detailed calculations of thick target heating show that if nonthermal electrons heat the chromosphere directly, then the evaporation process can result in explosive upward motion of X-ray emitting plasma if the heating rate exceeds a threshold value. In such a case, upflow velocities approach an upper limit of roughly $2.35 c_s$ as the heating rate is increased beyond the threshold, where $c_s$ is the sound speed in the evaporated plasma. This is known as explosive evaporation. If the flare heating rate is less than the threshold, evaporation takes place indirectly through thermal conduction of heat deposited in the corona by the energetic electrons. Upflows in this case are roughly 10 to 20\\% of the upper limit. Evaporation by thermal model heating always takes place through thermal conduction, and the computed upflow speeds seem to be about 10\\% to 20\\% of the upper limit, independent of the energy flux. The pressure increase in the evaporated plasma for either the thick target or thermal model leads to a number of interesting phenomena in the flare chromosphere. The sudden pressure increase initiates a downward moving ``chromospheric condensation'', an overdense region which gradually decelerates as it accretes material and propagates into the gravitationally stratified chromosphere. Solutions to an equation of motion for this condensation shows that its motion decays after about one minute of propagation into the chromosphere. When the front of this downflowing region is supersonic relative to the atmosphere ahead of it, a radiating shock will form. If the downflow is rapid enough, the shock strength should be sufficient to excite UV radiation normally associated with the transition region, and furthermore, the radiating shock will be brighter than the transition region. These results lead to a number of observationally testable relationships between the optical and ultraviolet spectra from the condensation and radiating shock. ", "introduction": " ", "conclusions": "" }, "1005/1005.5102_arXiv.txt": { "abstract": "\\noindent We calculate solar models including dark matter (DM) weakly-interacting massive particles (WIMPs) of mass 5-50\\gev\\;and test these models against helioseismic constraints on sound speed, convection zone depth, convection zone helium abundance, and small separations of low-degree p-modes. Our main conclusion is that both direct detection experiments and particle accelerators may be complemented by using the Sun as a probe for WIMP DM particles in the 5-50\\gev\\;mass range. The DM most sensitive to this probe has suppressed annihilations and a large spin-dependent elastic scattering cross section. For the WIMP cross-section parameters explored here, the lightest WIMP masses $<$\\,10\\gev\\;are ruled out by constraints on core sound speed and low-degree frequency spacings. For WIMP masses 30-50\\gev, the changes to the solar structure are confined to the inner 4\\% of the solar radius and so do not significantly affect the solar p-modes. Future helioseismology observations, most notably involving g-modes, and future solar neutrino experiments may be able to constrain the allowable DM parameter space in a mass range that is of current interest for direct detection. ", "introduction": "\\noindent We explore the role of DM WIMPs in modifying the thermal gradient of the Sun. A similar study involving standard WIMPS of mass $50\\gev$ or larger has been performed by Bottino et al. \\cite{Bottino}. However, here we wish to consider the effects of low mass WIMPs, with masses as low as $5\\gev$, with large trapped abundances within the Sun. To affect the Sun's thermal gradient, we need large elastic scattering rates. The solar sound speed can be affected, and helioseismology has been proposed as providing a possible constraint on supersymmetric WIMPs \\cite{lopesa,lopesb, lopesc}. For the range in masses in which we are interested, the limits on the size of spin-independent WIMP elastic cross sections from CRESST \\cite{cresst}, XENON-10 \\cite{xenon10}, and XENON-100 \\cite{xenon100} are already quite stringent making it unlikely for helioseismology to provide further restrictions. Moreover, recent COUPP results \\cite{Behnke} (especially at masses $\\gsim 10$\\gev) and the results from PICASSO \\cite{picasso} similarly restrict spin-dependent interactions. However, these limits become weaker as the DM mass is decreased, especially for masses of around $5$\\gev or less. Detailed solar models \\cite{Bottino} show that a WIMP signal is only possible for cross sections in a limited range. For a DM particle mass of $50 \\gev$ the relevant effective cross section for these signals was found to be of order $10^{-35}\\,{\\rm cm}^2$. The limits on spin-dependent scattering from direct detection experiments, e.g., COUPP \\cite{Behnke, coupp} restrict the spin-dependent elastic scattering cross section for a $50 \\gev$ DM particle to less than $\\sim10^{-37}\\,{\\rm cm}^2$. Moving to lighter masses alleviates these limits somewhat, and for around $5 \\gev$ we are in the interesting region of around $10^{-35}\\,{\\rm cm}^2$ \\cite{picasso}. In addition, given the astrophysical uncertainties that can affect these limits (e.g.\\,\\cite{green}) it is intriguing to ask whether helioseismology can complement direct detection limits at lower masses. Recently, there has been interest in exploring the low DM mass regime as a possible way to consistently combine the results from the DAMA direct detection experiment \\cite{dama} with those from others such as CDMS \\cite{cdms} and CoGeNT \\cite{cogent} (see e.g., \\cite{lowmasssector}). Although we do not attempt to do the same here, we simply note that this mass regime is of great interest with upper limits on the spin-dependent elastic scattering cross section for low-mass DM, reaching $\\sim10^{-32}\\,{\\rm cm}^2$ \\cite{Kopp} for certain models and assumptions. Solar effects are most pronounced for DM with a suppressed annihilation cross section such that after capture by the Sun, the DM candidates do not annihilate quickly. A prominent example of this is asymmetric DM where annihilation is completely suppressed. In this paper, we outline a class of models for WIMPs that are capable of modifying the temperature profile in the core of the Sun, and illustrate their effects on helioseismology and neutrino fluxes. The accumulation of these WIMPs in the solar core results in significant energy transfer to solar protons. We note that the effect is not large enough to account for discrepancies between observations and helioseismology for models that also predict the observed neutrino flux. However, given the current debate about the appropriate element abundances to be adopted in solar models \\cite{serenelli}, this effect may still play a role and should be included in the models. Indeed the recently revised solar abundances result in solar models that cannot reproduce currently observed helioseismic data \\cite{AGS05, AGSS09}. The effects on the Sun of low-mass asymmetric WIMPs possessing large self-interactions have been considered in \\cite{Frandsen}. While in \\cite{Frandsen} the authors focus on WIMPs with spin-independent interactions, in this study we specifically focus on WIMPs with purely axial interactions and consequently only spin-dependent elastic scattering. In the following sections, we outline the main features of DM models that can be potentially probed using solar properties and explore their effect on solar models. ", "conclusions": "Our main conclusion is that both direct detection and accelerator probes may be complemented by using the Sun as a probe of DM. Models of DM that have large spin-dependent interactions and an intrinsic asymmetry that prevents post freeze-out annihilations can significantly lower the central temperature of the Sun as well as the resulting $^8$B neutrino flux. For WIMP masses $m_{\\chi}>10\\gev$, the presence of WIMPs does not significantly affect currently available helioseismic constraints. However, for WIMP masses of 10\\,GeV or lighter, constraints on sound speed and small frequency separations between $\\ell=0$ and $\\ell=2$ p-modes can begin to constrain and rule out the presence of WIMPs with the cross sections utilised here. Our study is motivated in a large part by the recently revised solar abundances \\cite{AGS05, AGSS09} which result in solar models that cannot reproduce the currently observed helioseismic data, with numerous attempts to restore agreement being met with only partial success (see e.g., \\cite{BA08, GM10}). This means that additional physics must be incorporated into solar modelling, and dark matter is among the options that merit detailed consideration. Since the original submission of our paper in May 2010, an additional paper appeared on solar models including WIMPs and the implications for reconciling the new solar abundances with helioseismology \\cite{Taoso10}. In agreement with \\cite{Taoso10}, our explorations to date do not show any realistic path in which the inclusion of WIMPs will mitigate this problem. Even for the large interaction cross section and small annihilation cross section considered here, the inclusion of WIMPs of mass $m_{\\chi}>10\\gev$ has little effect on presently observable helioseismic signatures. The inclusion of WIMPs with masses of 10\\,GeV or lighter worsens the agreement with the helioseismically-inferred sound speed at radii $0.1\\le r\\le0.2\\,R_{\\odot}$, only slightly deepens the predicted convection-zone depth, and introduces a trend with radial order in the low-degree p-mode small separations that is not observed in the data. Our primary new result is that WIMP masses of $\\sim 5$ GeV may be excluded for spin-dependent interactions in a specified cross-section range, thereby complementing direct detection experiments in a region that they access only with great difficulty provided the WIMPs annihilation cross section is suppressed. While here we do not discuss whether these WIMP models could accommodate measurements of the $^8$B neutrino flux from current solar neutrino experiments such as Super-Kamiokande III \\cite{Yang:2009hp}, SNO \\cite{Collaboration:2009qz} or Borexino \\cite{Collaboration:2008mr}, with precisions of $\\sim$10\\% and theoretical expectations of up to $\\sim$20\\% depending on the solar composition \\cite{Bottino, Frandsen}, a more detailed study of the low-mass region of WIMP parameter space and its consistency with current experimental data is deferred to a later paper. There we will address the question of whether future helioseismic observations, most notably using g-modes, and solar neutrinos, may be able to constrain the allowable DM parameter space in a mass range that is of current interest for direct detection. Finally, we note that for solar mass stars near the centre of the Galaxy, where the WIMP density is enhanced by up to some 6 orders of magnitude relative to that in the solar neighbourhood, the effect of the redistribution of energy in the stellar core may generate a significant reduction of the main-sequence lifetimes. We leave an investigation of this scenario to our future work." }, "1005/1005.0511_arXiv.txt": { "abstract": "We compare the radial locations of 178 core-collapse supernovae to the $R$-band and H$\\alpha$ light distributions of their host galaxies. When the galaxies are split into `disturbed' and `undisturbed' categories, a striking difference emerges. The disturbed galaxies have a central excess of core-collapse supernovae, and this excess is almost completely dominated by supernovae of types Ib, Ic and Ib/c, whereas type II supernovae dominate in all other environments. The difference cannot easily be explained by metallicity or extinction effects, and thus we propose that this is direct evidence for a stellar initial mass function that is strongly weighted towards high mass stars, specifically in the central regions of disturbed galaxies. ", "introduction": "\\label{sec:intro} Following the pioneering work of \\citet{lars78}, many studies have confirmed that tidal disturbance following galaxy interactions is an efficient trigger of star formation in galaxies \\citep[e.g.][]{jose84,kenn84}. Such star formation frequently takes the form of centrally-concentrated nuclear starbursts \\citep{jose85}, fuelled by the central concentrations of molecular gas found to occur naturally in simulations of highly-disturbed systems \\citep{barn91,miho96}. The strength of the link between starbursts and interactions was highlighted by the finding that almost all of the `ultra-luminous infrared galaxies' (ULIRGs) display signs of interactions or mergers \\citep{sand88,born99}, and by correlations between galaxy-galaxy separations and starburst strength \\citep{bart00}. Even minor mergers with low-mass companions have been shown through simulations to result in significant nuclear star formation activity \\citep{miho94}. Several early studies of nuclear starbursts suggested that this star formation might require a top heavy initial mass function (IMF), preferentially producing high mass stars \\citep{riek80, doyo92}. There is theoretical support for this suggestion, with simulations showing that an IMF weighted to high-mass stars naturally arises in high-density regions, due to feedback processes heating the gas. In a recent study, \\citet{krum10} have demonstrated that such regions should have a high-mass stellar fraction at least 1.7 times larger, and possibly much more, than lower density, more quiescent regions. However, the observational evidence for this variation has to date proved controversial (see \\citealt{bast10} for a recent review). Some studies have found indirect evidence for top-heavy IMFs with, for example, \\citet{riek93} concluding that the nearby starburst galaxy M82 requires an IMF biased to high mass stars to explain its emission line ratios and total luminosity. Similar techniques have been used for NGC 3256 (an ongoing merger with a `super-starburst') which have again shown indications of a modified IMF with an excess of high mass stars \\citep{doyo94}. \\citet{gibs97} showed that, in order to reproduce the observed colour-luminosity relation of elliptical galaxies, an IMF much flatter than that of \\citet{salp55} needed to be adopted. \\citet{baug05} had to employ a top heavy IMF for the starbursts powering the distant population of highly luminous submillimetre galaxies in order to explain the number counts of these systems. Finally, \\citet{bras07} studied nine interacting galaxies from the $Chandra$ survey and found that highly disturbed systems showed a strongly enhanced infrared luminosity compared to that expected from the x-ray emission, again suggesting the need for a top-heavy IMF. More direct evidence of a variation in IMF has been found for the resolved stellar population of the young Arches cluster in the Galactic Centre. \\citet{fige99,stol02,paum06,espi09} all find evidence for stellar mass functions weighted towards high-mass stars in this cluster or the general Galactic Centre region. Such mass functions are parametrized as an IMF that is either much flatter than that found by \\citet{salp55}, or having a higher mass turnover than is found in the function for field stars. One possible tracer of the IMF that has not been fully exploited to date is the relative numbers of core-collapse supernovae (CCSNe) of different types. Their short progenitor lifetimes and high luminosities make them powerful indicators of recent or ongoing star formation, and indeed they provide the only direct tracer of recent star formation within unresolved stellar populations. Recent advances in the understanding of supernovae and their progenitors raise the possibility that they can provide information on the initial mass function of a young stellar population. Theoretical models of single star progenitors predict that SNII should have lower mass progenitors than SNIb or SNIc \\citep{hege03, eldr04}. This has received observational support from studies of the strength of association with H$\\alpha$ emission \\citep{ande08}, confirming that SNII have the lowest mass CC progenitors, but additionally indicating that the SNIc have still higher mass progenitors than the SNIb. The existence of this II-Ib-Ic progenitor mass sequence allows information on the IMF of the stellar population in the SN environments to be derived from the relative numbers of type II, Ib and Ic supernovae. \\citet{petr95} studied the distribution of SNe events in 32 interacting systems containing 12 known core collapse SNe. They found that the radial distribution of these core collapse events showed a higher concentration towards the nuclear regions of the interacting galaxies when compared to isolated galaxies. This confirmed the enhanced star formation around the central regions of the systems, but the sample was too small to analyse the separate types of CCSNe. This paper will therefore use a larger sample of local CCSNe to explore the IMF in nuclear starbursts, resulting from galaxy disturbance, by studying the ratio of type II/Ibc SNe in both disturbed and undisturbed host galaxies. Throughout this paper, we use `Ibc' to encompass all SNe with classifications of Ib, Ic or Ib/c. The structure of the paper is as follows: In Section~\\ref{sec:sample} we will define and discuss the sample used throughout this work. Section~\\ref{sec:results} will describe the results on the radial distributions, for disturbed and undisturbed hosts and looking separately at type II and Ibc SNe. In Section~\\ref{sec:disc} we discuss the possible interpretations of our results, in terms of metallicity, extinction and IMF effects. Finally, Section~\\ref{sec:conc} contains a summary of our conclusions. ", "conclusions": "\\label{sec:conc} We have analysed the spatial distribution of 178 CCSNe within a sample of host galaxies with recession velocities less than 6000~km/s. Host galaxies were classified by eye according to whether they show disturbance due to strong tidal interactions or mergers. The main results are as follows: \\begin{itemize} \\item CCSNe of all types show a strong degree of central concentration in the disturbed galaxies, probably as a result of nuclear starbursts in these galaxies. \\item This central excess is dominated by SNIbc. \\item The SNIbc in disturbed galaxies are more centrally concentrated than the H$\\alpha$ emission. \\item The SNIbc excess cannot easily be explained in terms of metallicity effects, extinction, or central incompleteness of SNe. \\item Our preferred explanation of the SNIbc excess is that the central regions of the disturbed galaxies are dominated by nuclear starbursts with IMFs biased towards high mass stars, although metallicity, binarity and stellar rotation may also play a role. \\end{itemize} \\ack This paper has made use of data provided by the Central Bureau for Astronomical Telegrams. We would like to acknowledge members of staff at the Astrophysics Research Institute, in particular Sue Percival and David Bersier for their helpful comments and discussion. The authors would also like to thank the referee for a constructive and helpful report. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The Isaac Newton Telescope is operated on the island of La Palma by the Isaac Newton Group in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof\\'isica de Canarias. The Liverpool Telescope is operated on the island of La Palma by Liverpool John Moores University in the Spanish Observatorio del Roque de los Muchachos of the Instituto de Astrof\\'isica de Canarias with financial support from the UK Science and Technology Facilities Council. SMH would like to acknowledge STFC for a research studentship. \\newpage \\begin{figure} \\noindent\\makebox[\\textwidth]{% \\includegraphics{figure1.eps}} \\caption{ Histogram showing the distribution of fractions of host galaxy $R$-band light lying within the locations of each CCSN in our undisturbed host galaxies. The top plot represents the distribution of all CCSNe, the middle SNII and the lower SNIbc.} \\label{figure1} \\end{figure} \\newpage \\begin{figure} \\noindent\\makebox[\\textwidth]{% \\includegraphics{figure2.eps}} \\caption{ As Figure 1, but for the disturbed host galaxies. } \\label{figure2} \\end{figure} \\newpage \\begin{figure} \\noindent\\makebox[\\textwidth]{% \\includegraphics{figure3.eps}} \\caption{ Histogram showing the distribution of fractions of host galaxy H$\\alpha$ light lying within the locations of each CCSN in our sample, for the undisturbed host galaxies. Again, the upper plot shows the overall CCSNe distribution, the middle SNII and the lower SNIbc. } \\label{figure3} \\end{figure} \\newpage \\begin{figure} \\noindent\\makebox[\\textwidth]{% \\includegraphics{figure4.eps}} \\caption{ As Figure 3, for the disturbed host galaxies. } \\label{figure4} \\end{figure} \\newpage \\begin{figure} \\noindent\\makebox[\\textwidth]{% \\includegraphics{figure5.eps}} \\caption{ Images of 12 of the host galaxies classified as disturbed and with centrally located CCSNe. } \\label{figure5} \\end{figure} \\newpage \\appendix \\setcounter{section}{1} \\Tables \\begin{indented} \\item \\begin{longtable}{lllll} \\caption{{\\label{Table 1}}Undisturbed host galaxy sample used in this analysis. Columns represent the host galaxy, the individual SNe, the spectral classification of the SNe and the fractional $R$-band light and fractional H$\\alpha$ values for each SNe.}\\\\ \\br Host&SN&SN type&Fr($R$)&Fr(H$\\alpha$)\\\\ \\mr NGC493&1971S&IIP&0.605&0.570\\\\ NGC918&2009js&IIP&0.703&-\\\\ NGC941&2005ad&II&0.831&0.864\\\\ NGC991&1984L&Ib&0.498&0.401\\\\ NGC1035&1990E&IIP&0.272&0.363\\\\ NGC1058&1961V&II&0.968&0.931\\\\ NGC1058&1969L&IIP&1.000&1.000\\\\ NGC1058&2007gr&Ib/c&0.421&-\\\\ NGC1073&1962L&Ic&0.754&0.518\\\\ NGC1087&1995V&II&0.368&0.497\\\\ NGC1187&1982R&Ib&0.695&0.760\\\\ NGC1187&2007Y&Ib&0.981&1.000\\\\ MCG-01-09-24&2002ei&IIP&0.195&0.195\\\\ NGC1343&2008dv&Ic&0.195&0.134\\\\ UGC2906&2008im&Ib&0.682&-\\\\ UGC2971&2003ig&Ic&0.176&0.108\\\\ IC381&2001ef&Ic&0.082&0.052\\\\ NGC1832&2004gq&Ib&0.672&0.328\\\\ NGC1832&2009kr&II&0.489&-\\\\ IC2152&2004ep&II&0.461&0.560\\\\ UGC3804&2002A&IIn&0.419&0.253\\\\ NGC2551&2003hr&II&0.914&1.000\\\\ NGC2596&2003bp&Ib&0.486&0.362\\\\ UGC4436&2004ak&II&0.887&0.882\\\\ NGC2726&1995F&Ic&0.037&0.050\\\\ NGC2742&2003Z&IIP&0.675&0.736\\\\ NGC2715&1987M&Ic&0.129&0.044\\\\ UGC4904&2006jc&Ib/c&0.332&0.525\\\\ NGC2841&1972R&Ib&0.855&0.904\\\\ NGC2906&2005ip&II&0.399&0.528\\\\ UGC5249&1989C&IIP&0.017&0.058\\\\ NGC3074&1965N&IIP&0.110&0.059\\\\ NGC3074&2002cp&Ib/c&0.936&0.961\\\\ NGC3147&2006gi&Ib&0.984&0.991\\\\ NGC3184&1921B&II&0.856&0.954\\\\ NGC3184&1937F&IIP&0.808&0.930\\\\ NGC3184&1999gi&IIP&0.276&0.112\\\\ NGC3198&1966J&Ib&0.898&0.963\\\\ NGC3198&1999bw&IIn&0.745&0.755\\\\ NGC3240&2001M&Ic&0.323&0.251\\\\ NGC3294&1990H&IIP&0.156&0.125\\\\ NGC3340&2005O&Ib&0.322&0.305\\\\ \\newpage \\br Host&SN&SN type&Fr($R$)&Fr(H$\\alpha$)\\\\ \\mr NGC3340&2007fp&II&0.170&0.125\\\\ NGC3430&2004ez&II&0.788&0.833\\\\ NGC3437&2004bm&Ic&0.073&0.076\\\\ NGC3451&1997dn&II&0.872&0.946\\\\ NGC3504&2001ac&IIn&0.826&0.992\\\\ NGC3512&2001fv&IIP&0.669&0.689\\\\ NGC3556&1969B&IIP&0.197&0.494\\\\ NGC3631&1964A&II&0.915&0.992\\\\ NGC3631&1965L&IIP&0.622&0.658\\\\ NGC3631&1996bu&IIn&0.923&0.993\\\\ NGC3655&2002ji&Ib/c&0.709&0.957\\\\ NGC3683&2004C&Ic&0.532&0.545\\\\ UGC6517&2006lv&Ib/c&0.480&-\\\\ NGC3756&1975T&IIP&0.846&0.856\\\\ NGC3810&2000ew&Ic&0.261&0.147\\\\ NGC3810&1997dq&Ib/c&0.774&0.734\\\\ NGC3949&2000db&II&0.364&0.253\\\\ NGC3963&1997ei&Ic&0.197&0.053\\\\ NGC4030&2007aa&II&0.942&0.828\\\\ NGC4041&1994W&IIn&0.491&0.541\\\\ NGC4051&1983I&Ic&0.498&0.473\\\\ NGC4051&2003ie&II&0.838&0.885\\\\ IC758&1999bg&IIP&0.669&0.657\\\\ NGC4136&1941C&II&0.880&0.882\\\\ NGC4210&2002ho&Ic&0.146&0.051\\\\ NGC4242&2002bu&IIn&0.896&0.930\\\\ NGC4303&1926A&IIL&0.607&0.736\\\\ NGC4303&1961I&II&0.697&0.877\\\\ NGC4303&1964F&II&0.189&0.106\\\\ NGC4303&1999gn&IIP&0.418&0.429\\\\ NGC4303&2006ov&IIP&0.418&0.429\\\\ NGC4303&2008in&IIP&0.845&-\\\\ NGC4369&2005kl&Ic&0.271&0.540\\\\ NGC4384&2000de&Ib&0.087&0.140\\\\ NGC4451&1985G&IIP&0.138&0.212\\\\ NGC4559&1941A&IIL&0.208&0.131\\\\ UGC7848&2006bv&IIn&0.579&-\\\\ NGC4666&1965H&IIP&0.324&0.198\\\\ NGC4708&2003ef&II&0.335&0.352\\\\ NGC4725&1940B&IIP&0.675&0.802\\\\ NGC4900&1999br&IIP&0.786&0.932\\\\ NGC4961&2005az&Ic&0.426&-\\\\ NGC4981&2007C&Ib&0.320&0.236\\\\ NGC5012&1997eg&IIn&0.503&0.449\\\\ NGC5033&1950C&Ib/c&0.972&1.000\\\\ NGC5033&1985L&IIL&0.585&0.571\\\\ NGC5334&2003gm&IIn&0.480&-\\\\ NGC5371&1994Y&IIn&0.355&0.212\\\\ NGC5468&2002ed&IIP&0.811&0.791\\\\ \\newpage \\br Host&SN&SN type&Fr($R$)&Fr(H$\\alpha$)\\\\ \\mr NGC5559&2001co&Ib/c&0.618&0.497\\\\ NGC5584&1996aq&Ic&0.178&0.086\\\\ NGC5630&2005dp&II&0.534&0.590\\\\ NGC5630&2006am&IIn&0.604&0.617\\\\ NGC5673&1996cc&II&0.924&0.934\\\\ NGC5668&2004G&II&0.657&0.595\\\\ NGC5775&1996ae&IIn&0.757&0.671\\\\ NGC5806&2004dg&IIP&0.484&0.378\\\\ NGC5850&1987B&IIn&0.995&-\\\\ NGC5879&1954C&II&0.615&0.511\\\\ NGC5921&2001X&IIP&0.579&0.369\\\\ NGC6118&2004dk&Ib&0.673&0.626\\\\ NGC6207&2004A&IIP&0.729&0.660\\\\ UGC10862&2004ao&Ib&-&0.215\\\\ NGC6643&2008ij&IIP&0.519&0.620\\\\ NGC6643&2008bo&Ib&0.451&0.510\\\\ NGC6700&2002cw&Ib&-&0.642\\\\ NGC6946&2004et&II&0.975&-\\\\ NGC6951&1999el&IIn&0.320&0.259\\\\ UGC11861&1995ag&II&0.343&0.170\\\\ UGC11861&1997db&II&0.633&0.396\\\\ UGC12160&1995X&II&0.564&0.484\\\\ UGC12182&2006fp&IIn&1.000&1.000\\\\ \\br \\end{longtable} \\end{indented} \\newpage \\begin{indented} \\item \\begin{longtable}{lllll} \\caption{\\label{Table 2}Disturbed host galaxy sample used in this analysis. Columns represent the host galaxy, the individual SNe, the spectral classification of the SNe and the fractional $R$-band light and fractional H$\\alpha$ values for each SNe, as in table 1.}\\\\ \\br Host&SN&SN type&Fr($R$)&Fr(H$\\alpha$)\\\\ \\mr NGC895&2003id&Ic&0.524&-\\\\ UGC2984&2002jz&Ic&0.091&0.099\\\\ NGC1614&1996D&Ic&0.275&-\\\\ NGC1637&1999em&IIP&0.276&0.268\\\\ IC391&2001B&Ib&0.062&0.060\\\\ NGC1961&2001is&Ib&-&0.749\\\\ NGC2207&1999ec&Ib&-&0.521\\\\ NGC2207&2003H&Ib&-&0.259\\\\ NGC2146&2005V&Ib/c&0.033&0.091\\\\ ESO492-G2&2005lr&Ic&-&0.005\\\\ UGC3829&2001ej&Ib&0.152&0.391\\\\ NGC2276&1968V&II&0.699&0.790\\\\ NGC2276&2005dl&II&0.247&0.099\\\\ NGC2276&1993X&II&0.899&0.619\\\\ NGC2532&1999gb&IIn&0.485&0.443\\\\ NGC2532&2002hn&Ic&0.023&0.011\\\\ NGC2604&2002ce&II&0.381&0.560\\\\ NGC2782&1994ak&IIn&0.725&0.977\\\\ NGC2993&2003ao&IIP&0.456&0.784\\\\ NGC3169&1984E&IIL&0.684&0.731\\\\ NGC3310&1991N&Ic&0.268&0.277\\\\ NGC3323&2004bs&Ib&0.191&0.119\\\\ NGC3323&2005kk&II&0.766&0.875\\\\ NGC3367&1992C&II&0.689&0.687\\\\ NGC3367&2007am&II&0.302&0.314\\\\ NGC3627&1973R&IIP&0.471&0.566\\\\ NGC3627&1997bs&IIn&0.362&0.348\\\\ NGC3627&2009hd&II&0.496&-\\\\ NGC3690&1993G&IIL&0.464&0.744\\\\ NGC3690&1998T&Ib&0.056&0.056\\\\ NGC3690&1999D&II&0.560&0.849\\\\ NGC3786&1999bu&Ic&0.180&0.522\\\\ NGC3811&1971K&IIP&0.809&0.900\\\\ IC2973&1991A&Ic&0.742&0.588\\\\ NGC4038&2004gt&Ib/c&0.834&0.991\\\\ NGC4088&1991G&IIP&0.466&0.453\\\\ NGC4088&2009dd&II&0.100&-\\\\ NGC4141&2008X&IIP&0.194&0.085\\\\ NGC4141&2009E&IIP&0.594&0.491\\\\ NGC4254&1967H&II&0.664&0.648\\\\ NGC4254&1972Q&IIP&0.811&0.791\\\\ NGC4254&1986I&IIP&0.334&0.318\\\\ NGC4273&2008N&IIP&0.300&0.316\\\\ NGC4273&1936A&IIP&0.569&0.598\\\\ NGC4490&1982F&IIP&0.277&0.202\\\\ \\newpage \\br Host&SN&SN type&Fr($R$)&Fr(H$\\alpha$)\\\\ \\mr NGC4568&1990B&Ic&0.302&-\\\\ NGC4568&2004cc&Ic&0.158&-\\\\ NGC4618&1985F&Ib&0.121&0.087\\\\ NGC4615&1987F&IIn&0.489&0.333\\\\ NGC4688&1966B&IIL&0.571&0.454\\\\ NGC4691&1997X&Ic&0.171&0.472\\\\ NGC5000&2003el&Ic&0.482&0.476\\\\ NGC5021&1996ak&II&0.619&0.659\\\\ MCG-04-32-07&2003am&II&0.211&0.339\\\\ NGC5194&1994I&Ic&-&0.122\\\\ NGC5194&2005cs&IIP&-&0.222\\\\ NGC5395&2000cr&Ic&0.538&0.549\\\\ NGC5480&1988L&Ib&0.230&0.369\\\\ NGC5682&2005ci&II&0.204&0.191\\\\ NGC7479&1990U&Ic&0.603&0.488\\\\ NGC7479&2009jf&Ib&0.764&-\\\\ NGC7537&2002gd&II&0.759&0.685\\\\ NGC7714&2007fo&Ib&0.377&-\\\\ UGC12846&2007od&IIP&0.945&1.000\\\\ \\br \\end{longtable} \\end{indented} \\clearpage \\def\\newblock{\\hskip .11em plus .33em minus .07em}" }, "1005/1005.3332_arXiv.txt": { "abstract": "s{We present an analytical derivation of the Sachs Wolfe effect sourced by a primordial magnetic field, generated by a causal process, such as a first order phase transition in the early universe. As for the topological defects case, we apply the general relativistic junction conditions to match the perturbation variables before and after the phase transition, in such a way that the total energy momentum tensor is conserved across the transition. We find that the relevant contribution to the magnetic Sachs Wolfe effect comes from the metric perturbations at next-to-leading order in the large scale limit. The leading order term is strongly suppressed due to the presence of free-streaming neutrinos. We derive the neutrino compensation effect and confirm that the magnetic Sachs Wolfe spectrum from a causal magnetic field behaves as $\\ell(\\ell+1)\\,C^\\B_\\ell \\propto \\ell^2$ as found in the latest numerical analyses.} ", "introduction": "The origin of the large scale magnetic fields observed in galaxies and clusters is still unknown: one of the possible explanations is that they have been generated in the primordial universe. A primordial magnetic field of the order of the nanoGauss could leave a detectable imprint in the cosmic microwave background (CMB) anisotropies~\\cite{Kahniashvili:2006hy,Yamazaki:2008gr,PFP,Shaw:2009nf}. Here we concentrate on its effect on the temperature CMB spectrum at large scales and more particularly on the Sachs Wolfe effect. The motivation is that conflicting results exist in the literature regarding the $\\ell$-dependence of the Sachs Wolfe effect induced by a causal primordial magnetic field: the analytical analysis of \\cite{Kahniashvili:2006hy} found $\\ell(\\ell+1)\\,C^\\B_\\ell$ scaling as $ \\ell^{-1}$ or more negative, and the same result was found in the numerical calculation of \\cite{Yamazaki:2008gr}; on the other hand, \\cite{PFP,Shaw:2009nf,Caprini:2009vk} found $\\ell(\\ell+1)\\,C^\\B_\\ell$ scaling as $\\ell^2$. The aim of this paper is to explain this discrepancy analytically (for a more detailed derivation see~\\cite{cmbmag}). We assume that a magnetic field is generated in the early universe by a sudden phase transition, as for example the electroweak (EW) phase transition. We consider a stochastic magnetic field with no background component and we suppose that the magnetic energy momentum tensor is first order in perturbation theory. We study the effect of the magnetic field on the metric and fluid (matter plus radiation) perturbations by solving analytically Einstein's and conservation equations in the long wavelength limit. We take into account the neutrinos in our derivation. In order to connect the solutions before and after the magnetic field generation, we match the geometry and the fluid variables at the phase transition time, so that the induced three metric and the extrinsic curvature are continuous~\\cite{Deruelle:1997py}. This implies the conservation of the total energy momentum tensor across the phase transition, and it completely determines the metric and fluid perturbation variables after the magnetic field generation. Before neutrino decoupling, we find that at leading order in the large scale expansion $k/\\mathcal{H}\\ll 1$, the metric perturbation $\\Phi$ is proportional to $\\Phi \\propto \\Pi_\\B (\\HH/k)^2$, where $\\Pi_\\B$ is the magnetic field anisotropic stress. This induces a contribution in the CMB spectrum scaling as $\\ell(\\ell+1)\\,C^\\B_\\ell\\propto \\ell^{-1}$, and consistent with~\\cite{Kahniashvili:2006hy,Yamazaki:2008gr}. However, once neutrinos decouple and start free-streaming, they acquire a non-zero anisotropic stress, which acts to compensate and reduce the magnetic field one~\\cite{Shaw:2009nf,Kojima:2009gw}. We demonstrate that this compensation drastically reduces the leading order contribution to the CMB spectrum, and that the dominant contribution becomes the one from the next-to-leading order in the $k/\\mathcal{H}\\ll 1$ expansion, which induces then $\\ell(\\ell+1)\\,C^\\B_\\ell \\propto \\ell^2$, as found in~\\cite{PFP,Shaw:2009nf,Caprini:2009vk}. ", "conclusions": "In this work we present an analytical computation of the Sachs Wolfe effect induced by a primordial magnetic field. We have restricted our analysis to a magnetic field generated by a causal process, such as a first order phase transition. In this case, the initial conditions for the metric and fluid variables are determined unambiguously by imposing conservation of the total energy momentum tensor across the transition. Using these initial conditions, we have computed analytically the leading order and next-to-leading order magnetic contribution to the Sachs Wolfe effect. We have found that the leading order contribution is sourced only by the magnetic field anisotropic stress, and leads to a CMB spectrum scaling as $1/\\ell$. However, this contribution is strongly suppressed once the magnetic field anisotropic stress is compensated by the one of the neutrinos. As a consequence, the dominant contribution to the Sachs Wolfe is the next-to-leading order one, that generates a CMB spectrum scaling as $\\ell^2$. Our analytical work solves therefore the discrepancy regarding the $\\ell$-dependence of the magnetic Sachs Wolfe." }, "1005/1005.1876_arXiv.txt": { "abstract": "Long-duration gamma-ray bursts(LGRBs) are believed to be linked with the star formation. We adopt a galactic evolution model, in which the star formation process inside the virialized dark halo at given redshift can be achieved. In this paper, the gamma-ray burst(GRB) host galaxies are assumed to be the star-forming galaxies within the small dark halos. The star formation rates(SFRs) in the host galaxies of LGRBs at different redshifts have been derived from our model with the galactic evolutionary time about a few times of $10^7$ yr and the dark halo mass of about $5\\times 10^{11}M_\\odot$. The related stellar masses, luminosities and metallicities of these hosts are estimated as well. We further calculate the X-ray and optical absorption of GRB afterglow emission. From our model calculation, at higher redshift, the SFR of host galaxy is larger, the absorption in X-ray band and optical band of GRB afterglow is stronger, in the condition that the dust and metal components are released locally, surrounding the GRB environment. These model predictions are compared with the {\\it Swift} and other observational data. At lower redshift $z<1$, as the merger and interaction of some host galaxies are involved, one monolithic physical process is not sufficient to fully explain all kinds of observed phenomena. ", "introduction": "Gamma-ray burst(GRB), the most violent explosion cosmic source, has been identified as the cosmological event since 1997(van Paradijs et al. 1997; Metzger et al. 1997). Recently GRB 090423 has been explored at high redshift above 8(Salvaterra et al. 2009a; Tanvir et al. 2009). The long-duration GRB(LGRB) progenitors are proposed to be the massive collapsing stars (Woosley 1993; Kumar, Narayan \\& Johnson 2008). Some long bursts have been observed to be associated with supernova events(Hjorth et al. 2003; Stanek et al. 2003; Malesani et al. 2004; Mazzali et al. 2006; Xu et al. 2008), hence having a common star-forming origin(Paczynski 1998). Indeed, long GRBs can be found in the star formation galaxies and these galaxies are dominated by the young stellar population(Christensen et al. 2004). In general, GRBs favor a metal-poor environment(Fynbo et al. 2006; Kewley et al. 2007) and the hosts have low stellar masses(Wiersema et al. 2007). Jakobsson et al. (2005) proposed that GRB host galaxies, at least those high redshift($z>2$) hosts, trace the star formation of the universe in an unbiased way. The high global star formation rate(SFR) history at redshift larger than 6 (Hopkins \\& Beacom 2006; Yan et al. 2009) indicates the possibility of high-redshift GRB production and the detection of host galaxies. From the research of Y\\\"{u}ksel et al. (2008) and Kistler et al. (2009), there could be a link from star formation to the GRB production in the high redshift universe, in which the GRB luminosity function is involved. Moreover, the evolution of the GRB luminosity function has been investigated by Salvaterra et al. (2009b). All of these evidence provide the strong clue to study the intrinsic link from SFR to GRB production and the possible evolutionary properties of GRBs and their hosts. The grains and metals produced by the host galaxy will take effects on the GRB afterglow emissions. Thus, the GRB progenitors and their environments can be expressed by the absorption features of GRB afterglows. The heavy attenuation in the X-ray band has been given in the statistic results from Campana et al. (2010), indicating a dense surrounding environment of those GRBs. In the mean while, it is also interesting to understand whether this kind of strong attenuation is intrinsically evolved with redshift. On the other hand, the characteristics of the corresponding absorption in the optical band are still under debate. Although the approximate dust extinction law of GRB host galaxies has been given by Chen, Li \\& Wei (2006) and Li et al. (2008), in order to have an explanation of dust obscuration and especially to interpret some X-ray detected but optical faint bursts(so-called dark bursts, Akerlof \\& Swan 2007; Kann et al. 2007; Perley et al. 2009), the physical origin associated with the star formation and galactic evolution should be studied in an unified scenario. In this paper, we specify one physical model of star-forming and metal-poor galaxies being as the hosts of long GRBs, exploiting the physical recipes from Granato et al. (2004). In the general scenario of Granato et al. (2004), at each redshift bin, the SFR and galaxy mass in the given dark halo potential well have been calculated, with the effects on the kinetic feedback of supernova and central black hole. Under this framework, the different evolutionary stages of galaxies and the central black holes with different physical conditions have been investigated(e.g., Cirasuolo et al. 2005 about the properties of E/S0 galaxies; Lapi et al. 2006 about the active galactic nucleus luminosity function; Granato et al. 2006 about the submillimeter galaxies). In particular, Mao et al. (2007) calculated the UV luminosities and the relative dust attenuation in the star-forming and metal-poor galaxies, Lapi et al (2008) estimated the long GRB progenitor rates and redshift distribution. Since the updated X-ray/optical observations on the GRB afterglows and host galaxies have been performed sequentially by Castro Cer\\'{o}n et al. (2008), Evans et al. (2009), Savaglio, Glazebrook \\& Le Borgne (2009), Levesque et al. (2009a) and Fynbo et al. (2009), in this context, it is necessary to further compare some properties calculated by our model with these updated observational data. We extend the former calculation from Mao et al. (2007), attempting to understand the physical origin of the long GRB production and the GRB environment, especially, we reveal that some properties from afterglow emissions and GRB hosts have shown the possible intrinsic cosmological evolution. Throughout the paper, we adopt cosmological parameters: $h=0.7$, $\\Omega_M=0.3$, and $\\Omega_\\Lambda =0.7$. ", "conclusions": "Under the framework of galaxy formation scenario, Lapi et al. (2008) predicted the GRB progenitor rate and redshift distribution. In this paper, without the information of GRB rates and the cosmological star formation density, we attempt to reveal some properties of GRBs and their host galaxies, which have intrinsic redshift distributions. The distributions of these properties with redshift are found to be originally from the star formation in the star-forming galaxies. Given a proper galactic evolutionary time and a reasonable dark halo mass, the final results can be obtained by the model calculation. These results are compared with all kinds of observational data. At high redshift, the GRB host galaxy has a plenty of neutral gas, suffering much violent star formation. After the short-time stellar evolution phase, the metal and dust are released by massive stars; thus, the optical and X-ray GRB emissions will have a strong attenuation locally at high redshift. The star formation activity, evolving from relatively massive hosts at high redshift to dwarf galaxies at low redshift, is a coincidence with the so-called downsizing scenario(e.g., Heavens et al. 2004). However, at lower redshift, the situation turns to be more complicated. From the morphological statistics by Conselice et al. (2005) and Wainwright et al. (2007), the GRB hosts present a broad diversity of galaxy types. About 1/3 host galaxies in the sample of Savaglio et al. (2009) are mergers, while in our model the merging and interaction processes are not taken into account. In fact, Conselice et al. (2005) found that the GRB hosts at $z>1$ are different from those at $z<1$ in terms of light concentration and the morphological size. Through the study of galaxy mass distribution, GRB hosts tracing star formation might be biased at low redshift(Kocevski et al. 2009). It is also complex that the hosts at $z<1$ are not representative of the general galaxy population(Levesque et al. 2009a). Thus, the properties of these low-redshift GRB hosts presented in this paper could not be reproduced by any monolithic process. At least, some low-redshift galaxies may undergo multiple star-forming processes during their whole lifetimes. GRB production can be accompanied with any single starburst event. From the analysis in this paper, we see that the absorption of GRB X-ray and optical emissions is relatively strong. The strong intrinsic attenuation of GRB host galaxies may produce some dark bursts, defined by the index $\\beta_{ox}<0.5$, where $\\beta_{ox}$ is the flux density ratio between optical and X-ray bands(Jakobsson et al. 2004). Rol et al. (2005) proposed several extinction origins from their preliminary results. From our calculations, we see that the heavy attenuation may occur due to the following three possibilities: (1) the local environment of the host is metal-enriched, metallicity is higher, and/or, the host galaxy in the massive dark halo larger than $10^{12}M_\\odot$ may have strong absorption. For example, at redshift 2.5, $Z=1.0Z_\\odot$, halo mass $M_{halo}=5.0\\times 10^{12}M_{\\odot}$, after the galactic evolving time $1.0\\times 10^8~yr$, we have dust extinction $A_v=1.0$ and the corresponding X-ray absorption $N_{H,x}=4.7\\times 10^{22}cm^{-2}$; (2) the dust and metals surrounding the GRB in the host galaxy are distributed in an inhomogeneous way; there could be heavy absorption through the line of sight, but in other directions the absorption is slight. Also, in our model, we assume that the $A_v$ and $N_{H,x}$ are measured locally and do not change significantly if the dust and gas extend out to a few tens to hundreds of pc from the burst(Perna \\& Lazzati 2002, D'Elia et al. 2009); however, suppose the observed optical extinction is due to the grain absorption far beyond this local region of GRB, the $A_v$-$N_{H,x}$ correlation obtained by Schady et al. (2007, 2010) may be invalid and our calculations are strongly biased; (3) as mentioned in Section 2.2.2, the dust produced by the AGB population at high redshift should be taken into account. In order to further understand the metal production of GRB environment, we roughly re-estimate the metallicity of GRB hosts under our framework. The mass of metal $M_{metal}=SFR\\cdot f\\cdot f_{dep}\\cdot M_{dust}/M_{star}$, where f is the ratio of massive stars to all stars, and $f_{dep}$ is the ratio of mental converted from the dust. $f=0.47$ is the case for the stars with the mass larger than $2M_\\odot$ by our adopted IMF, $f_{dep}=1.0$ means that all the dust can be transferred to metals. Metallicity is defined by $Z=M_{metal}/M_{gas}$. From the SFR calculated by Granato et al. (2004) and Mao et al. (2007), as an example, at redshift 6, we obtain the metallicity as $Z\\sim 2.75\\times 10^{-2}(M_{dust}/M_{star})$. If we take a supernova with the dust production of $10^{-3}$ solar mass(Pozzo et al. 2004), we obtain the upper limit of metallicity $Z\\sim 10^{-3}Z_{\\odot}$, which is lower than the measurement($Z>0.02Z_\\odot$) of GRB 050904(Campana et al. 2007). If we take the dust mass 0.08-0.3 solar mass per primordial massive supernova(Todini \\& Ferrara 2001), we have the result which is consistent with the observation. The estimation values of Population I/II metallicity are lower than the observational values at high redshift, meaning that the imprints from some primordial objects(Kawai et al. 2006), such as pop III stars and mini-quasars, have to be included in the possible cosmic evolution properties of these GRB host galaxies. According to this estimation, the metal-enriched environment of GRB host galaxy naturally gives the reason of the strong attenuation in X-ray and optical band measurements. In our model, the initial galactic evolutionary time of about $10^7$ yr of host galaxies is given, but the corresponding metallicity about $Z\\sim 0.3Z_\\odot$(Lapi et al. 2008) is not a necessary condition, as mentioned by Levesque et al. (2009b) that low metallicity may not be required for a relativistic explosion. With our model, the massive dark halo above $10^{12}M_\\odot$ can host the GRB galaxy in which the metallicity is relatively high, although most GRB host galaxies are inside the dark halos with the masses less than $10^{12}M_\\odot$. On the other hand, a host galaxy with a top-heavy IMF, meaning that much more massive stars are involved, can produce more metals in relatively short time during the galactic evolution phase. For instance, the Wolf-Rayet star with a mass of 80$M_\\odot$ and initial metallicity $Z=0.001$ has the possibility to self-enrich the HII region(Kr\\\"{o}ger, Hensler \\& Freyer 2006) and to produce the GRB event(Eldridge et al. 2006). In this paper, we have calculated the SFR, galactic mass, and metallicity of the GRB host galaxies. The absorption variations with redshift in the X-ray and optical bands are presented as well. Some selection effects have been taken into account through our calculation. Other observational biases should also be considered. All the redshift measurements come from the optical observations so that some optical-faint GRBs and host galaxies are ignored. Moreover, at high redshift only most luminous galaxies with high SFRs can be detected, indicating that some low-luminosity cases are not included. However, our calculations come from the intrinsic star formation of GRB host galaxies. Thus, due to all these selection effects and observational bias mentioned in the paper, the intrinsic properties of GRB afterglows and the hosts by the model calculations have some differences to those from observations. As SFR evolution plays a dominant role in the calculations, compared to the situation at low redshift, in general, star formation in the metal-poor environment at high redshift may provide more powerful GRB explosion. Therefore, although the effective threshold is given by Kistler et al. (2009), we speculate that improving the sensitivity of detectors on the high-energy telescopes is not strongly useful to catch more high-redshift but faint GRBs, since low-energy-released GRBs are almost absent in the high-redshift universe." }, "1005/1005.3042_arXiv.txt": { "abstract": "{In recent years mid- and far infrared spectra of planetary nebulae have been analysed and lead to more accurate abundances. It may be expected that these better abundances lead to a better understanding of the evolution of these objects.} {The observed abundances in planetary nebulae are compared to those predicted by the models of Karakas (2003) in order to predict the progenitor masses of the various PNe used. The morphology of the PNe is included in the comparison. Since the central stars play an important role in the evolution, it is expected that this comparison will yield additional information about them.} {First the nitrogen/oxygen ratio is discussed with relation to the helium/hydrogen ratio. The progenitor mass for each PNe can be found by a comparison with the models of Karakas. Then the present luminosity of the central stars is determined in two ways: first by computing the central star effective temperature and radius, and second by computing the nebular luminosity from the hydrogen and helium lines. This luminosity is also a function of the initial mass so that these two values of initial mass can be compared.} {Six of the seven bipolar nebulae can be identified as descendants of high mass stars (4M\\smallsun - 6M\\smallsun) while the seventh is ambiguous. Most of the elliptical PNe have central stars which descend from low initial mass stars, although there are a few caveats which are discussed. There is no observational evidence for a higher mass for central stars which have a high carbon/oxygen ratio. The evidence provided by the abundance comparison with the models of Karakas is consistent with the HR diagram to which it is compared. In the course of this discussion it is shown how `optically thin' nebulae can be separated from those which are 'optically thick'.} {} ", "introduction": "Planetary nebulae (hereafter PNe) are an advanced stage of stellar evolution of low and intermediate mass stars. After the asymptotic giant branch (AGB) phase is completed, these stars evolve through the PN stage before ending their lives as white dwarfs. The gaseous nebula seen now as PN is the remnant of the deep convective envelope which once surrounded the core. This core is now seen as the central star of the PN. The present abundances in the nebula reveal information about the chemical processes that took place during the AGB. These processes, which have first been discussed by Iben \\& Renzini \\cite{iben} and Renzini \\& Voli \\cite{renzini}, change the abundances according to the mass of the star involved and the initial abundances in the star. Thus by investigating the PN abundances it may be possible to assign an initial mass to the star. Models have been made of the evolution of stars of different masses. These were initiated with the discussion of Paczynski \\cite{pacz} followed by the detailed calculations of Sch\\\"{o}nberner \\cite{schon}, Vassiliadis and Wood \\cite{wood} and Bl\\\"{o}cker \\cite{blocker}. These models refer mostly to post AGB evolution. Models referring to evolution on the AGB have been made by several authors, e.g. Marigo et al. \\cite{marigo} and Karakas \\cite{karakas}. The latter models predict changes in the chemical composition which have occurred during the evolution and which have been brought to the surface and subsequently expelled as the nebula. It is these models which will be used to compare with observed PN abundances because not only do they follow a star of a given mass over its entire life, but the same is done for an entire sequence of possible masses for several different initial abundances. The purpose of the present paper is to compare these models with the abundances which we have observed. These abundances have been determined with the help of mid and far infrared observations either from {\\em ISO} or {\\em Spitzer} and are quite accurate because they are less affected by possible temperature variations or gradients in the nebula. These observations have already been used (Pottasch \\& Bernard-Salas\\,\\cite{pobs} to better determine PN abundance gradients in the galaxy. In an ideal case it might be expected that a comparison of models with observations will lead to: 1) knowledge of the individual properties of the central stars, and 2) confirmation or suggestion for improvement of the models. In practice these goals are rather difficult to reach because of shortcomings of both the observations as well as the models. On the observational side are uncertainties in the effective temperature of the central star, their distances, as well as the accuracy of the measurements. The models presently available are uncertain because the physical conditions in the actual star-nebula system is poorly known. For example, the mass loss along the AGB (and post AGB) is physically not well understood and the initial conditions may not be realistic. Thus models used for comparison are taken from different authors who may use different mass loss rates. Therefore core masses are used where possible although initial masses are given for the Karakas models because the author identifies them as such. The abundances observed are listed in Table 1. No indication is given there of the spectrum of the central star. A few of these stars are Wolf-Rayet stars for which it may be that some of the evolutionary calculations may not apply. These are the central stars of BD+30 3639, NGC\\,40, NGC\\,5315 and NGC\\,6369. This can be kept in mind when making the comparisons. The objects were selected to be IR bright (in the diaphragm of the instrument used). This was first done with the ISO spectrometer where almost all of the usable PN spectra were investigated. Later the Spitzer IR spectra of PNe have been investigated. Most of these spectra are as bright or nearly as bright as the ISO PNe. This may at first suggest a bias toward PNe with massive central stars because these initially evolve at the highest luminosity. But the period of high luminosity is expected to be very short so that very few, if any, high mass central star PNe are expected. We therefore may expect that many low mass central star PNe have been observed, not only because of much longer evolution time but also because of the much greater number of low mass objects present. It is expected that most of the observed PNe are reasonably local objects, within a few kpc of the sun. Nevertheless a confrontation of the models with the observations, even with these limitations, can give interesting insights into the evolution of the PN system. In Sect.\\,2 the morphology of the nebulae will be discussed, first in relation to the nitrogen/oxygen ratio observed in the nebula, and then the helium abundance will be introduced into the discussion. In Sect.\\,3 the effective temperature of the central star and the various ways of determining it will be discussed. Then the luminosity of the central stars will be discussed. Because the luminosity is dependent on the distance of the nebula this will also be discussed in this section. In addition a digression will be made into the long-standing question of whether or not a nebula is 'thick' to photons which ionize hydrogen. This can be done because two different methods of obtaining the nebular luminosity are available, one which makes no assumption concerning the nebular 'thickness' and one which is dependent on this assumption. In Sect.\\,4 we investigate whether the expected relation between abundance and luminosity can be seen. Conclusions and discussion are given in the last section. ", "conclusions": "We have determined the masses of a selection of 33 well-studied nebulae PNe by comparing the observed nebular abundances with that predicted by the evolutionary models of Karakas\\,\\cite{karakas}, which is the first systematic study of the evolution of lower mass stars. A secondary purpose is to see whether the masses determined in this way are consistent with the evolutionary tracks computed by Sch\\\"{o}nberner\\,\\cite{schon}, Bl\\\"{o}cker\\,\\cite{blocker} and Vassiliadis \\& Wood\\,\\cite{wood}. The abundances used in this comparison are helium, oxygen, nitrogen and carbon. The morphology is also considered in this comparison but in a simplified form. The PNe are divided into two categories: bipolar and elliptical (including round nebulae). The result of the comparison of He vs N/O (section 2.2) indicate four PNe have core masses greater than 0.9M\\smallsun; three of these nebulae are bipolar (NGC6302, NGC6537 and He2-111). There are also four nebulae which appear to have a slightly lower initial mass, between 0.85M\\smallsun~and 0.9M\\smallsun~and have also undergone hot-bottom burning. Again three of these PNe are bipolar (NGC2440, Hb5 and Hu1-2). Thus 6 of the 7 bipolar nebular in our selection have abundances which indicate that they originated from initially high mass stars. The only bipolar nebula which appears to be an exception is NGC6445 which has a high He/H ratio but a low N/O ratio which is difficult to understand on the basis of the models of Karakas. NGC5315 appears to be an elliptical PN with a high mass. There are three other elliptical nebulae which on the basis of a rather high N/O ratio seem to originate from stars of rather high mass. These are NGC6153, NGC6886 and NGC2392. A unlikely explanation is that these PNe are really bipolar seen edge on making them look round. Two of these PNe do not have a high He/H ratio however, making this interpretation appear too simplified. All the other elliptical PNe have N/O and He/H ratios which do not substantially differ from solar and therefore have a core mass of less than approximately 0.7M\\smallsun, using the evolutionary models of Karakas\\,\\cite{karakas}. We cannot determine the mass more precisely for these nebulae, except to say that in the models of Karakas those PNe with a high carbon abundance (C/O$\\geq$1) have the higher mass. An HR diagram for these nebulae was then constructed in order to see if the position of the PNe on this diagram can confirm the conclusions drawn from the comparison of the abundances with the predictions of Karakas. The effective temperature of the central stars is determined, usually with an error that is less than 10\\%. The determination of the luminosity can be made in two ways. First by computing it from the temperature and radius of the central star, and secondly by finding it from the nebular emission in both the hydrogen and helium lines. The second method has the additional assumption that the nebula is 'optically deep' to ionizing radiation of hydrogen or helium. We find that the luminosity determined from the ionized helium line is usually approximately equal to the luminosity found from the central star, indicating that the nebula absorbs all radiation which can doubly ionize helium. For about 30\\% of the nebulae the hydrogen ionizing radiation gives a lower luminosity indicating that in these cases the nebula is 'optically thin' to this radiation. In this way a method has been found to determine which nebulae are `optically deep' to the various radiation fields. Probably the largest uncertainty in the luminosity is the distance determination so that a subsection is devoted to a discussion of the distances used. The resultant HR diagram is shown in Fig.4. The high masses found from the abundances for the six bipolar PNe are consistent with their position on the HR diagram. The position of the seventh bipolar PN, NGC6445, is consistent with it being a high mass object, leaving its rather low N/O ratio as a problem. The position of NGC6886 is consistent with a rather high mass. Most of the elliptical PNe have positions consistent with low core masses. The biggest problems are: 1) NGC7027, whose HR position indicates a high mass while its abundances give a low mass; 2) NGC5315 whose position indicates a low mass while its abundance gives a high mass. M1-42 is also a problem but its distance is very poorly known and some of the observations are not very good. NGC2392 has a different problem since the luminosity determined from the star is improbably high. The effective temperature that we have determined for this central star does not apply to the star measured in the visual. The most likely solution is that this central star is a binary and the secondary, which is unseen in the visual, is of a much higher temperature and is responsible for the high degree of ionization found. The possible presence of a hotter star has been suspected earlier; Ciardullo et al.\\cite{ciar} have looked for it but have found only a very faint nearby star. If this is the source of ionization it cannot be a main sequence star because if this was so it would be too distant to be associated with the nebula. But the ionization source could be a star which is too close to the bright star to be observed. In general it appears that the initial masses as determined from the observed abundances in conjunction with the models of Karakas\\,\\cite{karakas} are consistent with the initial masses predicted using the evolutionary models of Sch\\\"{o}nberner\\,\\cite{schon}, Bl\\\"{o}cker\\,\\cite{blocker} and Vassiliadis \\& Wood\\,\\cite{wood}. But being consistent is only a first step and both the models and the observed abundances should both be improved. Furthermore the various cases which appear to give inconsistent results must be understood before we can speak of agreement." }, "1005/1005.3797_arXiv.txt": { "abstract": "Direct detection of dark matter (DM) requires an interaction of dark matter particles with nucleons. The same interaction can lead to dark matter pair production at a hadron collider, and with the addition of initial state radiation this may lead to mono-jet signals. Mono-jet searches at the Tevatron can thus place limits on DM direct detection rates. We study these bounds both in the case where there is a contact interaction between DM and the standard model and where there is a mediator kinematically accessible at the Tevatron. We find that in many cases the Tevatron provides the current best limit, particularly for light dark matter, below $\\sim 5$ GeV, a and for spin dependent interactions. Non-standard dark matter candidates are also constrained. The introduction of a light mediator significantly weakens the collider bound. A direct detection discovery that is in apparent conflict with mono-jet limits will thus point to a new light state coupling the standard model to the dark sector. Mono-jet searches with more luminosity and including the spectrum shape in the analysis can improve the constraints on DM-nucleon scattering cross section. ", "introduction": "\\label{sec:intro} From astronomical and cosmological observations it is now clear that $\\sim 25\\%$ of the matter-energy content of the universe if made up by dark matter (DM). Although DM has so far only been observed through its gravitational interactions the quest for a more direct observation of DM is taking place simultaneously on many fronts. Indirect searches look for signals of standard model (SM) particle production from DM annihilations in our galaxy, direct searches look for interactions of DM with SM particles in underground detectors and colliders attempt to produce the DM and measure it. We will concentrate here on direct detection and collider searches. If dark matter is to be observed in direct detection searches it must couple to quarks or gluons~\\footnote{DAMA is an exception as, unlike other experiments, it does not distinguish between nuclear and electron recoils of DM.}. The same couplings lead to direct DM production at hadronic colliders such as the Tevatron, and we wish to investigate the connection between the two types of search. We will do so in a model independent fashion~\\cite{Birkedal:2004xn}; we will assume that the DM is fermionic and that there is some massive state whose exchange couples DM to quarks. The mediator may be a SM gauge boson, the Higgs or a new particle (if the new particle is very heavy we can describe its effects with an effective contact operator). Although the processes that give direct detection and those that give DM production occur through s- and t-channel exchange of the same mediator, the regimes probed in the two types of experiment are very different. The momentum exchange during a DM-nucleus recoil is $\\sim 100\\ \\mev$ whereas at the Tevatron the typical momentum exchange is $10-100\\ \\gev$. This leads to two interesting regimes to consider when comparing bounds from the two types of experiments: heavy mediators $M\\gtrsim 100\\ \\gev$ and light mediators $M\\lesssim 100\\ \\gev$. The momentum exchange at direct detection experiments is sufficiently low that for all but the lightest mediators below ${\\cal O}(100~\\mbox{MeV})$, which we do not consider here, the mediator can effectively be integrated out and the scattering rate in both regimes scales as, \\be \\sigma_{\\rm DD}\\sim g_\\chi^2\\, g_q^2\\, \\frac{\\mu^2}{M^4}\\,, \\ee where, for simplicity, we have ignored form factors and possible momentum and velocity dependence in the cross section. Here, $g_\\chi$ and $g_q$ are couplings of the mediator to DM and quarks. $\\mu$ is the reduced mass of the DM-nucleon system. In contrast the two regimes behave very differently at colliders. Concentrating on direct production of a pair of DM particles and an initial state emission of a jet, we estimate the mono-jet + $\\slsh{E_T}$ partonic production cross section in the two cases to be \\be \\sigma_{1j} \\sim \\begin{cases} \\alpha_s \\,g_\\chi^2\\, g_q^2\\,\\frac{1}{p_T^2} & \\quad M\\lesssim p_T \\,, \\vspace{3mm}\\\\ \\alpha_s \\,g_\\chi^2\\, g_q^2\\, \\frac{p_T^2}{M^4} &\\quad M\\gtrsim p_T \\,, \\\\ \\end{cases} \\label{eq:colliderqual} \\ee where $\\alpha_s$ is the QCD coupling and $p_T$ is the transverse momentum of the jet, which is typically $\\sim\\mathcal{O}(100)\\,\\gev$ at the Tevatron\\footnote{Note that (\\ref{eq:colliderqual}) is only qualitative in nature. The limits are correct for mediator masses well above and below the $p_T$ of the jet.}. Thus, for the heavy mediator case the (partonic) production cross section at the Tevatron, where $p_T\\sim 100\\ \\gev$, is $\\mathcal{O}(1000)$ times larger than the direct detection cross section for $\\mu \\sim 1$~GeV when the DM is heavier than the nucleon mass. The CDF mono-jet search~\\cite{CDFmonojet} analysed $\\sim1\\ {\\rm fb}^{-1}$ and saw no significant discrepancy from the SM, thus limiting the DM + mono-jet production cross section to be smaller than $\\sim 500\\ {\\rm fb}$. Due to the factor of 1000 mentioned above, this will translate to bounds in the neighborhood of $0.5$ fb in direct detection experiments, the exact bound at direct detection experiments will depend upon the details of the parton density functions relating the partonic cross section of (\\ref{eq:colliderqual}) to the actual CDF mono-jet bound. This is to be compared with direct detection current searches. Null results from experiments such as CDMS~\\cite{Ahmed:2009zw}, XENON\\cite{Angle:2007uj,Aprile:2010um} and others, place strong constraints on the cross section of DM to recoil from a nucleus, $\\sigma \\lesssim 10^{-3}-10^{-4}~{\\rm fb}$ for a 10-100 GeV WIMP scattering elastically through a spin independent (SI) interaction. Thus, for this situation it seems that direct detection has greater reach. However, due to the threshold to detect a DM recoil in these experiments there is a DM mass below which these experiments are no longer sensitive, typically this lower bound is $m_\\chi\\sim 5-10\\ \\gev$, there is no such threshold in collider searches. Furthermore, the DAMA collaboration~\\cite{Bernabei:2008yi} have observed a signal consistent with DM scattering from NaI % which is inconsistent with bounds on a standard WIMP from CDMS and other experiments. This has motivated the introduction of non-standard DM scenarios that can make these seemingly discrepant results consistent. The cross sections necessary to explain DAMA are considerably larger than $10^{-3} {\\rm fb}$ and may allow these scenarios to be probed directly at the Tevatron, due to the increase in cross section described above. Another possibility that has been motivated both by DAMA and the recent CoGeNT~\\cite{Aalseth:2010vx} excess is that dark matter is light, below about 10 GeV, and is thus transfers small momenta to nuclei giving a signal near threshold. The Tevatron will place a strong bound for dark matter particles below 5 GeV. Finally, spin-dependent (SD) WIMP-nucleus scatterings are not coherent and therefore are not enhanced by an $A^2$ factor. Typical bounds on a SD WIMP-proton scatter from direct detection are $\\sim 1$ fb , and will be severely impacted by the mono-jet bounds presented here. We will begin our discussion with a model independent operator analysis, corresponding to very heavy mediation particles (such as a heavy $Z'$ or squarks). In Section~\\ref{sec:Ops} we will introduce some representative four fermion operators supressed by a cutoff scale. We will then place limits on the strengths of these operators from the Tevatron mono-jet search. In Section~\\ref{sec-DD} we will translate the Tevatron bounds to limits on direct detection cross section for different dark matter scenarios. In Section~\\ref{sec:lightmediator} we move on to introduce lighter mediators that are kinematically accesible at the Tevatron and find that these can either slightly enhance or severely weaken the Tevatron bounds. In Section~\\ref{sec-discussion} we will discuss possible enhancements to the Tevatron dark matter search using the mono-jet $p_T$ spectrum, and conclude. ", "conclusions": "\\label{sec-discussion} \\begin{figure}[t] \\begin{center} \\includegraphics[width=0.48\\textwidth]{Puu10GeV.eps} \\vspace{2mm} \\includegraphics[width=0.48\\textwidth]{PCalcMad.eps}\\\\ (a)\\qquad\\qquad\\qquad\\qquad\\qquad\\qquad\\qquad\\qquad\\qquad\\qquad(b) \\caption{(a) Comparisons of the shapes of the signal, the SM background and CDF measured events. The SM predictions are shown in the green and the CDF observed data are shown in red. (b) Comparisons of simulated signal events from two different Monte-Carlo tools and for the parton and the particle levels. The cutoff $\\Lambda \\equiv M/\\sqrt{g_\\chi g_q}$ is chosen to be 1 TeV.} \\label{fig-spectra} \\end{center} \\end{figure} It is worthwhile to consider possible improvements to the dark matter search at the Tevatron, and in the future at the LHC. Here we placed bounds on dark matter using only the total rate of mono-jet signal events above a certain $p_T$ cut. An analysis that takes the spectrum shape into account may yield more powerful bounds. We show the spectral shape of the signal compared to the background in Figure~\\ref{fig-spectra}(a). We find that the signal spectrum is somewhat harder than the background, especially when the messenger mass is much higher than the dark matter mass. We find that including showering, hadronization (using Pythia~\\cite{Sjostrand:2006za}) and a detector simulation (PGS~\\cite{PGS}) does not change the signal shape significantly, particularly above 100 GeV, as is shown in Figure~\\ref{fig-spectra}(b). This may allow us to place tighter constraints using a multi-bin analysis as compared with a simple counting experiment, since signal predicts more deviations in high $p_T$ bins. However, this would require knowledge of the theoretical uncertainty on a bin-by-bin basis which is not presently available. Furthermore, a bound may be extracted from mono-photon events. In this work we show that the Tevatron mono-jet search places competitive bounds on dark matter-nucleus cross sections relevant for direct detection experiments. In particular, the Tevatron limits are the current world-best for light dark matter, below a mass of 5 GeV. The Tevatron also sets the best limit spin dependent dark matter scattering. Various models built to explain the DAMA modulation signal such as inelastic and exothermic dark matter are also constrained by current Tevatron searches. In addition to considering dark matter that couples to quarks via contact interactions we have taken the possibility of light mediators, as motivated by cosmic ray excesses~\\cite{ArkaniHamed:2008qn} into account. We find that the introduction of a light mediator of mass $\\lesssim 10$ GeV alleviates the mono-jet bounds completely for most cases. This leads to an interesting conclusion - if a direct dark matter signal is established in a region that is in conflict with collider bounds, a new light state should be introduced to reconcile the data. The Tevatron is unable to search for DM above a few 100 GeV due to kinematics, an upper bound that will be raised at the LHC. The current powerful bounds using only 1 fb$^{-1}$ motivates a dedicated analysis using more Tevatron data as well as future analyses at the LHC. \\vspace{0.5cm} \\textbf{Note added:} While in the final stages of this work Ref.~\\cite{Goodman:2010yf} appeared on the arXiv. They address similar issues, focussing on Majorana DM, but do not consider light mediators. \\subsection*{Acknowledgements} We would like to thank John Campbell, Walter Giele, Joachim Kopp, Adam Martin and Neal Weiner for useful discussions. We are grateful to Joachim Kopp for providing the contours used in Figure~\\ref{fig-iDM-limits}. PJF would like to thank the CCCP at NYU, and the UC Davis HEFTI workshop on Light Dark Matter for hospitality while part of this work was undertaken. Fermilab is operated by Fermi Research Alliance, LLC, under Contract DE-AC02-07CH11359 with the United States Department of Energy. \\appendix" }, "1005/1005.1792_arXiv.txt": { "abstract": "In Ruderman \\& Sutherland (RS75) model, the normal neutron stars as pulsars bear a severe problem, namely the binding energy problem that both ions (e.g., ${}_{26}^{56}$Fe) and electrons on normal neutron star surface can be pulled out freely by the unipolar generator induced electric field so that sparking on polar cap can hardly occur. This problem could be solved within the Partially Screened Gap (PSG) model in the regime of neutron stars. However, in this paper we extensively study this problem in a bare strange quark star (BSS) model. We find that the huge potential barrier built by the electric field in the vacuum gap above polar cap could usually prevent electrons from streaming into the magnetosphere unless the electric potential of a pulsar is sufficiently lower than that at infinite interstellar medium. Other processes, such as the diffusion and thermionic emission of electrons have also been included here. Our conclusions are as follows: both positive and negative particles on a BSS's surface would be bound strongly enough to form a vacuum gap above its polar cap as long as the BSS is not charged (or not highly negative charged), and multi-accelerators could occur in a BSS's magnetosphere. Our results would be helpful to distinguish normal neutron stars and bare quark stars through pulsar's magnetospheric activities. ", "introduction": "Although pulsar-like stars have many different manifestations, they are populated mainly by rotation-powered radio pulsars. A lot of information about pulsar radiative process is inferred from the integrated and individual pulses, the sub-pulses, and even the micro-structures of radio pulses. Among the magnetospheric emission models, the user-friendly nature of Ruderman \\& Sutherland (1975; hereafter RS75) model is a virtue not shared by others~\\citep{Shukre92}. In RS75 and its modified versions~\\citep[e.g.,][]{QL98}, a vacuum gap exists above polar cap of a pulsar, in which charged particles (electrons and positrons) are accelerated because of ${\\bf E \\cdot B} \\neq 0 $. These accelerated charged particles, moving along the curved magnetic field lines, radiate curvature or inverse-Compton-scattering-induced high energy photons which are converted to $e^\\pm$ while propagating in strong magnetic field. A follow-up breakdown of the vacuum gap produces secondary electron-positron pairs plasma that radiate coherent radio emission. These models with gap-sparking provide a good framework to analyze observational phenomena, especially the drifting~\\citep{DC68,DR99,VJ99} and bi-drifting~\\citep{QLZXW2004} sub-pulses. However, the RS75-like vacuum gap models work only in strict conditions: strong magnetic field and low temperature on surface of pulsars~\\citep[e.g.,][]{GH08,ML07}. The necessary binding energy of positive ions (e.g., ${}_{26}^{56}$Fe) for RS75 model to work should be higher than $\\sim 10$ keV, while calculations showed that the cohesive energy of ${}_{26}^{56}$Fe at the neutron star surface is $<1$ keV~\\citep{FLRSHM77,L01}. This binding energy problem could be solved within a partially screened inner gap model~\\citep{GM03,GM06,MG09} for normal neutron stars. Alternatively, it is noted that the binding energy could be sufficiently high if pulsars are bare strange quark stars~\\citep{XQ98,XQZ99,XZQ01} although strange stars were previously supposed to exist with crusts~\\citep{AFO86}. Certainly, it is very meaningful in the elementary strong interaction between quarks and the phases of cold quark matter that the binding energy problem could be solved by bare quark stars as pulsars~\\citep{X09,X10}. Though the ideas of solving the binding energy problem in BSS model were presented and discussed in some literatures, there is no comprehensive study with quantitative calculations up to now. In this paper, we are going to investigate the BSS model in quantitative details and show the physical picture of binding of particles on BSS's surface. Our research results are that multi-accelerators could occur above the polar cap for (and only for) the curvature-radiation-induced (CR-induced) sparking normal pulsars (NPs), but for other cases, such as resonant inverse-Compton-scattering-induced (ICS-induced) sparking NPs and both CR-induced and ICS-induced millisecond pulsars (MSPs), particles on surface of BSSs are bound strongly enough to form vacuum gap and RS75-like models work well if pulsars are BSSs. ", "conclusions": "In RS75 model, the binding energy problem is one of the most serious problems in the normal neutron star model of pulsars. Arons and Scharlemann (1979) developed an alternative model, the space-charge limited flow (SCLF) model, in which the particles, both iron ions and electrons can be pulled out freely, and form a steady flow~\\citep{AS79}. In this SCLF model, the drifting sub-pulse phenomenon which has been commonly observed in pulsars can rarely be reproduced. The prerequisite for understanding this phenomenon could be the existence of a vacuum gap. In a very special case, through our calculations, we find that there is a new physical scenario for CR-induced sparking of normal pulsars (NPs) that free flow and vacuum gap may coexist above the polar cap. But in other cases, such as ICS-induced sparking of NPs and millisecond pulsars (MSPs), only vacuum gap exists. In general, if a pulsar is not highly negatively charged~\\citep{XCQ06}, vacuum gap survives at polar cap as well. One limitation is that our calculation is based on one-dimensional approximation and it might fail in some cases of MSPs. As far as we find, it is very difficult to deal with the high-dimensional cases. The one-dimensional approximation provides a good understanding of the geometry of polar cap of BSSs. In conclusion, the binding energy problem could be solved completely in the BSS model of pulsar as long as BSSs are neutral (or not highly negative charged), and the structure of polar cap of BSSs are very different with respect to that of NSs. Detailed information about the geometry of BSS's polar cap is given in Table~\\ref{AG}. \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{The accelerators above polar caps of BSSs. \\label{AG}} \\begin{tabular}{@{}llllll@{}} \\hline &\\multicolumn{2}{c}{[0,~$\\theta_{\\rm 0}$]$^{\\dag}$} &\\multicolumn{2}{c}{[$\\theta_{\\rm 0}$,~$\\theta_{\\rm A}$]} & \\\\ \\cline{2-5}\\\\ &CR &ICS &CR &ICS & \\\\ \\hline &SCLF &VG &VG &VG &$\\bf \\Omega \\cdot B > 0$ \\\\ NPs &VG &VG$^{\\ddag}$ &SCLF &VG$^{\\ddag}$& $\\bf \\Omega \\cdot B < 0$ \\\\ &VG &VG &VG &VG &$\\bf \\Omega \\cdot B >0 $ \\\\ MSPs &VG$^{\\ddag}$ &VG &VG$^{\\ddag}$ &VG &$\\bf \\Omega \\cdot B < 0$\\\\ \\hline \\end{tabular}\\\\ $^{\\dag}${$\\theta_{0}$ represents $\\theta_{\\rm 0,B}$ while choosing $\\phi_{\\rm B}=0$ and $\\theta_{\\rm 0,C}$ while choosing $\\phi_{\\rm C}=0$.}\\\\ $^{\\ddag}${for such cases, $\\theta_{0}$ $>$~$\\theta_{\\rm A}$, which represents the structure of the whole polar cap region.} \\end{minipage} \\end{table*} A more interesting region from pole to equator may locate between that polar angle where the total energy of electron equals the potential barrier and the polar angle of the foot of zero potential magnetic field line (i.e., [$\\theta_{0, \\rm C},\\theta_{\\rm C}$] or [$\\theta_{0, \\rm B},\\theta_{\\rm B}$], see Fig.~\\ref{PEC}) for CR-induced sparking NPs. After the birth of a NP, a vacuum gap exists at this region. When sparking starts, the potential in vacuum gap drops rapidly due to screen by electron-positron pairs and may become lower than that at the surface, namely $V_{\\rm i}(\\theta)$. As a result, the sparking converts vacuum gap to free flow at this region until the sparking ends, i.e., at [$\\theta_{0, \\rm C},\\theta_{\\rm C}$] or [$\\theta_{0, \\rm B},\\theta_{\\rm B}$], vacuum gap and free flow work alternately. This argument may have profound implications for us to distinguish neutron stars and quark stars by pulsar's magnetospheric activities (e.g., the diversity pulse profiles). Another issue to be discussed is about the drifting rate of subpulses when we use the height of pure vacuum gap in this work. The natural explanation of the drifting subpulse phenomena in vacuum gap is due to $\\bf E \\times B$. Unfortunately, these theoretical calculations gave higher drifting rate with respect to observations~\\citep[e.g.,][]{RS75, DR99, DR01, GM03, GMZ06}. Since it has been observed~\\citep{DC68}, the drifting subpulse phenomenon remains unclear which has been widely regarded as one of the most critical and potentially insightful aspects of pulsar emission~\\citep{DR01}. The PSG mechanism~\\citep[e.g.,][]{GM03, GM06, GMZ06} could be a way to understand lower drifting rates observed, but some complexities still exist which make the underlying physics of drifting subpulses keep complicated and far from knowing clearly. (1) in principle the drifting velocity of subpulses is the ratio of the drifting distance to the duration, while the expected velocity predicted by $\\bf E \\times B$ is only for electrons in separated emission units, namely the plasma filaments. These two velocities would not be the same if the plasma filaments may stop after sparking. When sparking starts, the electric field in the vacuum gap vanishes due to screen by plasmas; while sparking ends, the electric field appears again. Thus, the calculated drifting velocity with $\\bf E \\times B$ could be higher than that of observations. (2) the so-called aliasing effect: as one observes subpulses only once every rotation period, we can hardly determine their actual speed. The main obstacles in the aliasing problem are the under sampling of subpulse motion and our inability to distinguish between subpulses especially when the differences between subpulses formed by various subbeams are smaller than the fluctuations in subpulses from one single subbeam~\\citep{LSRR03}. Anyway, detailed studies are very necessary in the future works. We assume that the potential energy related to Eq.~\\ref{UGP}, $eV_{\\rm i}$, to be the constant, $\\phi_{0}$, in Eq.~\\ref{FD}. This assumption could be reasonable. For an uniformly magnetized, rotating conductor sphere, the unipolar generator will induce an electric field which is a function of polar angle, as described in Eq.~\\ref{UGP}. In the case of $\\bf \\Omega \\cdot B>0$ (Fig.~\\ref{antipulsar}), the potential energy of electron is highest at the polar region which means that those electrons there could be easier to escape. Alternatively, this conclusion could be quantitatively understood as following: because of Lorentz force inside a star, more electrons locate at the polar region so that the Fermi energy of electron is higher there and easier to escape into magnetosphere." }, "1005/1005.1955_arXiv.txt": { "abstract": "The Megamaser Cosmology Project (MCP) aims to determine $H_0$ by measuring angular-diameter distances to galaxies in the Hubble flow using observations of water vapor megamasers in the circumnuclear accretion disks of active galaxies. The technique is based only on geometry and determines $H_0$ in one step, independent of standard candles and the extragalactic distance ladder. In Paper I we presented a VLBI map of the maser emission from the Seyfert 2 galaxy UGC 3789. The map reveals an edge-on, sub-parsec disk in Keplerian rotation, analogous to the megamaser disk in NGC 4258. Here we present 3.2 years of monthly GBT observations of the megamaser disk in UGC 3789. We use these observations to measure the centripetal accelerations of both the systemic and high-velocity maser components. The measured accelerations suggest that maser emission lines near the systemic velocity originate on the front side of the accretion disk, primarily from segments of two narrow rings. Adopting a two-ring model for the systemic features, we determine the angular-diameter distance to UGC 3789 to be 49.9 $\\pm$ 7.0 Mpc. This is the most accurate geometric distance yet obtained to a galaxy in the Hubble flow. Based on this distance, we determine $H_0$ = 69 $\\pm$ 11 km s$^{-1}$ Mpc$^{-1}$. We also measure the mass of the central black hole to be 1.09 $\\times$ 10$^7$ $\\solmass$ $\\pm 14\\%$. With additional observations the uncertainty in the distance to this galaxy can be reduced to under 10\\%. Observations of megamaser disks in other galaxies will further reduce the uncertainty in $H_0$ as measured by the MCP. ", "introduction": "As a complement to observations of the cosmic microwave background (CMB), an independent measurement of the Hubble Constant, $H_0$, within 3\\% rms would place a valuable constraint on the dark energy equation-of-state parameter, $w$ (e.g. Hu 2005, Olling 2007). The most widely accepted value for $H_0$ is 72 $\\pm$ 7 km s$^{-1}$ Mpc$^{-1}$ (Freedman et al. 2001), based on the ``extragalactic distance ladder'' and using the Large Magellanic Cloud (LMC) to calibrate Cepheid variables which are treated as standard candles. The Cepheid metallicity correction has been controversial, as highlighted by Sandage et al. (2006), who used similar methods but a different metallicity correction to obtain $H_0$ = 62 $\\pm$ 6 km s$^{-1}$ Mpc$^{-1}$. If measured at the $\\lesssim$ 3\\% level, a value of $H_0$ = 72 km s$^{-1}$ Mpc$^{-1}$ would be consistent with the cosmological constant as an explanation of dark energy ($w$ = -1), whereas $H_0$ = 62 km s$^{-1}$ Mpc$^{-1}$ would challenge that model (e.g. Figure 14 of Riess et al., 2009). Macri et al. (2006) used NGC 4258 rather than the LMC to calibrate Cepheids for the extragalactic distance scale and measured $H_0$ = 74 $\\pm$ 7 km s$^{-1}$ Mpc$^{-1}$. Riess et al. (2009) subsequently refined the measurement, also using Cepheids in NGC 4258 to calibrate those in selected SN Ia host galaxies, and they determined $H_0$ = 74.2 $\\pm$ 3.6 km s$^{-1}$ Mpc$^{-1}$ (5\\%). Riess et al. are able to reduce the systematic uncertainty in $H_0$ compared to the LMC-based results because the Cepheids in NGC 4258 are more similar in metallicity and period to those in the galaxies being calibrated. NGC 4258 is a good foundation for the extragalactic distance scale because its distance (7.2 Mpc $\\pm$ 7\\%) is known to high accuracy from observations of circumnuclear water megamasers in its AGN accretion disk (Herrnstein et al. 1999). NGC 4258 is too close, though, for determining $H_0$ directly because its recession velocity may be highly influenced by its peculiar motion. Lo (2005) gives a review of megamasers, including a discussion of the maser distance technique used in NGC 4258. The Megamaser Cosmology Project (MCP) aims to determine $H_0$ by measuring angular-diameter distances to galaxies in the Hubble flow, $\\sim$ 50 -- 200 Mpc distant, using the maser technique. Being independent of Cepheids and the CMB, the MCP will provide a valuable check of these other methods for measuring $H_0$. The project is a multi-year effort that begins with surveys for new megamaser disks. Measuring a distance to each galaxy detected with a megamaser disk requires high-fidelity VLBI imaging, which we pursue with the High Sensitivity Array (VLBA+GBT+VLA+Effelsberg), and spectral-line monitoring to measure the centripetal accelerations of individual lines in the maser spectrum. The MCP aims to measure maser distances with $\\sim$ 10\\% or better accuracy to each of about 10 galaxies, thereby determining H$_0$ to $\\sim$ 3\\%. In a survey with the Green Bank Telescope, Braatz \\& Gugliucci (2008; hereafter BG08) discovered water maser emission from the Seyfert 2 galaxy UGC 3789. The maser spectrum has three sets of Doppler components spaced nearly symmetrically at and around the galaxy's systemic recession velocity (Figure 1). This profile is characteristic of megamasers in an edge-on accretion disk. The megamaser in UGC 3789 is similar to that in NGC~4258, where ``high-velocity'' blue-shifted and red-shifted maser components arise from the tangent points of the disk rotating toward and away from the observer, and components near the systemic recession velocity originate on the near side of the disk, along the line of sight to the central black hole. Rotation velocities indicated by the high-velocity maser lines in UGC 3789 are $\\sim$ 400--800 km s$^{-1}$. By monitoring the maser spectrum roughly monthly over 6 months, BG08 measured secular, redward drifts in radial velocity of 2 -- 8 km s$^{-1}$ yr$^{-1}$ among the systemic group of maser components, while the red- and blue-shifted features remained relatively fixed in velocity. The radial velocity drift is interpreted as centripetal acceleration in the maser gas as it orbits the central supermassive black hole. The spectral profile and the measured velocity drifts demonstrate that the maser disk in UGC 3789 is suitable for measuring an angular-diameter distance to the galaxy, using the method pioneered by Herrnstein et al. (1999). In Paper I from the Megamaser Cosmology Project (Reid et al. 2009), we presented a VLBI image of the maser disk in UGC 3789. The masers reveal nuclear gas in an edge-on disk ranging in radii from 0.08 to 0.30 pc. The high-velocity lines trace a Keplerian rotation curve with high precision. While the high-velocity lines indicate relatively little disk warping, the systemic features show some structure that suggests the front side of the disk may be warped, tilted, or geometrically thick. Here we present single-dish observations of the UGC 3789 maser made with the Green Bank Telescope over a period of 3.2 years. Using these observations we refine the acceleration measurements made by BG08 and we present a model to estimate the distance to UGC 3789 using the maser observations. ", "conclusions": "The peculiar radial velocity of UGC 3789 relative to the CMB is -151 $\\pm$ 163 km s$^{-1}$ (Masters, private communication), using a flow model based on the SFI++ sample (Masters et al. 2006; Springob et al. 2007). Here peculiar velocity is defined as $v_{pec} = v_{rec} - H_0 D$, where v$_{rec}$ is the galaxy's observed recession velocity. Writing the UGC 3789 maser radial velocity in the CMB frame ($v_{CMB} \\simeq v_{LSR}$ + 60 km s$^{-1}$) and correcting for the peculiar velocity, we obtain a relativistic, recessional flow velocity of 3481 $\\pm$ 163 km s$^{-1}$. Applying a standard $\\Lambda$CDM model with $\\Omega_m$ = 0.26, we therefore determine $H_0$ = 69 $\\pm$ 11 km s$^{-1}$ Mpc$^{-1}$. The main source of uncertainty in the distance to UGC 3789 comes from the measurement of the orbital curvature parameter $\\Omega$. The uncertainty in the acceleration is also a significant contributor, while the contribution from the Keplerian rotation constant is negligible. Therefore we can best reduce the total uncertainty in the distance to UGC 3789 by obtaining a more sensitive VLBI map. We are currently processing additional VLBI observations of UGC 3789, which we can average to increase the map sensitivity. Improving the measurement of the acceleration profile requires more sensitive monitoring observations. Recognizing this need, we increased the duration of GBT observations after May 2007. These efforts should reduce the overall distance uncertainty to $<$ 10\\%, which would determine $H_0$ with accuracy comparable to the Hubble Key Project, but with entirely independent data and methods. Ultimately, to reach an uncertainty in $H_0$ of $\\lesssim$ 3\\% from the Megamaser Cosmology Project, we require distances to at least 10 galaxies comparable to UGC 3789. We are currently pursuing observations of the megamaser galaxies Mrk 1419, NGC 6264, and NGC 6323, which could eventually lead to comparable distance measurements. There are another $\\sim$ 6 maser disk galaxies known that may also be suitable for distance measurements, but the quality of these (i.e. brightness of the masers, richness of the spectrum, and velocity coverage) is not as good. Surveys for additional maser galaxies therefore remain an important part of the MCP." }, "1005/1005.1487_arXiv.txt": { "abstract": "{The Photodetector Array Camera and Spectrometer (PACS) is one of the three science instruments on ESA's far infrared and submillimetre observatory. It employs two Ge:Ga photoconductor arrays (stressed and unstressed) with $16\\times25$ pixels, each, and two filled silicon bolometer arrays with $16\\times 32$ and $32\\times 64$ pixels, respectively, to perform integral-field spectroscopy and imaging photometry in the 60--210\\,$\\mu$m wavelength regime. In photometry mode, it simultaneously images two bands, 60--85$\\,\\mu$m or 85--125$\\,\\mu$m and 125--210$\\,\\mu$m, over a field of view of $\\sim 1.75'\\times 3.5'$, with close to Nyquist beam sampling in each band. In spectroscopy mode, it images a field of $47\\arcsec \\times 47\\arcsec$, resolved into $5\\times 5$ pixels, with an instantaneous spectral coverage of $\\sim 1500$~km/s and a spectral resolution of $\\sim 175$~km/s. We summarise the design of the instrument, describe observing modes, calibration, and data analysis methods, and present our current assessment of the in-orbit performance of the instrument based on the Performance Verification tests. PACS is fully operational, and the achieved performance is close to or better than the pre-launch predictions.} ", "introduction": "The PACS instrument was designed as a general-purpose science instrument covering the wavelength range $\\sim$60--210\\,$\\mu$m. It features both, a photometric multi-colour imaging mode, and an imaging spectrometer. Both instrument sections were designed with the goal of maximising the science return of the mission, given the constraints of the \\emph{Herschel} platform (telescope at $T \\approx 85\\ K$, diffraction limited for $\\lambda > 80\\,\\mu$m, limited real estate on the cryostat optical bench) and available FIR detector technology. \\subsection {Photometer rationale} Photometric colour diagnostics requires spectral bands with a relative bandwidth $\\Delta\\lambda/\\lambda < 0.5$. In coordination with the longer wavelength SPIRE bands, the PACS photometric bands have been defined as 60--85$\\,\\mu$m, 85--130$\\,\\mu$m, and 130--210$\\,\\mu$m, each spanning about half an octave in frequency. A large fraction of the \\emph{Herschel} observing time will be spent on deep and/or large scale photometric surveys. For these, mapping efficiency is of the highest priority. Mapping efficiency is determined by both, the field of view of the instrument (in the diffraction-sampled case, the number of pixels) and the sensitivity per pixel. The PACS photometer was therefore designed around the largest detector arrays available without compromising sensitivity. Simultaneous observation of several bands immediately multiplies observing efficiency. By implementing two camera arrays, PACS can observe a field in two bands at a time. Extracting very faint sources from the bright telescope background requires means to precisely flat-field the system responsivity on intermediate time-scales, as well as the use of spatial modulation techniques (chopping/nodding, scan-mapping) to move the signal frequency from ``DC'' into a domain above the 1/f ``knee'' of the system, including - most notably - the detectors. \\subsection{Spectrometer rationale} Key spectroscopic observations, particularly of extragalactic sources, ask for the detection of faint spectral lines with medium resolution ($R \\sim 1500$). The power emitted or absorbed by a single spectral line in the far-infrared is normally several orders of magnitudes lower than the power in the dust continuum over a typical photometric band. Sensitivity is thus the most important parameter for optimisation; with background-limited detector performance the best sensitivity is obtained if the spectrometer satisfies the following conditions: The detection bandwidth should not be greater than the resolution bandwidth, which in turn should be matched to the line width of the source, and, the line flux from the source must be detected with the highest possible efficiency in terms of system transmission, spatial and spectral multiplexing. Again, subtraction of the high telescope background has to be achieved by appropriate spatial and/or spectral modulation techniques. ", "conclusions": "With the PACS instrument, we have -- for the first time -- introduced large, filled focal plane arrays, as well as integral-field spectroscopy with diffraction-limited resolution, in the far infrared. Without any prior demonstration, several major, technological developments have found their first application with PACS in space. The success of this path -- documented by the results in this volume -- should encourage our community to defend this more ``pioneering'' approach against trends towards an ``industrial'' (i.e., minimal-risk) approach, which will not allow us to take advantage of the latest, experimental developments, on which our scientific progress often depends." }, "1005/1005.3357_arXiv.txt": { "abstract": "We report the result of VLBI observation of the giant radio galaxy J1313+696 (4C +69.15) at 2.3/8.4 GHz, only the core component of the giant radio galaxy was detected in the VLBI observation at the dual frequencies. The result shows a steep spectrum core with $\\alpha=-0.82$ ($S \\propto \\nu^{\\alpha}$) between 2.3 GHz and 8.4 GHz. The steep spectrum core may be a sign of renewed activity. Considering also the upper limit flux density of 2.0 mJy at 0.6 GHz from Konar et al. 2004 the core has a GHz-peaked spectrum, implying that the core is compact and absorbed. Further high resolution VLBI observations are needed to identify if the steep spectrum core is consisting of a core and steep spectrum jet. ", "introduction": "It's known that the large radio sources are classified into two types i.e. FRI and FRII-type according to the morphology of the radio sources. Among the large radio sources, the radio galaxies whose lobes span a (projected) distance of above 1 Mpc are called giant radio galaxies (GRGs), and the majority of them are FRII sources (Schoenmakers et al. 2001). A small number of the large extended sources consisting of a pair of double lobes have been called double-double radio galaxies (DDRGs), and the observed structures of DDRGs suggest recurrent or interrupted central activity as the origin of these sources (Schoenmakers et al. 2000a; Saikia et al. 2006). The large radio sources have been observed with telescopes or arrays at relatively low resolution but not well observed with the very long baseline interferometry (VLBI). The VLBI observation is able to identify the core of the large radio sources at high resolution. We have included the GRG J1313+696 in our sample of VLBI observation with the European VLBI Network (EVN) at dual frequency 2.3/8.4 GHz simultaneously. ", "conclusions": "J1313+696 (B1312+698, DA340, 4C +69.15, z=0.106), is an FRII type giant radio galaxy. In VLA observation at 1.4 GHz the emission from the core to the lobes forms a continuous bridge in SE-NW direction (Lara et al. 2001), the GMRT observation at 605 MHz shows a similar structure (Konar et al. 2004). But at 4.9 GHz only the core and the lobe extremes are detected with the VLA (Lara et al. 2001), and the core position is consistent with the position of the associated galaxy. The source was not classified as a DDRG according to the definition of Schoenmakers et al. (2000a) which an inner pair of edge-brightened lobes should be detected. The VLBI images (Fig.~\\ref{fig1}, Fig.~\\ref{fig2}) of J1313+696 at 2.3/8.4 GHz show a point source. We checked the position of the point source in the VLBI maps with the VLA images at 1.6 and 5 GHz (Lara et al. 2001), confirmed that the VLBI point source is the core of J1313+696 in the VLA images. We collected the core flux densities of J1313+696 from our data and literature as shown in Fig.~\\ref{fig3}, the error bars are 1$\\sigma$ (per beam) from our data and the literature. The source flux densities in the VLBI observation at 2.3 GHz and 8.4 GHz are 7.5$\\pm$1.1 mJy and 2.6$\\pm$0.2 mJy respectively. It is 10.2 mJy at 1.4 GHz and 3.8 mJy at 4.9 GHz in Lara et al. (2001), and 6 mJy at 2.7 GHz in Saunders et al. (1987). It is 7 mJy at 1.4 GHz, 4.3 mJy at 4.9 GHz, and an upper limit of 2 mJy at 605 MHz in Konar et al. (2004). The spectral index between 2.3 and 8.4 GHz in the VLBI observation is $-0.82$ (we use $S \\propto \\nu^{\\alpha}$). The spectral index from the collected data above 1.4 GHz can be fitted linearly (Fig.~\\ref{fig3}) with $\\alpha=-0.78\\pm0.03$. It is interesting that this value is consistent with the spectral indices of the east, west lobe of J1313+696 (Schoenmakers et al. 2000b). However, because an upper limit 2 mJy at 605 MHz was estimated by Konar et al. (2004) the integrated spectrum in Fig.~\\ref{fig3} can also be an inverted spectrum which peaked around 1.4 GHz. We note that the 1.4 GHz core flux (10.2 mJy) from Lara et al. (2001), might be overestimated due to contamination of diffuse emission. As Lara et al. mentioned, for this core the ratio peak/total $<0.8$. An ideal point source should have the ratio peak/total $\\sim$1. The core is sitting on top of a diffuse emission, as is evident from the map in Lara et al.. Whereas, the measurement of Konar et al. (2004) has less contamination as they re-mapped the field with lower uv-cutoff to lose the diffuse flux and get the core flux as correct as possible. Konar et al. (2004) have discussed the steep-spectrum cores (SSCs, $\\alpha_{core} < -0.5$) of GRGs, in their sample at least 3 out of 17 sources show SSCs. In Lara et al. (2004) sample a compact radio core was detected at 4.9 GHz with VLA in all sources, with an average spectral index of $0.07\\pm0.41$ for FRII and $-0.24\\pm0.52$ for FRI sources; hence FRIs seem to have more SSCs than FRIIs according to the averaged spectral indices. However, Lara et al. (2004) explained that the core spectral index in FRIs suffers from more contamination from the steeper jet emission than in FRIIs. In Konar et al. (2004) sample, 3 FRII-type sources have been suggested to have SSCs besides J1313+696. They suggested that SSC is preferentially occurred in giant radio galaxies, and SSC is related to the interrupted activity of GRGs. Schoenmakers et al. (2000b) found a polarized component of J1313+696 near but not at the core position, they suggested it is a jet component. The VLBI images of J1313+696 show a point source, not as expected to show a core-jet, but reveal a steep spectrum core which may imply a renewed jet in the core. Considering the upper limit flux density at 605 MHz the core of the J1313+696 has a GHz-peaked spectrum (GPS), indicating the core is very compact and absorbed at lower frequency. We estimated an upper limit of the core size at 8.4 GHz with the beam size of our VLBI map, it is 3.8 pc (in $H_0=71~km s^{-1} Mpc^{-1}$, $\\Omega_m$=0.27, and $\\Omega_{vac}$=0.73 cosmology). Another example is the DDRG B1834+620, its core shows a steep spectrum at higher frequencies with spectral index the same as that of lobes, and the core spectrum is inverted at lower frequencies, showing a GPS shape (Schoenmakers et al. 2000c). Konar et al. (2008) found a GPS core in the GRG J1155+4029 with GMRT and VLA observations." }, "1005/1005.3816_arXiv.txt": { "abstract": "{ We present an early broad-brush analysis of \\emph{Herschel}/PACS observations of star-forming galaxies in 8 galaxy clusters drawn from our survey of 30 clusters at $z{\\simeq}0.2$. We define a complete sample of $192$ spectroscopically confirmed cluster members down to $L_{\\rm TIR}{>}3{\\times}10^{10}{\\Lsol}$ and $L_K{>}0.25{\\Lsol}$. The average $K$-band and bolometric infrared luminosities of these galaxies both fade by a factor of ${\\sim}2$ from clustercentric radii of ${\\sim}2r_{200}$ to ${\\sim}0.5r_{200}$, indicating that as galaxies enter the clusters ongoing star-formation stops first in the most massive galaxies, and that the specific star-formation rate (SSFR) is conserved. On smaller scales the average SSFR jumps by ${\\sim}25\\%$, suggesting that in cluster cores processes including ram pressure stripping may trigger a final episode of star-formation that presumably exhausts the remaining gas. This picture is consistent with our comparison of the \\emph{Herschel}-detected cluster members with the cluster mass distributions, as measured in our previous weak-lensing study of these clusters. For example, the spatial distribution of the \\emph{Herschel} sources is positively correlated with the structures in the weak-lensing mass maps at ${\\sim}5{\\sigma}$ significance, with the strongest signal seen at intermediate group-like densities. The strong dependence of the total cluster IR luminosity on cluster mass -- $L_{\\rm TIR}{\\propto}M_{\\rm virial}^2$ -- is also consistent with accretion of galaxies and groups of galaxies (i.e.\\ the substructure mass function) driving the cluster IR luminosity. The most surprising result is that roughly half of the \\emph{Herschel}-detected cluster members have redder $S_{100}/S_{24}$ flux ratios than expected, based on the Rieke et al.\\ models. On average cluster members are redder than non-members, and the fraction of red galaxies increases towards the cluster centers, both of which indicate that these colors are not attributable to systematic photometric errors. Our future goals include to intepret physically these red galaxies, and to exploit this unique large sample of clusters with unprecedented multi-wavelength observations to measure the cluster-cluster scatter in S0 progenitor populations, and to intepret that scatter in the context of the hierarchical assembly of clusters. } ", "introduction": "Lenticular galaxies (hereafter S0s) are mainly found in the cores of galaxy clusters at low redshift (e.g.\\ Dressler et al.\\ 1997; Smith et al.\\ 2005; Postman et al.\\ 2005). There is a broad consensus that they are the descendants of gas rich spiral galaxies that have been accreted from the surrounding filamentary structure. However the physics of how spirals are transformed into S0s remains largely unconstrained, with numerous ``S0 progenitor'' populations (e.g.\\ Moran et al.\\ 2006; Poggianti et al.\\ 2000; Geach et al.\\ 2006; Haines et al.\\ 2009a -- hereafter H09a) and physical processes (e.g.\\ Gunn \\& Gott 1972; Moore et al.\\ 1999) discussed in the literature. The broad range of cluster-centric radii at which various S0 progenitors are found reflects the fact that different physical processes act in different environments, for example ram pressure stripping is more effective closer to cluster centers where the intracluster medium (ICM) is denser, and galaxy-galaxy merging is more effective in galaxy groups that are falling into the cluster than in the cluster cores. Moreover, the observational signatures of S0 progenitors are diverse, ranging in wavelength from ultraviolet (UV) emission from A stars in galaxies whose star-formation (SF) has been recently quenched, through optical spectral features including Balmer absorption lines, to mid/far-infrared (IR) emission from dust heated by SF (e.g.\\ Moran et al.\\ 2007; Poggianti et al.\\ 2000; Haines et al.\\ 2009b). Mid- and far-IR properties of cluster galaxies have been studied previously with \\emph{IRAS} (e.g.\\ Leggett et al.\\ 1987; Doyon \\& Joseph 1989), \\emph{ISO} (see Metcalfe et al.\\ 2005 for a review), and \\emph{Spitzer} (e.g.\\ Geach et al.\\ 2006; Fadda et al.\\ 2008; Haines et al.\\ 2009a,b; Bai et al.\\ 2009). A key result from these IR studies is that a significant fraction of the total SF in galaxy clusters is obscured by dust. The inferred levels of SF naturally fit the hypothesis that bulge dominated S0s are descended from late-type spirals. It has also been suggested that dusty S0 progenitors are more common in dynamically active, i.e.\\ merging, galaxy clusters than in so-called ``relaxed'' clusters (e.g.\\ Metcalfe et al.\\ 2005; Geach et al.\\ 2006; Miller et al.\\ 2006). However it has thus far been difficult to test this idea robustly because the intrinsic scatter in levels of SF in clusters appears to be large, (as noted by Kodama et al.\\ 2004), and the sample sizes observed to date are small (i.e.\\ $\\ls2$) within any given redshift bin -- although see H09a for a recent counter-example. We are therefore conducting a systematic wide-field survey of a large statistically well-defined sample of galaxy clusters in a narrow redshift slice at $z{\\simeq}0.2$, as part of the Local Cluster Substructure Survey (LoCuSS\\footnote{http://www.sr.bham.ac.uk/locuss/}). Our goals are to compile a complete inventory of S0 progenitors using data from the far-UV to far-IR, and to relate these populations to the underlying gas physics and hierarchical structure of the host galaxy clusters. We aim to delineate the different physical processes responsible for galaxy transformation in clusters and their surrounding large scale structure, and thus constrain the amplitude of the different physical pathways from spiral to S0 morphology, and how these relate to the dynamical state of the clusters. Our Open Time Key Programme observations with \\emph{Herschel} (Pilbratt et al., 2010), supplemented by existing \\emph{Spitzer} mid-IR observations provide the all-important measurements of the bolometric IR luminosity and mid/far-IR colors of dust-reddened/obscured S0 progenitors. We assume ${\\Ho}{=}70{\\kms}$, ${\\Om}{=}0.3$, ${\\Ol}{=}0.7$. In this cosmology $1{\\kpc}$ at $z{=}0.2$, subtends $0.3''$. All cluster masses and radii relative to an over-density are derived from the weak-lensing analysis of Okabe et al.\\ (2010; hereafter Ok10). ", "conclusions": "We have presented an initial broad-brush analysis of \\emph{Herschel}/PACS observations of $25\\%$ of our sample of 30 galaxy clusters at $z{\\simeq}0.2$, and combined these data with our existing \\emph{Spitzer}, Subaru, \\emph{Chandra}, UKIRT, and MMT data. The main analysis concentrates on a sample of 192 spectroscopically confirmed cluster members with $L_{\\rm TIR}{>}3{\\times}10^{10}L_\\odot$, $L_K{>}0.25L_K^\\star$, $R{<}1.5r_{200}$. The average $K$-band luminosity of these galaxies fades by a factor of almost 2 from the cluster outskirts (${\\sim}1{-}2r_{200}$) to the cluster cores (${\\ls}0.5r_{200}$), although the average specific star-formation rate, as probed by ${\\langle}L_{\\rm TIR}{\\rangle}/{\\langle}L_K{\\rangle}$, is constant across most of this radial range (${\\sim}0.5{-}2r_{200}$), before jumping by 25\\% on smaller scales. This suggests that as gas rich galaxies fall into the clusters (typically in groups -- Fig.~1) ongoing star-formation stops first in the most massive galaxies. As galaxies reach the cluster cores physical processes that operate in high density environments, for example ram pressure stripping and harrassment, then appear to trigger a final episode of star-formation that presumably exhausts the remaining gas supply. This picture is consistent with our comparison of the \\emph{Herschel}-detected cluster members with the cluster mass distributions, as probed by Okabe et al.'s (2010) weak-lensing analysis. First, the spatial distribution of the \\emph{Herschel} sources is positively correlated with the structures in the weak-lensing mass maps at ${\\sim}5{\\sigma}$ significance, with the strongest signal seen at intermediate, group-like densities. Second, the strong dependence of the total cluster IR luminosity on cluster mass ($L_{\\rm TIR}{\\propto}M_{\\rm virial}^2$) is consistent with accretion of galaxies and groups of galaxies driving the cluster IR luminosity. This is because, assuming that IR galaxy mass-to-light ratios are independent of the cluster mass, the scaling relation can be understood as stemming from the $M^2$ dependence of the substructure mass function seen in theoretical models (e.g.\\ Taylor \\& Babul 2005). Third, the IR luminosity density profiles of the clusters generally increase to large radii, with some clusters showing a peak at ${\\sim}r_{200}$. This is qualitatively consistent with the enhanced star-formation rates seen in in-falling galaxy populations by Moran et al.\\ (2005). The most surprising result is that roughly half of the \\emph{Herschel}-detected cluster galaxies have excess flux at $100{\\um}$ over that predicted from current SED models. Cluster members are redder than non-members, and we find a shallow trend of increasing fraction of red IR galaxies towards the cluster centers, both of which suggest that this is a physical effect and not caused by systematic photometric uncertainties. This result will be the focus of more detailed future investigation. Finally, we note that, contrary to previous speculation in the literature, we do not find a strong relationship between cluster IR luminosity and cluster dynamical state. Observations of the full sample will allow us to investigate this issue in more detail." }, "1005/1005.3027_arXiv.txt": { "abstract": "We present near-infrared Ks-band photometry bracketing the secondary eclipse of the hot Jupiter TrES-2b using the Wide-field Infrared Camera on the Canada-France-Hawaii Telescope. We detect its thermal emission with an eclipse depth of \\FpOverFStarPercentAbstractTrESTwo$^{+\\FpOverFStarPercentAbstractPlusTrESTwo}_{-\\FpOverFStarPercentAbstractMinusTrESTwo}$\\% (\\XSigmaTrESTwo $\\sigma$). Our best-fit secondary eclipse is consistent with a circular orbit (a 3$\\sigma$ upper limit on the eccentricity, $e$, and argument or periastron, $\\omega$, of $|$$e$$\\cos$$\\omega$$|$ $<$ \\ECosOmegaAbsoluteThreeSigmaLimitTrESTwo), in agreement with mid-infrared detections of the secondary eclipse of this planet. A secondary eclipse of this depth corresponds to a day-side Ks-band brightness temperature of $T_B$ = \\TBrightTrESTwo$^{+\\TBrightPlusTrESTwo}_{-\\TBrightMinusTrESTwo}$ $K$. Our thermal emission measurement when combined with the thermal emission measurements using Spitzer/IRAC from O'Donovan and collaborators suggest that this planet exhibits relatively efficient day to night-side redistribution of heat and a near isothermal dayside atmospheric temperature structure, with a spectrum that is well approximated by a blackbody. It is unclear if the atmosphere of TrES-2b requires a temperature inversion; if it does it is likely due to chemical species other than TiO/VO as the atmosphere of TrES-2b is too cool to allow TiO/VO to remain in gaseous form. Our secondary eclipse has the smallest depth of any detected from the ground at around 2 $\\mu m$ to date. ", "introduction": "The first detection of the transit of an exoplanet in front of its parent star (\\citealt{Charb00}; \\citealt{Henry00}) opened a new avenue to determine the characteristics of these exotic worlds. For all but the most eccentric cases, approximately half-an-orbit after their transits these planets pass behind their star along our line of sight allowing their thermal flux to be measured in the infrared. The first detections of an exoplanet's thermal emission (\\citealt{Charb05}; \\citealt{Deming05}) came from observations in space with Spitzer using the Infrared Array Camera (IRAC; \\citealt{Fazio04}). Since then the vast majority of such measurements have been made using Spitzer at wavelengths longer than 3 $\\mu m$, and thus longwards of the blackbody peak of these ``hot'' exoplanets. Recent observations have extended secondary eclipse detections into the near-infrared; the first detection was from space with NICMOS on the Hubble Space Telescope (\\citealt{Swain09} at $\\sim$2 $\\mu m$). More recently, near-infrared detections have been achieved from the ground; the first of these detections include a $\\sim$6$\\sigma$ detection in K-band of TrES-3b using the William Herschel Telescope \\citep{deMooij09}, a $\\sim$4$\\sigma$ detection in z'-band of OGLE-TR-56b using Magellan and the Very Large Telescope (VLT; \\citealt{Sing09}), and a $\\sim$5$\\sigma$ detection at $\\sim$2.1 $\\mu m$ of CoRoT-1b also with the VLT \\citep{Gillon09}. Thermal emission measurements in the near-infrared are crucial to our understanding of these planets' atmospheres, as they allow us to constrain hot Jupiters' thermal emission near their blackbody peaks. The combination of Spitzer/IRAC and near-infrared thermal emission measurements allows us to constrain the temperature-pressure profiles of these planets' atmospheres over a range of pressures \\citep{Fortney08}, better estimate the bolometric luminosity of these planets' dayside emission, and thus contributes to a more complete understanding of how these planets transport heat from the day to nightside at a variety of depths and pressures in their atmospheres \\citep{Barman08}. The transiting hot Jupiter TrES-2b orbits a G0 V star with a period of $\\sim$2.47 $d$ \\citep{ODonovan06}. According to the \\citet{Fortney08} theory this places TrES-2b marginally in the hottest, mostly highly irradiated class (the pM-class) of hot Jupiters and close to the dividing line between this hottest class and the merely warm class of hot Jupiters (the pL-class). Thus TrES-2b could be a key object to refine the dividing line between these two classes, and indicate the physical cause of this demarcation, or reveal whether this divide even exists. Recently \\citet{ODonovan09} used Spitzer/IRAC to measure the depth of the secondary eclipse of TrES-2b in the four IRAC bands. Their best-fit eclipses are consistent with a circular orbit, and collectively they are able to place a 3$\\sigma$ limit on the eccentricity, $e$, and argument of periastron, $\\omega$, of $|$$e$cos$\\omega$$|$ $<$ 0.0036. Their best-fit eclipses at 3.6, 5.8 and 8.0 $\\mu m$ are well-fit by a blackbody. At 4.5 $\\mu m$ they detect excess emission, in agreement with the theory of several researchers (\\citealt{Fortney08, Burrows08}) that predicts such excess due to water emission, rather than absorption, at this wavelength due to a temperature inversion in the atmosphere. One-dimensional radiative-equilibrium models for hot Jupiter planets generally show that the atmospheric opacity is dominated by water vapor, which is especially high in the mid-infrared, but has prominent windows (the JHK bands) in the near infrared \\citep{Fortney08,Burrows08}. One can probe more deeply, to gas at higher pressure, in these opacity windows. Models without temperature inversions feature strong emission in the JHK bands, since one sees down to the hotter gas. Models with temperature inversions, since they feature a relatively hotter upper atmosphere and relatively cooler lower atmosphere, yield weaker emission in the near-IR (JHK), but stronger emission in the mid-infrared \\citep{Hubeny03,Fortney06}. Near-infrared thermal emission measurements should thus be useful to determine whether TrES-2b does or does not harbour a temperature inversion. Owing to its high irradiation, with an incident flux of $\\sim$$1.1$$\\times$$10^{9}$ $erg$$s^{-1}$$cm^{-2}$, and favourable planet-to-star radius ratio ($R_{P}/R_{*}$$\\sim$0.13), we included TrES-2b in our program observing the secondary eclipses of some of the hottest of the hot Jupiters from the ground. Here we present Ks-band observations bracketing TrES-2b's secondary eclipse using the Wide-field InfraRed Camera (WIRCam) on the Canada-France-Hawaii Telescope (CFHT). We report a \\XSigmaTrESTwo $\\sigma$ detection of its thermal emission. ", "conclusions": "The depth of our best-fit secondary eclipse is \\FpOverFStarPercentAbstractTrESTwo$^{+\\FpOverFStarPercentAbstractPlusTrESTwo}_{-\\FpOverFStarPercentAbstractMinusTrESTwo}$\\%. The reduced $\\chi^{2}$ is \\ChiTrESTwo. Our best-fit secondary eclipse is consistent with a circular orbit; the offset from the expected eclipse center is: $t_{offset}$ = \\TOffsetTrESTwo$^{+\\TOffsetPlusTrESTwo}_{-\\TOffsetMinusTrESTwo}$ minutes (or at a phase of $\\phi$=\\PhaseAbstractTrESTwo$^{+\\PhaseAbstractPlusTrESTwo}_{-\\PhaseAbstractMinusTrESTwo}$). This corresponds to a limit on the eccentricity and argument of periastron of $e \\cos \\omega$ = \\ECosOmegaTrESTwo$^{+\\ECosOmegaPlusTrESTwo}_{-\\ECosOmegaMinusTrESTwo}$, or a 3$\\sigma$ limit of $|$$e$$\\cos$$\\omega$$|$ $<$ \\ECosOmegaAbsoluteThreeSigmaLimitTrESTwo). Our result is fully consistent with the more sensitive $e$cos$\\omega$ limits reported from the secondary eclipse detections at the four Spitzer/IRAC wavelengths \\citep{ODonovan09}. Thus our result bolsters the conclusion of \\citet{ODonovan09} that tidal damping of the orbital eccentricity is unlikely to be responsible for ``puffing up'' the radius of this exoplanet. A secondary eclipse depth of \\FpOverFStarPercentAbstractTrESTwo$^{+\\FpOverFStarPercentAbstractPlusTrESTwo}_{-\\FpOverFStarPercentAbstractMinusTrESTwo}$\\% corresponds to a brightness temperature of $T_{B}$ = \\TBrightTrESTwo$^{+\\TBrightPlusTrESTwo}_{-\\TBrightMinusTrESTwo}$ $K$ in the Ks-band assuming the planet radiates as a blackbody, and adopting a stellar effective temperature of $T_{eff}$ = 5850 $\\pm$ 50 \\citep{Sozzetti07}. This compares to the equilibrium temperature of TrES-2b of $T_{eq}$$\\sim$1472 $K$ assuming isotropic reradiation, and a zero Bond albedo. Hot Jupiter thermal emission measurements allow joint constraints on the Bond albedo, $A_B$, and the efficiency of day to nightside redistribution of heat on these presumably tidally locked planets. The Bond albedo, $A_B$ is the fraction of the bolometric, incident stellar irradiation that is reflected by the planet's atmosphere. We parameterize the redistribution of dayside stellar radiation absorbed by the planet's atmosphere to the nightside by the reradiation factor, $f$, following the \\citet{Lopez07} definition. If we assume a Bond albedo near zero, consistent with observations of other hot Jupiters \\citep{Charbonneau99,Rowe08} and with model predictions \\citep{Burrows08b}, we find a reradiation factor of $f_{Ks}$ = \\fReradiationTrESTwo$^{+\\fReradiationPlusTrESTwo}_{-\\fReradiationMinusTrESTwo}$ from our Ks-band eclipse photometry only, indicative of relatively efficient advection of heat from the day-to-nightside at this wavelength. In comparison, the reradiation factor for an atmosphere that reradiates isotropically is $f$=$\\frac{1}{4}$, while $f$=$\\frac{1}{2}$ denotes redistribution and reradiation over the dayside face only. \\begin{figure} \\centering \\includegraphics[scale=0.65,angle=270]{f8.eps} \\caption{\tThe 68.3\\% (1$\\sigma$; solid-line), 95.5\\% (2$\\sigma$; dashed-line) and 99.7\\% (3$\\sigma$; short dashed-line) $\\chi^2$ confidence regions on the reradiation factor, $f_{tot}$, and Bond albedo from the combination of our Ks-band point and the Spitzer/IRAC measurements \\citep{ODonovan09}.} \\label{FigBondReradiation} \\end{figure} Our secondary eclipse depth, when combined with the secondary eclipse depths at the Spitzer/IRAC wavelengths from \\citet{ODonovan09}, is consistent with a range of Bond albedos, $A_B$, and efficiencies of the day to nightside redistribution on this presumably tidally locked planet (Figure \\ref{FigBondReradiation}). The best-fit total reradiation factor, $f_{tot}$, that results from a $\\chi^{2}$ analysis of all the eclipse depths for TrES-2b assuming a zero Bond albedo is $f_{tot}$ = \\fReradiationTrESTwoALL$^{+\\fReradiationPlusTrESTwoALL}_{-\\fReradiationMinusTrESTwoALL}$. Thus our Ks-band brightness temperature ($T_{B}$ = \\TBrightTrESTwo$^{+\\TBrightPlusTrESTwo}_{-\\TBrightMinusTrESTwo}$ $K$) and reradiation factor $f_{Ks}$=\\fReradiationTrESTwo$^{+\\fReradiationPlusTrESTwo}_{-\\fReradiationMinusTrESTwo}$, reveal an atmospheric layer that is similar to, and perhaps slightly hotter, than the atmospheric layers probed by longer wavelength Spitzer observations ($T_B$$\\sim$1500 $K$ from Spitzer/IRAC observations of TrES-2b [\\citealt{ODonovan09}]). The Ks-band is expected to be at a minimum in the water opacity \\citep{Fortney08,Burrows08}, and thus our Ks-band observations are expected to be able to see deep into the atmosphere of TrES-2b. Our observations suggest that the deep, high pressure atmosphere of TrES-2b displays a similar temperature -- perhaps a slightly warmer temperature -- to lower pressure regions. Another way of parameterizing the level of day-to-nightside heat redistribution is calculating the percentage of the bolometric luminosity emitted by the planet's dayside, $L_{day}$, compared to the nightside emission, $L_{night}$. Measurements of the thermal emission of a hot Jupiter at its blackbody peak provide a valuable constraint on the bolometric luminosity of the planet's dayside emission, and by inference its nightside emission \\citep{Barman08}. From simple thermal equilibrium arguments if TrES-2b has a zero Bond albedo and it is in thermal equilibrium with its surroundings it should have a total bolometric luminosity of $L_{tot}$ = 7.7$\\times$10$^{-5}$$L_{\\odot}$. By integrating the luminosity per unit frequency of our best-fit blackbody model across a wide wavelength range we are able to calculate the percentage of the total luminosity reradiated by the dayside as $\\sim$\\DaysidePercentage\\% ($L_{day}$ = \\Lday$\\times$10$^{-5}$$L_{\\odot}$). The remainder, presumably, is advected via winds to the nightside. \\begin{figure} \\centering \\includegraphics[scale=0.465,angle=270]{f9a.eps} \\includegraphics[scale=0.465,angle=270]{f9b.eps} \\caption{ Dayside planet-to-star flux ratios (top) and dayside flux at the planet's surface (bottom). The Ks-band point (black point; $\\sim$2.15 $\\mu m$) is our own, while the Spitzer/IRAC red points are from \\citet{ODonovan09}. Blackbody curves for isotropic reradiation ($f$=$\\frac{1}{4}$; $T_{eq}$$\\sim$1496 $K$; blue dashed-line) and for our best-fit reradiation factor ($f$=0.346; $T_{eq}$$\\sim$1622 $K$; grey dotted-line) are also plotted. We also plot one-dimensional, radiative transfer spectral models \\citep{Fortney06,Fortney08} for various reradiation factors and with and without TiO/VO. The models with TiO/VO include $f$=$\\frac{1}{4}$ (purple dotted line), $f$=0.31 (green dashed line), and $f$ =$\\frac{1}{2}$ (orange dotted line); only the last of the models has a temperature inversion. The model without TiO/VO features emission from the dayside only ($f$=$\\frac{1}{2}$; cyan dot-dashed line). The models on the top panel are divided by a stellar atmosphere model \\citep{Hauschildt99} of TrES-2 using the parameters from \\citet{Torres08} ($M_{*}$=0.98 $M_{\\odot}$, $R_{*}$=1.00 $R_{\\odot}$, $T_{eff}$=5850 $K$, and log $g$= 4.43). We plot the Ks-band WIRCam transmission curve (dotted black lines) and Spitzer/IRAC curves (solid red lines) inverted at arbitrary scale at the top of both panels. } \\label{FigModel} \\end{figure} We compare the depth of our Ks-band eclipse and the Spitzer/IRAC eclipses \\citep{ODonovan09} to a series of planetary atmosphere models in Figure \\ref{FigModel}. This comparison is made quantitatively as well as qualitatively by integrating the models over the WIRCam Ks band-pass as well as the Spitzer/IRAC channels, and calculating the $\\chi^{2}$ of the thermal emission data compared to the models. We first plot blackbody models with an isotropic reradiation factor ($f$=$\\frac{1}{4}$; blue dotted-line) and that of our best-fit value ($f$=0.346; grey dotted-line) these models have dayside temperatures of $T_{day}$$\\sim$1496$K$ and $T_{day}$$\\sim$1622$K$, respectively. Both blackbody models provide reasonable fits to the data, although the latter model ($f$=0.346; $\\chi^{2}$=\\BlackbodyTwoChi) provides a definitively better fit than the former isothermal model ($f$=$\\frac{1}{4}$; $\\chi^2$=\\BlackbodyOneChi) as it better predicts our Ks-band emission and the Spitzer/IRAC 8.0 $\\mu m$ emission. This suggests that overall TrES-2b has a near-isothermal dayside temperature-pressure profile and is well-fit by a blackbody. We thus also compare the data to a number of one-dimensional, radiative transfer, spectral models \\citep{Fortney06,Fortney08} with different reradiation factors that specifically include or exclude gaseous TiO and VO into the chemical equilibrium and opacity calculations. In these models when TiO and VO are present they act as absorbers at high altitude and lead to a hot stratosphere and a temperature inversion \\citep{Hubeny03}. However, if the temperature becomes too cool (TiO and VO start to condense at 1670 $K$ at 1 mbar [\\citealt{Fortney08}]), TiO and VO condense out and the models with and without TiO/VO are very similar. In the case of TrES-2b, for all the models we calculate, except our model that features dayside emission only ($f$=$\\frac{1}{2}$), they do not harbour temperature inversions because the atmospheres are slightly too cool and TiO/VO has condensed out of their stratospheres. We plot models with TiO/VO and reradiation factors of $f$=$\\frac{1}{4}$ (purple dotted line), $f$=0.31 (green dashed line), and $f$=$\\frac{1}{2}$ (orange dotted line), and without TiO/VO with a reradiation factor of $f$=$\\frac{1}{2}$ (cyan dot-dashed line). \\citet{ODonovan09} argued that TrES-2b experienced a temperature inversion due to the high 4.5 $\\mu m$ emission compared to the low 3.6 $\\mu m$ emission, which was predicted to be a sign of water and CO in emission, rather than absorption, in TrES-2b's presumably inverted atmosphere. We also find that our models without a temperature inversion have difficultly matching the Spitzer/IRAC 5.6 and 8.0 $\\mu m$ thermal emission ($\\chi^{2}$=\\FortneyTwoChi \\ for $f$=$\\frac{1}{4}$ with TiO/VO, $\\chi^{2}$=\\FortneyOneChi \\ for $f$=0.31 with TiO/VO, and $\\chi^{2}$=\\FortneyThreeChi \\ for $f$=$\\frac{1}{2}$ without TiO/VO). If the temperature inversion is due to TiO/VO, by the time the atmosphere becomes hot enough that TiO/VO remains in gaseous form, the thermal emission is too bright to fit the 3.6, and 5.8 $\\mu m$ thermal emission ($\\chi^{2}$=\\FortneyFourChi \\ for $f$=$\\frac{1}{2}$ with TiO/VO). The combination of our blackbody and radiative transfer models with our own eclipse depth and those from the Spitzer/IRAC instrument \\citep{ODonovan09} thus suggest that the atmosphere of TrES-2b likely features modest redistribution of heat from the day to the nightside. It is unclear whether the atmosphere of TrES-2b requires a temperature inversion. A simple blackbody model ($f$=0.346 and $T_{eq}$$\\sim$1622 K) provides an exemplary fit to the data; this may indicate that TrES-2b has a fairly isothermal dayside temperature structure, perhaps similar to HAT-P-1b \\citep{Todorov10}. An important caveat, on the above result is that our $f$=$\\frac{1}{2}$ model without TiO/VO ($\\chi^{2}$=\\FortneyThreeChi) and thus without a temperature inversion returns nearly as good of fit as our best-fit blackbody model ($f$=0.346; $\\chi^{2}$=\\BlackbodyTwoChi); thermal emission measurements at other wavelengths, and repeated measurements at the above wavelengths, are thus necessary to differentiate a blackbody-like spectrum, from significant departures from blackbody-like behaviour, and to confirm that TrES-2b efficiently redistributes heat to the nightside of the planet. Specifically, the variations between the models displayed in Figure \\ref{FigModel} are largest in the near-infrared J \\& H-bands and thus further near-infrared constraints -- if they are able to achieve sufficient accuracy to measure the small thermal emission signal of TrES-2b in the near-infrared -- should prove eminently useful to constrain the atmospheric characteristics of this planet. If the excess emission at 4.5 $\\mu m$ is due to water emission, rather than absorption, due to a temperature inversion in the atmosphere of TrES-2b then the inversion is unlikely to be due to TiO/VO. This is because the atmosphere of TrES-2b appears too cool to allow TiO/VO to remain in gaseous form in its upper atmosphere. If there is a temperature inversion then the high altitude optical absorber is likely to be due to another chemical species than TiO/VO. For instance, \\citet{Zahnle09} have investigated the photochemistry of sulphur-bearing species as another alternative. TrES-2b is a promising target for the characterization of its thermal emission across a wide wavelength range. In addition to orbiting a relatively bright star, and having a favourable planet-to-star radius ratio, TrES-2 lies within the Kepler field. The combination of secondary eclipse measurements already published using Spitzer/IRAC, upcoming measurements with Kepler ($\\sim$430 - 900 $nm$; \\citealt{Borucki08}), and J, H and K-band near-infrared measurements that could be obtained from the ground, will allow us to fully constrain TrES-2b's energy budget. At the shorter end of this wavelength range it should also be possible to constrain the combination of reflected light and thermal emission. Our results predict that even if the geometric albedo of TrES-2b is as low as 5\\% in the Kepler bandpass, if Kepler is able to detect the secondary eclipse of this planet then it will be detecting a significant fraction of reflected light in addition to thermal emission. This will largely break the degeneracy on the Bond albedo and the reradiation factor for this planet, facilitating a more complete understanding of its energy budget." }, "1005/1005.1091_arXiv.txt": { "abstract": "We, first, analytically work out the long-term, i.e. averaged over one orbital revolution, perturbations on the orbit of a test particle moving in a local Fermi frame induced therein by the cosmological tidal effects of the inhomogeneous Lema\\^{\\i}tre-Tolman-Bondi (LTB) model. The LTB solution has recently attracted attention, among other things, as a possible explanation of the observed cosmic acceleration without resorting to dark energy. Then, we phenomenologically constrain both the parameters $K_1\\doteq -\\ddot{\\mathfrak{R}}/{\\mathfrak{R}}$ and $K_2\\doteq -\\ddot{\\mathfrak{R}}^{'}/{\\mathfrak{R}}^{'}$ of the LTB metric in the Fermi frame by using different kinds of solar system data. The corrections $\\Delta\\dot\\varpi$ to the standard Newtonian/Einsteinian precessions of the perihelia of the inner planets recently estimated with the EPM ephemerides, compared to our predictions for them, yield $K_1 = (4\\pm 8)\\times 10^{-26}$ s$^{-2}$, $K_2 = (3\\pm 7)\\times 10^{-23}$ s$^{-2}$. The residuals of the Cassini-based Earth-Saturn range, compared with the numerically integrated LTB range signature, allow to obtain $K_1\\approx K_2\\approx 10^{-27}$ s$^{-2}$. The LTB-induced distortions of the orbit of a typical object of the Oort cloud with respect to the commonly accepted Newtonian picture, based on the observations of the comet showers from that remote region of the solar system, point towards $K_1\\approx K_2\\lesssim 10^{-30}-10^{-32}$ s$^{-2}$. Such figures have to be compared with those inferred from cosmological data which are of the order of $K_1\\approx K_2=-4\\times 10^{-36}$ s$^{-2}$. ", "introduction": "Inhomogeneous cosmological models \\citep{Hellaby} have recently attracted much attention because, among other things, some of them may potentially be useful in explaining the observed Universe's acceleration without resorting to dark energy. Thus, it is important to put them on the test independently of the phenomenon itself for which they have been purposely introduced. In the framework of the general theory of relativity, the most general form of the line element $(ds)^2$ of a spherically symmetric inhomogeneous space-time in which the source in the Einstein's field equations is a perfect fluid is, in\\footnote{Recall that, by definition, the comoving coordinates are those in which the vector field $u^{\\alpha},\\ \\alpha=0,1,2,3$ has only the time component, i.e. $u^{\\alpha}\\propto \\delta^{\\alpha}_{\\ 0}$. If the vector field $u^{\\alpha}$ has zero rotation, as it happens if the space-time is spherically symmetric and the metric obeys the Einstein equations with a perfect fluid source, the comoving coordinates can be chosen so that they are synchronous, i.e. in them the metric tensor has no off-diagonal components $g_{0i},\\ i=1,2,3$.} comoving-synchronous spherical coordinates $\\{t,r,\\theta,\\phi\\}$ \\citep{Krasi}, {{\\eqi (ds)^2 = e^{{\\mathfrak{C}}(r,t)}(cdt)^2-e^{\\mathfrak{A}(r,t)}(dr)^2 - {\\mathfrak{R}}^2(r,t)[(d\\theta)^2+\\sin^2\\theta (d\\phi)^2].\\lb{metrica}\\eqf}} In order to solve the resulting Einstein field equations, an equation of state has to be assumed. The most natural choice consists of setting the pressure equal to zero, corresponding to a dust evolution driven by gravitation only. From the equations of motion of a perfect fluid it turns out that, for $p=0$, the dust moves along timelike geodesics. As a consequence, $\\mathfrak{C}=0$ and, for\\footnote{Here and in the following, $\\mathfrak{R}^{'}$ denotes the partial derivative of $\\mathfrak{R}$ with respect to $r$.} ${\\mathfrak{R}}^{'}\\neq 0$, \\eqi e^{\\mathfrak{A}}=\\rp{ {{\\mathfrak{R}}^{'}}^2 }{1+2\\mathfrak{E}(r)},\\eqf where $\\mathfrak{E}(r)$ is an arbitrary function such that $\\mathfrak{E}\\geq -1/2$ for all $r$ in order to preserve the right signature of the metric of \\rfr{metrica}. The resulting space-time line element is the so-called Lema\\^{\\i}tre-Tolman-Bondi (LTB) model. Indeed, such spherically symmetric inhomogeneous dust models were first discovered by \\citet{Lem} and further studied by \\citet{Tol} and \\citet{Bon}; for a detailed discussion of several properties see \\citet{Krasi}. Such models are among the best known and most useful exact solutions of Einstein's equations. Since they allow us to examine non-linear effects analytically, or at least in a tractable way, there is an extensive literature using them mostly to describe cosmological inhomogeneities \\citep{Hella}, but also as in other theoretical contexts, such as gravitational collapse \\citep{colla} and censorship of singularities \\citep{Cens} or quantum gravity \\citep{quantum}. Several authors considered LTB models as tools to probe how the cosmic acceleration associated to recent observations can be accounted for inhomogeneities, without introducing dark energy \\citep{Paran,Sark,Enq,Bellido}. LTB models are also a standard choice to apply Buchert's scalar averaging formalism \\citep{avera}, in which the effects of dark energy could be mimicked by \\virg{backreaction} terms. See \\citet{Noelle} for a comprehensive review of such aspects. Here we are interested in putting constraints in a phenomenological way on some parameters of the LTB metric from its local tidal effects on the dynamics of test bodies of the solar system. To this aim, it is necessary to set up a standard orthonormal tetrad frame parallel transported along the worldline of a fundamental observer representing, in this case, the Sun, and explicitly derive the expression of the LTB metric in such a local Fermi frame. This has been recently done by \\citet{Mash07}. In general, the motion of test bodies can be studied in such Fermi coordinate systems following ideas and methods developed by \\citet{Syn,Mana,Mash}. The paper is organized as follows. In Section \\ref{dua} we analytically work out the LTB effects on the orbital motion of a test particle in the local Fermi frame and compare them with some results present in literature. In Section \\ref{tria} we compare our theoretical predictions to different types of bodies and data of the solar system to phenomenologically put constraints on \\citep{Mash07} \\begin{eqnarray} K_1 &\\doteq & -\\rp{\\ddot{\\mathfrak{R}}}{\\mathfrak{R}}, \\\\ \\nonumber \\\\ K_2 &\\doteq & -\\rp{\\ddot{\\mathfrak{R}}^{'}}{\\mathfrak{R}^{'}} \\end{eqnarray} entering the LTB metric written in terms of the Fermi coordinates of a fundamental observer. Concerning their values infered from cosmological data, from (62)-(64) by \\citet{Mash07} it turns out that, in proximity of the origin of the local Fermi frame, i.e. for $r\\rightarrow 0$, $K_1$ and $K_2$ are rather similar being of the order of\\footnote{$K_{1/2}$ means that both $K_1$ and $K_2$ have approximately the same values.} \\eqi K_{1/2}\\approx q_0 H_0^2.\\eqf In it \\citep{hubblo} \\eqi H_0=71.0 \\ {\\rm km s}^{-1}\\ {\\rm Mpc}^{-1}=2.3\\times 10^{-18}\\ {\\rm s}^{-1}\\eqf is the present value of the Hubble parameter, and \\citep{Xu} \\eqi q_0 \\doteq\\rp{1+3\\widetilde{\\Omega}_{\\rm DE}w_{\\rm DE}}{2}=-0.7\\eqf is the deceleration parameter in which \\citep{hubblo} $\\widetilde{\\Omega}_{\\rm DE}=0.734\\pm 0.029$ is the dark energy density and \\citep{hubblo} $w_{\\rm DE}=-1.12^{+0.42}_{-0.43}$ is the equation of state. Thus, the order of magnitude of $K_1$ and $K_2$ inferred from cosmological observations is \\eqi K_{1/2}\\approx -4\\times 10^{-36}\\ {\\rm s}^{-2}.\\eqf Section \\ref{quatra} is devoted to summarizing our findings and to the conclusions. ", "conclusions": "\\lb{quatra} We, first, analytically worked out the long-term, i.e. averaged over one orbital revolution, perturbations on the orbit of a test particle moving in a local Fermi frame induced by the cosmological tidal effects of the inhomogeneous Lema\\^{\\i}tre-Tolman-Bondi model which has recently attracted attention as possible explanation of the observed cosmic acceleration without resorting to dark energy. In particular, we computed the variations of the semimajor axis $a$, the eccentricity $e$, the inclination $I$, the longitude of the ascending node $\\Omega$, the argumet of pericenter $\\omega$ and the mean anomaly $\\mathcal{M}$ by means of the Lagrange's variational equations with the eccentric anomaly $E$ as fast variable. We found that $a$ does not experience any change, on average, while the other Keplerian orbital elements undergo long-term variations including both secular and harmonic terms with frequency $2\\omega$; $e$ and $I$ exhibit only sinusoidal changes, while $I,\\Omega,\\mathcal{M}$ secularly precess as well. We repeated the calculation also with the Gauss perturbative approach by finding the same results. We also explicitly computed the changes over one orbital revolution of the position and velocity vectors along the radial $\\bds{\\hat R}$, transverse $\\bds{\\hat \\Xi}$ and normal directions $\\bds{\\hat \\Upsilon}$ of a frame co-moving with the test particle. While the radial and normal components $\\Delta R$ and $\\Delta\\Upsilon$ of the perturbation of the position vector experience only harmonic variations, the transverse one $\\Delta\\Xi$ shows secular changes as well. Concerning the velocity, its radial perturbation $\\Delta V_R$ undergoes both secular and sinusoidal modifications, while its transverse and normal components $\\Delta V_{\\Xi}$ and $\\Delta V_{\\Upsilon}$ exhibit harmonic signatures only. In obtaining our results we made no approximations on $e$ and $I$ by retaining all terms. Then, we phenomenologically put constraints on both the parameters $K_1$ and $K_2$ of the LTB metric in the local Fermi frame by looking at various astronomical bodies and data of the solar system. By comparing our analytical prediction for the rate of the longitude of pericenter $\\varpi$ to different sets of the corrections $\\Delta\\dot\\varpi$ to the standard Newtonian/Einsteinian precessions of the perihelia of the inner planets recently estimated with the EPM ephemerides we found that the tightest constraints are $K_1 = (4\\pm 8)\\times 10^{-26}$ s$^{-2}$, $K_2 = (3\\pm 7)\\times 10^{-23}$ s$^{-2}$. The confrontation of the residuals of the Earth-Saturn range, obtained by processing some years of radiotechnical data from the Cassini spacecraft as well, with the numerically computed LTB-induced Earth-Saturn range signal allowed to set $K_1\\approx K_2\\approx 10^{-27}$ s$^{-2}$. By looking at the LTB-type distortions of the orbit of a typical object of the Oort cloud with respect to the commonly accepted Newtonian picture, based on the observations of the comet showers from that remote region of the solar system, pointed towards even lower bounds, i.e. $K_1\\approx K_2\\lesssim 10^{-30}-10^{-32}$ s$^{-2}$. According to cosmological data, $K_1\\approx K_2=-4\\times 10^{-36}$ s$^{-2}$." }, "1005/1005.4999_arXiv.txt": { "abstract": "A time-dependent Synchrotron Self Compton model (SSC) which is able to motivate the used electron spectra of many SSC models as a balance of acceleration and radiative losses is introduced. Using stochastic acceleration as well as Fermi-I processes even electron spectra with a rising part can be explained, which are mandatory to fit the lowstate spectral energy distribution (SED) of PKS 2155-304 as constrained from Fermi LAT observations. Due to the time resolution the outburst of PKS 2155-304 observed by H.E.S.S. in 2006 can be modelled selfconsistently as fluctuations along the jet axis without introducing new sets of parameters. The model makes the time evolution of the SED also accessible. Hence giving new insights into the flaring behavior of blazars. ", "introduction": " ", "conclusions": "" }, "1005/1005.1572_arXiv.txt": { "abstract": "Using three epochs of VLA observations of the Galactic Plane in the first quadrant taken $\\sim15$ years apart, we have conducted a search for a population of variable Galactic radio emitters in the flux density range 1--100 mJy at 6~cm. We find 39 variable sources in a total survey area of 23.2 deg$^2$. Correcting for various selection effects and for the extragalactic variable population of active galactic nuclei, we conclude there are $\\sim1.6$ Galactic sources deg$^{-2}$ which vary by more than 50\\% on a time scale of years (or shorter). We show that these sources are much more highly variable than extragalactic objects; more than 50\\% show variability by a factor $>2$ compared to $<10\\%$ for extragalactic objects in the same flux density range. We also show that the fraction of variable sources increases toward the Galactic center (another indication that this is a Galactic population), and that the spectral indices of many of these sources are flat or inverted. A small number of the variables are coincident with mid-IR sources and two are coincident with X-ray emitters, but most have no known counterparts at other wavelengths. Intriguingly, one lies at the center of a supernova remnant, while another appears to be a very compact planetary nebula; several are likely to represent activity associated with star formation regions. We discuss the possible source classes which could contribute to the variable cohort and followup observations which could clarify the nature of these sources. ", "introduction": "Variable radio emission is a hallmark of energetic objects such as coronally active stars, supernovae, neutron stars, black holes, and active galactic nuclei (AGN). Indeed, radio variability is often indicative of high-energy processes and, in principle, can be valuable for finding examples of relatively rare objects. However, surveys for variability are themselves quite rare --- blind sky surveys are almost never repeated owing to the scarcity of telescope time. The exceptions are mostly in the optical regime. Comparisons between POSS1, POSS2, and SDSS have been useful for studying variability (de~Vries et al.\\ 2005). Gravitational microlensing studies (e.g., Alcock et al.\\ 1997) and supernova searches (e.g., Astier et al.\\ 2006, Miknaitis et al.\\ 2007) have produced a wealth of data on optical variability from targeted sky regions, and several upcoming experiments such as Pan-STARRS (Kaiser et al.\\ 2002), the Palomar Transit Factory (Rau et al.\\ 2009), and LSST (Tyson 2002) will make the coming decade one in which time-domain astronomy plays a prominent role. Variability studies in the radio band have typically targeted bright extragalactic sources (see de~Vries et al.\\ 2004 for a review of searches for, and mechanisms of, radio variability). Comparisons between blind radio surveys are often hampered by differences in angular resolution and the confusing presence of interferometric sidelobe patterns. For example, there has been no systematic search for radio variability between the two largest radio sky surveys, FIRST (Becker et al.\\ 1995) and NVSS (Condon et al.\\ 1998). The FIRST survey did observe one area twice at 1400 MHz, an equatorial strip $\\sim1.5$ degrees wide in the range $21^{\\mathrm h}20^{\\mathrm m} < RA < 03^{\\mathrm h} 20^{\\mathrm m}$. A search for variable sources in this area was reported in de~Vries et al.\\ (2004). The search covered $\\sim120$~deg$^2$ of extragalactic sky with a sensitivity similar to the Galactic plane search reported here; it thus serves as a useful control from which to estimate how many of the sources we find are background extragalactic radio sources. The most systematic search for radio variablity in the Galactic plane used the NRAO 91-m telescope in Greenbank, WV, operating at a frequency of 5~GHz (Gregory \\& Taylor 1986). Over a five-year period the plane was observed 16 times, leading to the detection of 32 variable radio sources. The survey had a flux density threshold of $\\sim20$~mJy and an angular resolution of 3\\arcmin. Using the Very Large Array\\footnote{The Very Large Array is an instrument of the National Radio Astronomy Observatory, a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.} (VLA), the Galactic plane has been surveyed at this same frequency but with much higher sensitivity and angular resolution, although with minimal repetition (Becker et al.\\ 1994). Recently, a new Galactic plane survey at 6~cm (4.86~GHz) has begun at the VLA. The new survey (CORNISH\\footnote{The Co-Ordinated Radio 'N' Infrared Survey for High-mass star formation; see \\url{http://www.ast.leeds.ac.uk/Cornish}}; Purcell et al.\\ 2008) has substantial overlap with our previous survey; both surveys have a flux density threshold of $\\sim1$~mJy. Comparison between these two data sets allows for a search for Galactic radio sources that exhibit variability over the fifteen-year interval between the surveys. The search is complicated because the two surveys use different VLA configurations and hence have different angular resolutions (5\\arcsec\\ versus 1.5\\arcsec). Nonetheless, it is possible to identify strongly varying sources. In this paper we will compare results between the original survey and two epochs of data from the new survey. In section 2 we discuss the parameters of the two surveys, while in section 3 we present the results from a comparison of the two samples and adduce evidence for variability. We describe the properties of the variable sources including their spatial distribution, spectral indices, and counterparts at other wavelengths (\\S4) and end with a discussion of our limited knowledge of the nature of these objects (\\S5). \\begin{figure*} \\epsscale{1.0} \\plotone{f1.eps} \\caption{ Sky coverage for overlapping regions from the 1990${+}$ (red), 2005 (blue), and 2006 (gray) 6~cm survey epochs. Darker regions have higher rms noise values, while white areas are outside the survey. The noise is higher at the edges of the coverage and in fields with bright or complex extended sources. Typical rms values are $\\sim 0.1$~mJy in both surveys, but the old 6~cm data were acquired with a more widely spaced pointing grid and so display greater variation with position. The areas of overlap are given in Table~1. } \\label{fig-coverage} \\end{figure*} ", "conclusions": "\\subsection{Source Identification -- What They Are Not} Having established the existence of a population of highly variable Galactic sources, the obvious question is, What are they? Three classes of Galactic variable radio emitters can be easily eliminated from consideration: coronally active radio-emitting stars, pulsars, and masers. We justify our exclusion of these source classes in turn. In a survey of 122 RS CVn and related active binary systems which are among the most luminous stellar radio sources, Drake et al. (1989) found only 18 detected above a quiescent flux density of 1~mJy at 6~cm; the faintest optical counterpart was $V=10.0$. Even assuming an extreme flare of a factor of 100 (Osten 2008), the faintest possible counterpart would have $V=15$; none of our variables has a counterpart this bright. As for dMe flare stars, the other main class of variable stellar radio emitters, the most luminous quiescent emission is $\\sim10^{14.2}$ erg s$^{-1}$ Hz$^{-1}$ (Gudel et al.\\ 1993) corresponding to a flux density of $\\sim1$~mJy at a distance of 13~pc. Even an extreme flare with an increase of a factor of 500 over the quiescent level (Osten 2008 and references therein) would fall below our flux density threshold for a distance $>290$~pc. Stars with spectral types later than M6 would have counterparts fainter than 20th magnitude and could be represented in our sample. However, statistically, M-stars cannot be a significant contributor; Helfand et al.\\ (1999) found only $\\sim5$ M stars in 5000 deg$^2$ of the {\\sl FIRST} survey to a flux density limit of 0.7~mJy, whereas our variables have a surface density of 1.6~deg$^{-2}$. While nearby radio pulsars scintillate strongly in the ISM leading to large-amplitude variability, pulsars have very steep radio spectra, and most have not been detected at 6~cm (none of our objects are coincident with one of the 1827 known pulsars; Manchester et al. 2005). For a typical spectral index of $-1.5$, our weakest source would be a $\\sim100$~mJy pulsar at 400~MHz, and most unlikely to have been missed in pulsar surveys. The small duty cycle of the recently discovered RRATs (Rotating RAdio Transients -- McLaughlin et al.\\ 2006) makes them equally unlikely to explain our variable sources. Finally, radio masers are known to be highly variable, but no known maser transitions fall within our bandpass. As noted below, however, three of our variables are coincident with methanol masers. Two classes of extragalactic radio transients --- supernovae and GRB afterglows --- are also highly improbable counterparts for our events. Both have rise times of at most tens of days and cannot, in the absence of a steady underlying source of radio emission, account for the bulk of our sources which show a flux density increase over many years. In addition, their rarity makes them statistically unlikely counterparts. The one extragalactic population that does show variability on the time scales we probe, AGN, are shown above to have variability amplitudes which exclude them from explaining all but a handful of our events. The remaining known classes of variable radio sources include microquasars (accreting, high-mass X-ray binaries that produce relativistic jets: e.g., SS433, Cyg X-3 and GRS 1915${+}$105), radio magnetars (Camilo et al.\\ 2006; Camilo et al.\\ 2007), and the recently described Galactic Center Transient sources (Hyman et al.\\ 2009 and references therein). The first two of these have signatures at other wavelengths; we explore below the fragmentary data outside the radio band that is available for our variable objects. \\subsection{Source Identification -- Multiwavelength Data} Counterparts at other wavelengths can be useful in suggesting the origin of radio variability. At our MAGPIS website (Helfand et al. 2006), we have collected the following Galactic plane data in addition to the three-epoch 6~cm data described herein: two epochs of 20~cm observations for these same fields including the principal MAGPIS survey, 90~cm observations of the same regions, the 3.6, 4.5, 5.8, and 8.0 $\\mu$m data from the Spitzer {\\sl GLIMPSE} survey (Benjamin et al.\\ 2003), 24 $\\mu$m images from MIPSGAL (Carey et al.\\ 2009), 20 $\\mu$m data from the MSX survey (Price et al.\\ 2001) and the 1.1~mm Bolocam Galactic plane survey (Aguirre et al.\\ 2009). In addition, we have queried the SIMBAD database for each of our sources and have examined the Digitized Sky Survey images; in one case, we have obtained optical observations of a source. We report the results of this multi-wavelength inquiry here. \\subsubsection{Mid-IR and mm observations} Seven of our variables are detected at 24 $\\mu$m in the MIPSGAL survey, and six of these are also detected in at least one GLIMPSE mid-IR band. Four of the objects are also detected at 1.1~mm in the Bolocam survey. In all seven cases at least two bands are available, and in all seven cases the sources are red; i.e., they are faintest in the short-wavelength bands and brightest in the long-wavelength bands. In two cases (G31.1595${+}$0.0449 and G37.7347${-}$0.1126) multiple components with different IR spectral shapes are present, with the radio source identified with the brighter component in the first case, and the redder component in the second. Three of the IR-detected objects have associated methanol masers; this, coupled with their IR spectra demonstrate they represent activity associated with star formation in compact or ultracompact \\ion{H}{2} regions. For one IR-detected source, G29.5779${-}$0.2685, we have obtained followup observations at the MDM Observatory (J.~Halpern, private communication). R-band and H$\\alpha$ images were obtained on 23~August 2009, and show a barely resolved ($\\sim1\\arcsec$) object, brighter in H$\\alpha$ and coincident with the radio source. A spectrum obtained the same night shows no continuum, but very strong nebular emission lines. The object appears to be a very compact planetary nebula. Its radio flux history is thus perplexing: 6.9~mJy at 6~cm in $\\sim1990$, rising to 10.5~mJy in 2005 and falling again to 5.8~mJy in 2006. The 20~cm flux density in the MAGPIS survey (epoch 2001--2004) is only 1.3~mJy, suggesting the source may be optically thick. Further simultaneous multi-frequency observations are required to measure the radio spectrum and derive clues as to the nature of the source's variability. \\begin{figure*} \\epsscale{1.0} \\plotone{f7.eps} \\caption{ MAGPIS 20~cm image of supernova remnants W41 (G23.3${-}$0.3) and G22.7${-}$0.2. The boxes mark the positions of three variable 6~cm sources (G22.7194${-}$0.1939, G22.9116${-}$0.2878, and G22.9743${-}$0.3920). } \\label{fig-snrs} \\end{figure*} \\subsubsection{X-ray observations} The brightest variable, G21.6552${-}$0.3611, is coincident with a point-like X-ray source catalogued in the XMM Galactic Plane survey (Hands et al.\\ 2004). It has a hard band (2--6~keV) flux of 0.0051 ct s$^{-1}$ and is undetected in the soft (0.4--2.0~keV) band. For an intrinsic power-law spectrum with spectral index $\\Gamma = 1.9$, the expected absorption column density through the Galactic plane of $\\sim10^{23}$ cm$^{-2}$ is consistent with the non-detection in the soft band; the inferred intrinsic flux in the 0.2--10~keV band would be $7 \\times 10^{-13}$ erg cm$^{-2}$ s$^{-1}$. For an extragalactic AGN at 1 Gpc, this corresponds to a luminosity of $8 \\times 10^{42}$ erg s$^{-1}$, while for a Galactic object at 5~kpc, the X-ray luminosity would be a modest $2 \\times 10^{33}$ erg s$^{-1}$; for a column density of only $10^{22}$ cm$^{-2}$, the luminosity estimates are lower by a factor of 3. While this source is the brightest of our variables, it has one of the lowest modulation factors (decreasing by just $\\sim50\\%$ over 16 years). The (non-contemporaneous) 20~cm flux density is lower than either of the 6~cm values, suggesting a mildly inverted spectrum source. It is not detected at any other wavelength. The most likely explanation of this object is a flat-spectrum extragalactic radio source, one of a handful we expect in our sample. One other source, G30.4460${-}$0.2148, lies 27\\arcsec\\ from the position of an ASCA Galactic Plane Survey catalog entry (Sugizaki et al.\\ 2001). The uncertainty in the X-ray position is 1\\arcmin; one other (brighter) radio source lies within the X-ray error circle although at twice the distance from its centroid. The X-ray source is a marginal detection ($4.6 \\sigma$) with a 0.7--7.0~keV unabsorbed flux of $2.6 \\times 10^{-12}$ erg cm$^{-2}$ s$^{-1}$ for an intrinsic power law index of $\\Gamma = 1.9$ and an absorption column density of $10^{23}$ cm$^{-2}$; the flux is roughly four times lower for $N_H = 10^{22}$ cm$^{-2}$. Assuming the identification is correct, the X-ray to radio flux ratio is thus $\\sim20$ times greater than our other X-ray detection, although still within the X-ray to radio luminosity ratios characteristic of AGN. The primary distinguishing feature, however, is that the radio source is coincident with a very bright mid-IR source (saturated in all but the 3.6 $\\mu$m band) which is also detected at 1.1~mm. The ASCA Galactic Plane Survey covered an area encompassing all but six of our variables to a flux density level of approximately $10^{-12.5}$ erg cm$^{-2}$ s$^{-1}$; no other X-ray sources are coincident to within 1\\arcmin. The higher-resolution coverage of the Einstein, ROSAT, XMM, and Chandra is much spottier; no further matches are found within the 10\\arcsec\\ error circles of these other catalogs. \\subsubsection{Low-frequency radio detections} Three of the variable sources are detected at 90~cm. G22.9116${-}$0.2878 (Fig.~\\ref{fig-snrs}) has a 90~cm flux density of $\\sim180$~mJy; this is consistent with a nonthermal spectral index of $\\sim-0.9$ if one takes the most recent (but far from contemporaneous) 6, 20, and 90~cm measurements. The 6~cm flux density increased by more than a factor of three since 1990, making it one of the higher amplitude variables, but no other information is available on this source. G30.6724${+}$0.9637 (the highest latitude source detected) has a 90~cm flux density of $\\sim70$ mJy, below that of the 20~cm flux density (90~mJy), possible additional evidence for variability, as the 20:6~cm flux density ratio is 3:1 (again, all non-contemporaneous). This is the smallest amplitude variable in our sample and, given its distance from the Galactic plane, an extragalactic counterpart is the most likely explanation. The third 90~cm detection is G22.7194${-}$0.1939, perhaps the most intriguing source in our sample. An image of the region surrounding this source is given in Figure~\\ref{fig-snrs}. The source lies very near to the geometric center of a 30\\arcmin-diameter supernova remnant, G22.7${-}$0.2 (Green 2004 and references therein) and 4\\arcmin\\ from a fairly bright \\ion{H}{2} region. There is no counterpart detected at mm, IR, or optical wavelengths. The source brightened by a factor of four between 2003 and 2006 at 6~cm; its 20~cm flux density is 12~mJy, three times higher than the higher of the two 6~cm measurements. The distance to the remnant is unknown, although its large angular diameter would suggest it is not very remote (its diameter would be $\\sim45$~pc at 5~kpc). X-ray observations could reveal whether or not this source is likely to be a compact object associated with the supernova remnant. \\subsection{Summary} We have discovered a relatively high surface density (2 deg$^{-2}$) of variable radio sources in the Galactic plane and have argued that the large majority of these ($\\sim80$\\%) are Galactic objects. While a few are associated with young star formation activity, the identity of the majority is unknown. Follow up radio observations are required to confirm the variability in these sources, establish the variability time scale(s), and obtain contemporaneous spectral indices. Observations at optical, infrared, and X-ray wavelengths could help establish counterparts and identify the origin of the variable radio emission." }, "1005/1005.4161_arXiv.txt": { "abstract": "The \\r-mode instability in rotating compact stars is used to constrain the phase of matter at high density. The color-flavor-locked phase with kaon condensation (CFL-K0) and without (CFL) is considered in the temperature range $10^8$K $\\lesssim T\\lesssim 10^{11}$K. While the bulk viscosity in either phase is only effective at damping the \\r-mode at temperatures $T\\gtrsim10^{11}$K, the shear viscosity in the CFL-K0 phase is the only effective damping agent all the way down to temperatures $T\\gtrsim 10^8$K characteristic of cooling neutron stars. However, it cannot keep the star from becoming unstable to gravitational wave emission for rotation frequencies $\\nu\\approx 56-11$Hz at $T\\approx10^8-10^9$K. Stars composed almost entirely of CFL or CFL-K0 matter are ruled out by observation of rapidly rotating neutron stars, indicating that dissipation at the quark-hadron interface or nuclear crust interface must play a key role in damping the instability. ", "introduction": "Introduction} The $r$-mode instability of rotating compact stars is of wide astrophysical interest for several reasons~\\cite{Andersson:1997xt,*Friedman:1997uh,*Lindblom:1998wf}. The majority of neutron stars have rotation period $P\\sim 0.5$seconds, much slower than theoretical limits: $P\\sim 0.001$s ($\\nu=1/P\\sim 1$kHz)~\\cite{Shapiro:1984}. Even the fastest spinning neutron star PSR J1748-2446ad with $\\nu=716$Hz~\\cite{Hessels:2006} is not near the theoretical stability limit, with the majority of ``spun-up'' neutron stars in low mass X-ray binaries (LMXBs) catalogued between 300-640 Hz. The \\r-mode provides a mechanism for spinning down young neutron stars~\\cite{Andersson:1998qs} and limiting spin-up frequencies in millisecond pulsars (see eg.~\\cite{Lorimer:2008se} for a review) through angular momentum loss in gravitational wave emissions. These gravitational waves may be detected given the expected improved sensitivities of ground-based interferometers eg. VIRGO, advanced LIGO~\\cite{Sa:2007zz,*Watts:2008qw,*Stergioulas:lrr-2003-3}. In addition, \\r-modes are probes of the equation of state of cold and dense matter through viscous damping effects and advance our understanding of nuclear and quark interactions by providing an astrophysical and phenomenological link to Quantum Chromodynamics (QCD) over a certain range of temperatures ($10^7$K $\\lesssim T\\lesssim10^{11}$K) and at super-saturation densities [$n_B\\sim$ (2-5)$n_{\\rm sat}$]. Only a few studies of \\r-modes linking them to phenomenology of quark matter in neutron stars have been carried out thus far~\\cite{Madsen:1999ci,Andersson:2001ev,Jaikumar:2008kh,Sa'd:2008gf,Mannarelli:2008je}. In this work we use \\r-mode damping to constrain the presence of a realistic phase of superconducting quark matter~\\cite{Alford:1998mk,Rapp:1997zu} in the core of neutron stars: viz., one that includes the effects of a sizable strange quark mass $m_s$ at intermediate baryon density. A key parameter in understanding the \\r-mode instability in rotating compact stars is the critical frequency $\\Omega_c$ for the onset of the instability ($\\Omega=2\\pi/P$~rad/s). Gravitational radiation tends to make the $r$-mode grow on a characteristic time scale $\\tau_{GR}$. Internal friction such as viscosity or mutual friction tends to damp $r$-mode growth on a characteristic time scale $\\tau_F$. The competition between the two determines $\\Omega_c$ at any temperature. At $\\Omega\\geq\\Omega_c$, the $r$-mode develops an instability ($1/\\tau_{\\rm total}=1/\\tau_{GR}+1/\\tau_{F}<0$). Previous works~\\cite{Sa'd:2008gf,Jaikumar:2008kh,Mannarelli:2008je} have studied the \\r-mode damping in normal (ungapped) quark matter as well as some superconducting (gapped) phases such as the color-flavor-locked phase (CFL), and found that bulk and shear viscosity in the fully gapped phase is large enough to render rapidly spinning quark stars ($\\Omega\\geq 0.1\\Omega_\\mathrm{Kepler}$) stable. However, this conclusion applies only in a restricted temperature regime $T\\gtrsim 10^{10}$K, since the shearing mean free path due to phonons otherwise becomes larger than the star's radius ($\\sim 10$km). Furthermore, for temperatures $T\\lesssim 10^8$K, mutual friction in the CFL phase has been shown to be too weak to damp \\r-modes~\\cite{Mannarelli:2008je}, arguing for rapid spin-down of CFL stars to less than a few Hz, effectively ruling our cold CFL matter in any rapidly rotating neutron stars. The \\r-mode damping in the temperature regime $10^8$K $\\lesssim T \\lesssim10^{10}$K for color-superconducting phases of quark matter is therefore an open question, and is the focus of this work. Here we study $r$-mode damping and the critical frequency curve for a neutron star made up mostly or entirely of the kaon condensed CFL (CFL-K0) phase. The CFL phase with symmetric pairing of up-, down- and strange-quarks is severely stressed at realistic chemical potential $\\mu_q\\sim 300$ MeV because the strange quark mass $m_s\\sim 100$ MeV is non-negligible and costs extra energy $m_s^2/(2\\mu_q)$ compared to the much lighter up- and down-quarks. In Ref.~\\cite{Bedaque:2001je}, it was shown that it is energetically favorable for the CFL vacuum to be in a rotated chiral state with a non-zero kaon condensate when the cost $m_s^2/(2\\mu_q)$ exceeds the kaon mass $m_K$, the lightest meson in high-density QCD~\\cite{Son:1999cm}. Realistic neutron star densities could support the CFL-K0 phase. We find that in the CFL-K0 phase, while the bulk viscosity has little role to play in the temperature region relevant to \\r-mode damping in cooling neutron stars, the shear viscosity from kaons in the CFL-K0 phase is important even below $T\\lesssim 10^9$K, unlike the CFL phase. In essence, this fact coupled with the large mean free path of the phonons controls the main features of the critical frequency curve (Fig.~\\ref{fig:critfreq}). ", "conclusions": "Conclusions} \\r-mode oscillations of compact stars in the kaon-condensed CFL phase are considered. The mode frequency is almost exactly the same as that for a CFL star~\\cite{Jaikumar:2008kh} since the equation of state for the CFL-K0 phase only differs at ${\\cal O}(m_S^4)$. The mode growth and viscous damping timescales in the temperature range $10^8$K $\\lesssim T\\lesssim 10^{11}$K is studied based on the dominant bulk and shear viscosity contributions. Compared to the pure CFL phase, the bulk viscosity in CFL-K0 is smaller and does not play a significant role. At temperatures $T\\gtrsim 10^{10}$K shear viscosity associated with phonon scattering is important and determines the temperature dependence of the critical frequency curve. At lower temperatures, the phonon mean free path exceeds the star radius, so that the new mechanism for shear viscosity associated with kaon condensation~\\cite{Alford:2009jm} provides the dominant dissipation. In comparison to pure CFL stars, CFL-K0 stars are more stable against the \\r-mode instability at the lower end of the temperature range considered. However, given the observed LMXB spin rates, the critical frequency curve predicts that even pure CFL-K0 stars are unlikely to exist. Just as in the neutron matter case, viscous damping just beneath the core-crust interface (Ekman layer) appears to be necessary to provide a large enough stable rotation rate consistent with LMXB data~\\cite{Bildsten:1999zn,*Levin:2000vq}. The main result of the current work is that given the known dissipative mechanism in CFL or CFL-K0 phase, a pure quark star with these phases is ruled out by observed LMXB spin rates. Mutual friction associated with kaon-vortex scattering could provide an additional dissipative mechanism, but the critical frequency curves calculated here provide a lower bound. Future related work should investigate hybrid stars with a CFL-K0 core. The role of the crust-core interface as well as the quark-hadronic matter layer should be key ingredients in obtaining constraints on the persistence of gravitational waves from stars containing superfluid quark matter in their interior." }, "1005/1005.1791.txt": { "abstract": "We present a new set of cooling models and isochrones for both H- and He-atmosphere white dwarfs, incorporating accurate boundary conditions from detailed model atmosphere calculations, and carbon-oxygen chemical abundance profiles based on updated stellar evolution calculations from the BaSTI stellar evolution archive - a theoretical data center for the Virtual Observatory. We discuss and quantify the uncertainties in the cooling times predicted by the models, arising from the treatment of mixing during the central H- and He-burning phases, number of thermal pulses experienced by the progenitors, progenitor metallicity and the $^{12}C(\\alpha,\\gamma)^{16}O$ reaction rate. The largest sources of uncertainty turn out to be related to the treatment of convection during the last stages of the progenitor central He-burning phase, and the $^{12}C(\\alpha,\\gamma)^{16}O$ reaction rate. We compare our new models to previous calculations performed with the same stellar evolution code, and discuss their application to the estimate of the age of the solar neighborhood, and the interpretation of the observed number ratios between H- and He-atmosphere white dwarfs. The new white dwarf sequences and an extensive set of white dwarf isochrones that cover a large range of ages and progenitor metallicities are made publicly available at the official BaSTI website. ", "introduction": "\\label{intro} % %%%%%%%%%%%%%%%%%%%%%% The interpretation of photometric and spectroscopic observations of stellar populations relies on the use of grids of stellar models and isochrones, that have to cover a wide range of initial chemical compositions, stellar masses and evolutionary phases. The BaSTI (a Bag of Stellar Tracks and Isochrones) project\\footnote{Official website at \\url{http://www.oa-teramo.inaf.it/BASTI}} started in 2004 has delivered, to date, an homogeneous database of stellar evolution models, isochrones and integrated spectra for single-age, single-metallicity populations, encompassing a large chemical composition range appropriate for stellar populations harboured in star clusters and galaxies of various morphological types (Pietrinferni et al.~2004, 2006, 2009, Cordier et al.~2007, Percival et al.~2009). Results from BaSTI projects have been used by a large number of authors to address very diverse astrophysical problems like, among others, fitting eclipsing binary systems in the mass-radius plane, determining the ages of star clusters from their color-magnitude-diagrams (CMDs) or comparing integrated colors of elliptical galaxies with theoretical predictions. BaSTI models and isochrones in their present form cover all relevant evolutionary phases until either the end of the thermal pulse regime along the Asymptotic Giant Branch (AGB), or central carbon ignition for masses without electron degenerate carbon-oxygen (CO) cores. In this paper we extend the evolutionary phase coverage of our database to include cooling models of CO-core White Dwarfs (WDs), the final evolutionary phase of stars with initial masses smaller than about 6-7 $M_{\\odot}$. %Typical CO-core WDs (hereafter denoted simply as WDs) have masses between $\\sim$0.55 and %$\\sim 1.0~M_{\\odot}$ and are made of a nearly %isothermal electron degenerate core surrounded by a layer of pure He with mass of the order %of $M_{He} \\sim 10^{-2} \\ M_{WD}$ (where $M_{WD}$ is the total mass of the WD) %or less. This He-layer can be, in turn , surrounded by a H-layer with mass %of the order of $M_H \\sim 10^{-4} \\ M_{WD}$ or less. %In absence of metal accretion from the environment, given the high efficiency of atomic diffusion %at the very beginning of the cooling sequence (see.e.g., Koester~2009 for %a tabulation of metal diffusion timescales in WD envelopes, as a function of $T_{eff}$) all metals in the %progenitor's envelope have settled at the bottom of the He-layer. %The WD evolution is a cooling process, whereby the core acts as energy reservoir (stored as %the internal energy of the CO ions) while the outer non-degenerate layers control the rate of energy outflow. As a result, %the luminosity and core temperature decrease with time, at approximately constant radius. %Given that most stars are or will become WDs, plus the existence of %a well defined relationship between cooling time and luminosity, and the large cooling timescales, WDs have been since long %recognized as very attractive candidates to unveil the history of star formation in the Galaxy (e.g., Schmidt~1959). During the last two decades observations and theory have improved to a level that has made finally possible to employ WDs for determining ages of the stellar populations in the solar neighborhood (e.g., Winget et al.~1987, Garcia-Berro et al.~1988, Wood 1992, Oswalt et al.~1996), and in the nearest open (e.g., Richer et al.~1998, von Hippel~2005, Bedin et al.~2008, 2010) and globular (e.g. Hansen et al. 2004, 2007, Bedin et al.~2009) clusters. Methods to determine stellar population ages from their WD cooling sequences are usually based on the comparison of either the observed WD luminosity function (LF - star counts as a function of magnitude, e.g.,Winget et al.~1987, Bedin et al.~2010) or the actual bidimensional WD distribution in the CMD, with their theoretical counterparts (see, e.g., Hansen et al.~2007). Both techniques rely on an extensive use of grids of WD cooling sequences. The sets of WD models largely employed in the more recent investigations on the age of WDs in Galactic stellar populations are those by Hansen~(1999, hereafter H99) and Salaris et al.~(2000, hereafter S00), computed with completely independent evolutionary codes and largely independent input physics. Additional recent large sets of WD evolutionary cooling models can be found in Althaus \\& Benvenuto~(1998 -- and later updates from the same group, who has also produced extensive libraries of He-core WD models, presented in Serenelli et al.~2002 and Althaus et al.~2009) and Fontaine, Brassard \\& Bergeron~(2001). The new grid of cooling models we present here to extend the evolutionary phase coverage of BaSTI, is an update of the results by S00. We include a complete set of both H- and He-atmosphere WD models (and isochrones, for a range of progenitor initial chemical compositions) that take advantage of the updated CO stratifications obtained from BaSTI AGB models, and employ boundary conditions from new sets of calculations of WD H and He atmospheres. Along with the presentation of our new models, we will discuss critically how WD cooling times are affected by the progenitor metallicity, uncertainties on the current estimate of the $^{12}C(\\alpha,\\gamma)^{16}O$ reaction rate and treatment of convection during the progenitor evolution. A similar analysis (albeit with several differences in the details) can be found in Prada-Moroni \\& Straniero~(2002), without taking into account CO phase separation upon crystallization in the WD cooling models. The paper is structured as follows. Section~2 presents briefly the updates in the input physics compared to S00, and discusses critically our choices for the core chemical stratifications. Section~3 analyzes the main properties of the cooling models and WD isochrones, shows comparisons with S00 calculations and an example of application to study WDs in the solar neighborhood. A summary follows in Sect.~4. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % ", "conclusions": "\\label{summary} We have expanded our BaSTI stellar evolution archive by including new, updated WD cooling models, computed using the CO stratification obtained from BaSTI AGB progenitor calculations. Improvements with respect to the S00 set of WD models concern the CO chemical profiles, that have been obtained employing an updated estimate of the $^{12}C(\\alpha,\\gamma)^{16}O$ reaction rate, and the inclusion of a full set of He-atmosphere WD models, computed with appropriate boundary conditions from non-gray model atmospheres. The reference set of WD models that will be made public at the BaSTI website makes use of the CO stratification at the first thermal pulse from progenitor models calculated with intial metal mass fraction Z=0.0198, and the inclusion of convective core overshooting during the main sequence. To assess how sensitive the models are to these assumptions, we have tested the effect of uncertainties on the recent determination of the $^{12}C(\\alpha,\\gamma)^{16}O$ reaction rate employed in the progenitor models, the inclusion/exclusion of core convective overshooting during the main sequence, different approaches for quenching the breathing pulses at the end of core He-burning, a variation of the metallicity of the progenitor, a variation of the number of pulses experienced by the progenitor models. The results of this analysis indicate that the uncertainty on the $^{12}C(\\alpha,\\gamma)^{16}O$ reaction rate and the numerical approach used for inhibiting the breathing pulses have on the whole the largest impact on the WD cooling times, of about 7\\% at most- or of about 3\\% when the effect of phase separation upon crystallization is neglected. The progenitor metallicity, convective core overshooting during the main sequence phase, number of pulses before the WD formation, have overall a smaller effect. We have discussed quantitatively differences in the mass-radius relationships and cooling speed of H- and He-atmosphere cooling models. The radii of the He-atmosphere models of a given mass are systematically lower than their H-atmosphere counterparts. Differences range between $\\sim$ 9\\% and $\\sim$ 3\\%, increasing with decreasing $M_{WD}$ and/or increasing temperature. He-atmosphere models show typically longer cooling times down to log($L/L_{\\odot}$)$\\approx -$4, before starting to cool down much faster at lower luminosities. We have also estimated the differences between ages of star clusters obtained employing our new H- and He-atmosphere WD models, as well as the S00 H-atmosphere WD calculations. Ages derived from S00 H-atmosphere models show only relatively small differences when compared to our new calculations. As an example of application of our new set of models to real data, we have estimated an age of $\\sim$ 12~Gyr for the onset of star formation in the solar neighborhood, by fitting the local WD LF compiled by Catalan et al.~(2008). We have also studied the variation of the number ratio (N(He)/N(H)) with $T_{eff}$, predicted by our simulation of the local WDs. Due to the different cooling times of H- and He-atmosphere models, we show how this ratio changes with $T_{eff}$, increasing below $T_{eff}\\sim$10000~K, as observed. However, at least with our assumptions about the formation of the local WDs -- a constant progenitor formation rate and a (N(He)/N(H)) ratio at the onset of the WD phase that is constant with time -- the predicted ratio drops well below the observed value when $T_{eff}$ is lower than 7000-8000~K. This result can be explained in terms of the spectral transformation of a fraction of H-atmosphere objects, that increases with decreasing $T_{eff}$ below 7000-8000~K, reaching a maximum of $\\sim$24\\% at the lowest temperatures sampled by the observational data. As a consequence, one needs a broad range of H-layer thickness in solar neighborhood H-atmosphere WDs to explain these spectral changes, thicker envelopes being mixed with the underlying more massive He-layers at increasingly lower $T_{eff}$. All cooling tracks and the reference chemical stratifications will be made publicly available at the official BaSTI website (\\url{http://www.oa-teramo.inaf.it/BASTI}). In addition, we provide WD isochrones for ages between 200~Myr and 14~Gyr for both H- and He-atmosphere objects (with and without the inclusion of phase separation) using as a reference the IFMR by Salaris et al.~(2009) and the progenitor lifetimes from BaSTI models including convective core overshooting on the main sequence. The isochrones will be available for progenitors with both scaled solar and $\\alpha$-enhanced mixtures, and 11 values of the metal fraction Z, ranging from Z=0.0001 to Z=0.04. For both cooling tracks and isochrones we provide magnitudes in the UBVRIJHK, and HST ACS photometric systems. % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "1005/1005.5398_arXiv.txt": { "abstract": "Cygnus X-1 was the first X-ray source widely accepted to be a black hole candidate and remains among the most studied astronomical objects in its class. The detection of non-thermal radio, hard X-rays and gamma rays reveals the fact that this kind of objects are capable of accelerating particles up to very high energies. In order to explain the electromagnetic emission from Cygnus X-1 in the low-hard state we present a model of a black hole corona with both relativistic lepton and hadron content. We characterize the corona as a two-temperature hot plasma plus a mixed non-thermal population in which energetic particles interact with magnetic, photon and matter fields. Our calculations include the radiation emitted by secondary particles (pions, muons and electron/positron pairs). Finally, we take into account the effects of photon absorption. We compare the results obtained from our model with data of Cygnus X-1 obtained by the COMPTEL instrument. ", "introduction": "The low-hard state of accreting black holes is characterized by the presence of a hot corona around the compact object. Figure \\ref{fig:geometria} shows a scheme of the main components of the system. For this geometry, we assume a spherical corona with a radius $R_{\\rm{c}}$ and an accretion disk that penetrates the corona up to a radius $R_{\\rm{d}}