{ "0911/0911.0935_arXiv.txt": { "abstract": "The energy and momentum deposited by the radiation from accretion flows onto the supermassive black holes (BHs) that reside at the centres of virtually all galaxies can halt or even reverse gas inflow, providing a natural mechanism for supermassive BHs to regulate their growth and to couple their properties to those of their host galaxies. However, it remains unclear whether this self-regulation occurs on the scale at which the BH is gravitationally dominant, on that of the stellar bulge, the galaxy, or that of the entire dark matter halo. To answer this question, we use self-consistent simulations of the co-evolution of the BH and galaxy populations that reproduce the observed correlations between the masses of the BHs and the properties of their host galaxies. We first confirm unambiguously that the BHs regulate their growth: the amount of energy that the BHs inject into their surroundings remains unchanged when the fraction of the accreted rest mass energy that is injected, is varied by four orders of magnitude. The BHs simply adjust their masses so as to inject the same amount of energy. We then use simulations with artificially reduced star formation rates to demonstrate explicitly that BH mass is not set by the stellar mass. Instead, we find that it is determined by the mass of the dark matter 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. We predict that the black hole mass, $\\mbh$, scales with halo mass as $\\mbh \\propto \\mhalo^\\alpha$, with $\\alpha \\approx 1.55 \\pm 0.05$ and that the scatter around the mean relation in part reflects the scatter in the halo concentration-mass relation. ", "introduction": "Almost all massive galaxies are thought to contain a central supermassive black hole (BH) and the properties of these BHs are tightly correlated with those of the galaxies in which they reside \\citep[e.g.][]{magg98,ferr00,gebh00,trem02,hari04,hopk07b,ho08}. It is known that most of the mass of the BHs is assembled via luminous accretion of matter \\citep{solt82}. The energy emitted by this process provides a natural mechanism by which BHs can couple their properties to those of their host galaxies. Analytic \\citep[e.g.][]{silk98,haeh98,fabi99,adam01,king03,wyit03,murr05,merl08}, semi-analytic \\citep[e.g.][]{kauf00,catt01,gran04,bowe06} and hydrodynamical \\citep[e.g.][]{spri05,dima05,robe06,sija07,hopk07,dima08,okam08,boot09} studies have used this coupling between the energy emitted by luminous accretion and the gas local to the BH to investigate the origin of the observed correlation between BH and galaxy properties, and the buildup of the supermassive BH population. BHs are expected to regulate the rate at which they accrete gas down to the scale on which they are gravitationally dominant. For example, gas flowing in through an accretion disk can become so hot that its thermal emission becomes energetically important. Scattering of the photons emitted by the accreting matter by free electrons gives rise to the so-called Eddington limit. If the accretion rate exceeds this limit, which is inversely proportional to the assumed radiative efficiency of the accretion disk, then the radiative force exceeds the gravitational attraction of the BH and the inflow is quenched, at least within the region that is optically thin to the radiation. However, observations indicate that the time-averaged accretion rate is far below Eddington \\citep{koll06}, suggesting the presence of processes acting on larger scales. Indeed, the existence of tight correlations between the mass of the BH and the properties of the stellar bulge indicates that self-regulation may happen on the scale of the bulge \\citep[$\\sim1$~kpc; ][]{adam01,hopk07}, far exceeding the radius within which the BH is gravitationally dominant. However, since galaxy-wide processes such as galaxy mergers can trigger gas flows into the bulge \\citep{sand88,miho94}, it is conceivable that BHs could regulate their growth on the scale of the entire galaxy \\citep[$\\sim10$~kpc; ][]{haeh98,fabi99,wyit03} or even on that of the DM haloes hosting the galaxies \\citep[$\\sim10^2$~kpc; ][]{silk98,ferr02}. Finally, it is possible, perhaps even likely, that self-regulation takes place simultaneously on multiple scales. For example, frequent, short, Eddington-limited outbursts may be able to regulate the inflow of gas on the scale of the bulge averaged over much longer time scales. In this paper we investigate, using self-consistent simulations of the co-evolution of the BH and galaxy populations, on what scale the self-regulation of BHs takes place. In Sec.~\\ref{sec:method} we describe the numerical techniques and simulation set employed in this study. In Sec.~\\ref{sec:results} we demonstrate that BH self-regulation takes place on the scale of the DM halo, and that the BH mass is determined by the binding energy of the DM halo rather than by the stellar mass of the host galaxy. Throughout we assume a flat $\\Lambda$CDM cosmology with the cosmological parameters: $\\{\\Omega_{\\rm m},\\Omega_{\\rm b},\\Omega_\\Lambda,\\sigma_8,n_{\\rm s},h\\}=\\{0.238,0.0418,0.762,0.74,0.951,0.73\\}$, as determined from the WMAP 3-year data \\citep{sper07}. ", "conclusions": "\\label{sec:results} It is instructive to first consider under what conditions BHs can regulate their growth. To regulate its growth on a mass scale $\\msr$, a BH of mass $\\mbh$ must be able to inject energy (or momentum) at a rate that is sufficient to counteract the force of gravity on the scale $\\msr$, averaged over the dynamical time associated with this scale. The mass $\\msr$ could, for example, correspond to that of the BH, the stellar bulge, or the dark matter (DM) halo. If the BH cannot inject energy sufficiently rapidly, then gravity will win and its mass will increase. Provided that the maximum rate at which it can inject energy increases with $\\mbh$ (as is for example the case for Bondi-Hoyle and Eddington-limited accretion with a constant radiative efficiency) and provided that this rate increases sufficiently rapidly to counteract the growth of $\\msr$, the BH will ultimately reach the critical mass $m_{\\rm BH,crit}(\\msr)$ required to halt the inflow on the scale $\\msr$. If, on the other hand, $\\mbh \\gg m_{\\rm BH,crit}(\\msr)$, then the BH will quickly quench the accretion flow and its mass will consequently remain nearly unchanged. The BH will in that case return to the equilibrium value $m_{\\rm BH,crit}(\\msr)$ on the time scale which characterises the growth of $\\msr$. If the BH regulates its growth on the mass scale $\\msr$ and if $\\mbh \\ll \\msr$, then the critical rate of energy injection required for self-regulation is independent of the mass of the BH. It then follows from Eq.~\\ref{eq:edot} that $\\dot{m}_{\\rm BH}\\propto \\epsf^{-1}$, which implies \\begin{equation} \\left (\\mbh - \\mseed\\right ) \\propto \\epsf^{-1}\\,, \\label{eq:prop} \\end{equation} where $\\mseed$ is the initial mass of the BH. Hence, if the self-gravity of BHs is negligible on the maximum scale on which they regulate their growth and if $\\mbh \\gg \\mseed$, then we expect $\\mbh\\propto \\epsf^{-1}$. \\begin{figure} \\begin{center} \\includegraphics[width=8.3cm,clip]{massscale.eps} \\end{center} \\vspace{-0.4cm} \\caption{\\label{fig:massscale}Predicted redshift zero global BH mass density (black diamonds) and normalization of the $\\mbh-\\sigma$ relation (black plus signs) as a function of the assumed efficiency of BH feedback, $\\epsf$. Both quantities are normalized to their values in the simulation with $\\epsf=0.15$, which reproduces the observed relations between the mass of the BH and properties of the stellar bulge. Each point represents a different simulation. For $10^{-4}<\\epsf<1$ all data points track the dotted black line, which is a power-law with index minus one. This implies that in this regime BH mass is inversely proportional to $\\mbh$, and thus that the BHs inject energy into their surroundings at a rate that is independent of $\\epsf$, as expected for self-regulated growth on scales that are sufficiently large for the gravity of the BH to be unimportant. The red data points show results from simulations with a mass resolution that is 8 times worse than the fiducial simulation. The blue data points correspond to simulations with 64 times better resolution than our fiducial resolution, but show results for redshift 2 rather than zero. The agreement between the black, red and blue points confirms numerical convergence and demonstrates that the BHs are already self-regulating at redshift 2.} \\vspace{-0.4cm} \\end{figure} \\begin{figure*} \\begin{center} \\includegraphics[width=16.6cm,clip]{changing_baryons.eps} \\end{center} \\vspace{-0.8cm} \\caption{\\label{fig:changing_baryons}Median $\\mbh-\\mhalo$ (\\emph{left panel}) and $\\mbh-\\ms$ (\\emph{right panel}) relations for all BHs more massive than $10m_{\\rm seed}$. The black curves correspond to a simulation using our fiducial star formation law and the red, dashed curves show the result for a run in which the star formation efficiency was decreased by a factor of 100. In order to isolate the effect of stellar mass, we turned off supernova feedback in both runs. The BH scaling relations therefore differ somewhat from those predicted by our fiducial model, which does include supernova feedback. Baryons dominate the gravitational potential in the central regions of the galaxy when we use our fiducial star formation law, but DM dominates everywhere in the run with the reduced star formation efficiency. While the $\\mbh-\\ms$ relation is strongly affected by the change in the star formation efficiency, the relation between BH and halo mass remains invariant. This demonstrates that the BH mass is insensitive to the mass distribution on scales where the stellar mass dominates, and must instead be determined by the mass distribution on larger ($\\gg 10\\,{\\rm kpc}$) scales.} \\vspace{-0.4cm} \\end{figure*} The black diamonds plotted in Fig.~\\ref{fig:massscale} show the predicted global mass density in BHs at redshift zero as a function of $\\epsf$, the efficiency with which BHs couple energy into the ISM, normalised to the density obtained for $\\epsf=0.15$. Similarly, the black plus signs indicate the normalisation of the $\\mbh-\\sigma$ relation divided by that for the $\\epsf=0.15$ run. The feedback efficiency, $\\epsf$, is varied, in factors of 4, from $\\epsf=9.2\\times10^{-6}$ to $\\epsf=9.6$, which implies that the fraction of the accreted rest mass energy that is injected (i.e.\\ $\\epsr\\epsf$) varies from $9.2\\times 10^{-7}$ to 0.96. BH mass is clearly inversely proportional to the assumed feedback efficiency for $10^{-4}<\\epsf< 1$. For $\\epsf>1$ the trend breaks down because the BH masses remain similar to the assumed seed mass, in accord with Eq.~\\ref{eq:prop}. If we had used a lower seed mass, then the trend would have extended to greater values of $\\epsf$. The deviation from inverse proportionality that sets in below $\\epsf =10^{-4}$ is more interesting. Such low values yield BH masses that are more than $0.15/10^{-4}\\sim 10^{3}$ times greater than observed, in which case they are no longer negligible compared to the masses of their host galaxies. In that case the critical rate of energy deposition will no longer be independent of $\\mbh$ and we do not expect Eq.~\\ref{eq:prop} to hold. We have thus confirmed that feedback enables BHs to regulate their growth. Moreover, we demonstrated that this self-regulation takes places on scales over which the gravitational influence of the BHs is negligible, provided that the fraction of the accreted rest mass energy that is coupled back into the interstellar medium is $\\gtrsim 10^{-5}$. To test whether it is the stellar or the dark matter distribution that determines the mass of BHs, we compare the BH masses in two simulations that are identical except for the assumed efficiency of star formation. One uses our fiducial star formation law, but in the other simulation we reduced its amplitude by a factor of 100, making the gas consumption time scale much longer than the age of the Universe. Because changing the amount of stars would imply changing the rate of injection of supernova energy, which could affect the efficiency of BH feedback, we neglected feedback from star formation in both runs. In the simulation with \\lq normal\\rq\\ star formation the central regions of the galaxies are dominated gravitationally by the baryonic component of the galaxy, whereas in the simulation with reduced star formation the DM dominates everywhere. Fig.~\\ref{fig:changing_baryons} shows the $\\mbh-\\mhalo$ and $\\mbh-\\ms$ relations at redshift 0. While the two runs produce nearly identical BH masses for a fixed halo mass, the $\\mbh-\\ms$ relation is shifted to lower stellar masses by more than an order of magnitude in the model with reduced star formation. The insensitivity of the relation between $\\mbh$ and $\\mhalo$ to the assumed star formation efficiency demonstrates that the BH mass is not set by the gravitational potential on the scale of the galaxy. We have verified that the same result holds at redshift two for the simulations with 64 times better mass resolution. Clearly, stellar mass does not significantly influence the relation between the mass of the BH and that of its host halo. This implies that BH self-regulation occurs on the scale of DM haloes. If the rate by which the BHs inject energy is independent of the assumed feedback efficiency, then we expect the same to be true for the factor by which BH feedback suppresses star formation. This is confirmed by comparison of the global SFRs in runs with different values of $\\epsf$ (see Fig.~6 of BS09). \\begin{figure} \\begin{center} \\includegraphics[width=8.3cm,clip]{mhalombh.eps} \\end{center} \\vspace{-0.4cm} \\caption{\\label{fig:mhalombh}The relation between BH mass and DM halo mass for all BHs that belong to central galaxies and have masses greater than $10m_{\\rm seed}$. The DM halo mass, $m_{200}$, is defined as the mass enclosed within a sphere, centred on the potential minimum of the DM halo, that has a mean internal density of 200 times the critical density of the Universe. The grey pixels show the results from our fiducial simulation ($\\epsf=0.15$), with the colour of each pixel set by the logarithm of the number of BHs in that pixel. The solid, red line shows the observational determination of the $\\mbh-\\mhalo$ relation (Bandara et al. 2009) and has a slope of 1.55. The dotted, red lines show the 1$\\sigma$ errors on the observations. The simulation agrees very well with the observed relation. The value of the slope and the scatter (which correlates with the concentration of the DM halo) suggest that the halo binding energy, rather than mass, determines the masses of BHs.} \\vspace{-0.4cm} \\end{figure} Fig.~\\ref{fig:mhalombh} compares the predicted $\\log_{10}\\mbh-\\log_{10}\\mhalo$ relation with observation \\citep{band09}. The agreement is striking. The slope and normalization of the observed $\\log_{10}(\\mbh/\\msun)-\\log_{10}(\\mhalo/10^{13}\\,\\msun)$ relation are $1.55\\pm0.31$ and $8.18\\pm0.11$ respectively, whereas the simulation predicts $1.55\\pm0.05$ and $8.01\\pm0.04$. Note that the simulation was only tuned to match the normalization of the relations between $\\mbh$ and the galaxy stellar properties. If the energy injected by a BH is proportional to the halo gravitational binding energy, then, for isothermal models \\citep{silk98}, $\\mbh\\propto\\mhalo^{5/3}$. Here we extend these models to the more realistic universal halo density profile \\citep{nava97}, whose shape is specified by a concentration parameter, $c$ (we assumed $c\\propto v_{\\rm max}^2/v_{\\rm v}^2$, where $v_{\\rm max}$ and $v_{{\\rm v}}$ are the maximum halo circular velocity and the circular velocity at the virial radius respectively). It is known that concentration decreases with increasing halo mass, $c\\propto\\mhalo^{-0.1}$ \\citep{bulo01,duff08}, which then affects BH mass through the dependence of halo binding energy on concentration. If the total energy injected by a BH of a given mass is proportional to the energy required to unbind gas from a DM halo \\citep{loka01} out to some fraction of the virial radius, $\\re/r_{\\rm v}$ then \\begin{align} \\mbh\\propto&\\Bigg(\\frac{c}{\\big(\\ln(1+c)-c/(1+c)\\big)^2}\\Bigg)\\times \\nonumber\\\\ &\\Bigg(1-\\frac{1}{(1+c\\frac{\\re}{r_{\\rm v}})^2}-\\frac{2\\ln(1+c\\frac{\\re}{r_{\\rm v}})}{1+c\\frac{\\re}{r_{\\rm v}}}\\Bigg)m_{\\rm v}^{5/3}\\,. \\end{align} Inserting $c\\propto m_{\\rm v}^{-0.1}$ and computing the logarithmic derivative with respect to $m_{\\rm v}$ in the mass range $10^{10}\\,\\msun < m_{\\rm v} < 10^{14}\\,\\msun$, we find that the slope is a weak function of $\\re/r_{\\rm v}$ that varies from 1.50 at $\\re=10^{-1}r_{\\rm v}$ to 1.61 at $\\re=r_{\\rm v}$. The close match between theory, simulation and observation suggests that the halo binding energy, rather than halo mass, determines the mass of the BH. The residuals from the $\\mbh-\\mhalo$ relation ($\\Delta\\log_{10}\\mbh$) are correlated with halo concentration (Spearman rank correlation coefficient $\\rho=0.29$, probability of significance $P= 0.9998$) as would be expected if $\\mbh$ is sensitive to the halo binding energy. The residuals are also correlated with galaxy stellar mass, though much less strongly ($\\rho=0.09$; $P=0.96$). Taken together, these correlations tell us that, at a given halo mass, galaxies with BHs more massive than the average will also contain a larger than average amount of stars, and are hosted by more concentrated haloes. This suggests that the galaxy stellar mass is also determined by the halo binding energy. Thus, outliers in the $\\mbh-\\mhalo$ relation may still lie close to the mean $\\mbh-\\ms$ relation. Furthermore, higher concentrations imply earlier formation times and spheroidal components do indeed typically host old stellar populations. In addition to the \\lq quasar mode\\rq\\ of feedback discussed in this work, it has recently become clear that a second \\lq radio mode\\rq\\ may be required to quench cooling flows in galaxy groups and clusters \\citep[see e.g.][for a review]{catt09}. Although we do not explicitly include a \\lq radio mode\\rq\\ in the current work, the AGN feedback prescription explored here is capable of suppressing cooling flows, at least on group scales, providing excellent matches to observed group density and temperature profiles as well as galaxy stellar masses and age distributions \\citep{mcca09}. It is known that BHs obtain most of their mass in the \\lq quasar mode\\rq\\ \\citep{solt82} so any discussion of what detemines the masses of BHs must focus primarily on this mode of accretion. Finally, the ability of a BH to quench cooling flows in the \\lq radio mode\\rq\\ is expected to be closely related the virial properties of the hot halo \\citep{catt09} and would therefore provide an additional link between BHs and DM haloes over and above what we discuss here and so serve to make any fundamental connection between BH mass and the properties of the DM halo even stronger. We conclude that our simulation results suggest that in order to effectively halt BH (and galaxy) growth, gas must not return to the galaxy on a short timescale. This requires that the BH injects enough energy to eject gas out to scales where the DM halo potential is dominant. The mass of the BH is therefore determined primarily by the mass of the DM halo with a secondary dependence on halo concentration, of the form that would be expected if the BH mass were controlled by the halo binding energy. The tight correlation between $\\mbh$ and $\\ms$ is then a consequence of the more fundamental relations between halo binding energy and both $\\mbh$ and $\\ms$. \\vspace{-0.4cm}" }, "0911/0911.2792_arXiv.txt": { "abstract": "In cold dark matter cosmologies, small mass halos outnumber larger mass halos at any redshift. However, the lower bound for the mass of a galaxy is unknown, as are the typical luminosity of the smallest galaxies and their numbers in the universe. The answers depend on the extent to which star formation in the first population of small mass halos may be suppressed by radiative feedback loops operating over cosmological distance scales. If early populations of dwarf galaxies did form in significant number, their relics should be found today in the Local Group. These relics have been termed ``fossils of the first galaxies''. This paper is a review that summarizes our ongoing efforts to simulate and identify these fossils around the Milky Way and Andromeda. It is widely believed that reionization of the intergalactic medium would have stopped star formation in the fossils of the first galaxies. Thus, they should be among the oldest objects in the Universe. However, here we dispute this idea and discuss a physical mechanism whereby relatively recent episodes of gas accretion and star formation would be produced in some fossils of the first galaxies. We argue that fossils may be characterized either by a single old population of stars or by a bimodal star formation history. We also propose that the same mechanism could turn small mass dark halos formed before reionization into gas-rich but starless ``dark galaxies''. We believe that current observational data supports the thesis that a fraction of the new ultra-faint dwarfs recently discovered in the Local Group are in fact fossils of the first galaxies. ", "introduction": "There are many questions that remain open in cosmology with regard to the mass, number and properties of the smallest galaxies in the universe. Have we already discovered the smallest galaxies in the universe or we are still missing an elusive but large population of ultra-faint dwarf galaxies? In cold dark matter (CDM) cosmologies most of the dark halos that formed before reionization had masses smaller than $10^8-10^9$~M$_{\\odot}$ \\citep[\\eg,][]{GnedinO:97}. The small mass halos that survived tidal destruction to the modern epoch, were they able to form stars, would constitute a sub-population of dwarf satellites orbiting larger halos. Small mass dark halos significantly outnumber more massive galaxies like the Milky Way and can be located in the voids between luminous galaxies \\citep[\\eg,][]{Hoeft:06, Ricotti:09}. However, until recently (\\ie, before 2005) observations did not show a large number of satellites around massive galaxies like the Milky Way and Andromeda. This became known as the ``missing galactic satellite problem, \\citep{Klypin:99, Moore:99}. The voids between bright galaxies appear to be devoid of dwarf galaxies \\citep{Karachentsev:04, Karachentsevetal06, Tullyetal06}. While the abundance of dwarfs in large voids may not pose a problem to CDM cosmology, as shown by \\cite{TinkerC:09}, it is unclear whether the predictions of the number of faint dwarfs in the Local Group is consistent with both the number of observed Milky Way dwarf satellites and the number of relatively isolated dwarfs in the local voids. Historically, the discrepancy between observation and theory on the number of dwarf galaxies has been interpreted in two ways: 1) as a problem with the CDM paradigm that could be solved by a modification of the dark matter properties -- for instance by introducing warm dark matter \\citep[\\eg,][]{BodeO:01} -- or 2) as an indicator of feedback processes that are exceptionally efficient in preventing star formation in small mass halos, which remain mostly dark \\citep[\\eg,][]{HaimanAR:00}. The recently discovered population of ultra-faint dwarfs \\citep{Belokurovetal06a, Belokurovetal07, Irwinetal07, Walshetal07, Willmanetal05ApJ, Willmanetal05AJ, Zuckeretal06a, Zuckeretal06b, Ibataetal07, Majewskietal07, Martinetal06} in combination with a proper treatment of observational incompleteness \\citep{SimonGeha07, Koposov:08, Tollerudetal08, Gehaetal08} has increased the estimated number of Milky Way satellites to a level that can be more easily reconciled with theoretical expectations. For instance, the suppression of dwarf galaxy formation due to intergalactic medium (IGM) reheating during reionization \\citep{Babul:92, Efstathiou:92b, ShapiroG:94, HaimanRL:96, Thoul:96, Quinn:96, Weinberg:97, NavarroS:97, Bullock:00, Gnedin:00, Somerville:02, Dijkstra:04, Shapiro:04, Hoeft:06, OkamotoGT08, Ricotti:09}, in conjunction with a strong suppression of star formation in small mass pre-reionization dwarfs, may be sufficient to explain the observed number of Milky Way satellites. In the near future we can hope to answer perhaps a more interesting question: what is the minimum mass that a galaxy can have? This is a non trivial and fundamental question in cosmology. Answering it requires a better understanding of the feedback mechanisms that regulate the formation of the first galaxies before reionization and the details of the process of reionization feedback itself. The formation of the first dwarf galaxies - before reionization - is self-regulated on cosmological distance scales. This means that the fate of small mass halos (\\ie, whether they remain dark or form stars) depends on local and global feedback effects. This type of galaxy feedback differs from the more familiar model operating in normal galaxies (\\eg, SN explosions, AGN feedback, etc), where the feedback is responsible for regulating the star formation rate within the galaxy itself but does not impact star formation in other distant galaxies. Rather, before reionization, each proto-dwarf galaxy reacts to the existence of all the others. Different theoretical assumptions and models for the cosmological self-regulation mechanisms will, of course, produce different predictions for the number and luminosity of the first dwarf galaxies \\citep{HaimanAR:00, RicottiGnedinShull02b, WiseA:08, OShea:08, RicottiGnedinShull08}. We now introduce the basic concepts on how feedback-regulated galaxy formation operates in the early universe (\\ie, before reionization):\\\\ A cooling mechanism for the gas is required in order to initiate star formation in dark halos. In proto-galaxies that form after reionization this is initially provided by hydrogen Lyman-alpha emission. This cooling is efficient at gas temperatures of 20,000 K but becomes negligible below $T \\sim 10,000$~K. Later, as the temperature drops below 10,000 K, the cooling is typically provided by metal line cooling. In the first galaxies, however, both these cooling mechanisms may be absent. This is because the first dwarf galaxies differ when compared to present-day galaxies in two respects: 1) they lack important coolants -- such as carbon and oxygen -- because the gas is nearly primordial in composition, and 2) due to the smaller typical masses of the first dark halos, the gas initially has a temperature that is too low to cool by Lyman-alpha emission. The gas in small mass halos with circular velocity $v_{\\rm vir}=(GM_{\\rm tot}/r_{\\rm vir})^{1/2} \\simlt 20$~km~s$^{-1}$, where $r_{\\rm vir}$ is the virial radius -- roughly corresponding to a mass $M_{\\rm tot} \\simlt 10^8$ M$_{\\odot}$ at the typical redshift of virialization -- has a temperature at virialization $T \\simlt 10,000$ K. Hence, if the gas has primordial composition it is unable to cool by Lyman-alpha emission and initiate star formation unless it can form a sufficient amount of primordial H$_2$ (an abundance $x_{H_2} \\simgt 10^{-4}$ is required). Because molecular hydrogen is easily destroyed by far-ultraviolet (FUV) radiation in the Lyman-Werner bands ($11.3 13.6$~eV) in the extreme-ultraviolet (EUV) plays a far more important role in regulating the formation of the first galaxies than FUV radiation. Thus, in our opinion, models that do not include 3D radiative transfer of H and He ionizing radiation cannot capture the most relevant feedback mechanism that regulates galaxy formation in the early universe \\citep{RicottiGnedinShull02a, RicottiGnedinShull02b}. After reionization, the formation of dwarf galaxies with $v_{\\rm vir}<20$~km~s$^{-1}$ is strongly inhibited by the increase in the Jeans mass in the IGM. Thus, according to this model, reionization feedback and negative feedback due to H$_2$ photo-dissociation by the FUV background (important before reionization) determine the mass of the smallest galactic building blocks. The resultant circular velocity of the smallest galactic building blocks is $v_{\\rm vir} \\sim 20$ km~s$^{-1}$, roughly corresponding to masses $M_{\\rm tot} \\sim 10^8-10^9$~M$_\\odot$. If this is what really happens in the early universe, the ``missing Galactic satellite problem'' can be considered qualitatively solved because the predicted number of Milky Way satellites with $v_{\\rm vir}>20$ km~s$^{-1}$ is already comparable to the estimated number of observed satellites after applying completeness corrections (although this model may still have problems reproducing the observations in detail). However, as briefly mentioned above, we have argued for some time that most simulations of the first galaxies cannot capture the main feedback mechanism operating in the early universe because they do not include a key physical ingredient: radiative transfer of H and He ionizing radiation. Our simulations of the formation of the first galaxies are to date the only simulations of a cosmologically representative volume of the universe (at $z \\sim 10$) that include 3D radiative transfer of H and He ionizing radiation \\citep{RicottiGnedinShull02a, RicottiGnedinShull02b, RicottiGS:08}. Figure~\\ref{fig:pfr2} shows the evolution of ionized bubbles around the first galaxies in a cubic volume of 1.5~Mpc in size at redshifts $z=21.2, 17.2, 15.7, 13.3$ from one of our simulations. The results suggest that negative feedback from the FUV background is not the dominant feedback mechanism that regulates galaxy formation before reionization. Rather, ``positive feedback'' on H$_2$ formation from ionizing radiation \\citep{HaimanRL:96, RicottiGS:01} dominates over the negative feedback of H$_2$ dissociating radiation. Hence, a strong suppression of galaxy formation in halos with $v_{\\rm vir}<20$~km~s$^{-1}$ does not take place. In this latter case, some galactic satellites would be the fossil remnants of the first galaxies. Comparisons of simulated pre-reionization fossils to dwarf spheroidals in the Local Group show remarkable agreement in properties \\citep[][hereafter RG05]{RicottiGnedin05}. Based on the results of the simulations, we also suggested the existence of the ultra-faint population before it was discovered about a year later \\citep[see RG05,][]{BovillR:09}. \\begin{figure*}[pth] \\epsscale{1.1} \\plotone{4tile1.eps} \\caption{3D rendering of cosmological \\HII~ regions (fully ionized gas is red and partially ionized gas is blue) around the first galaxies in a box of 1.5~Mpc. The four boxes show a time sequence at redshifts $z=21.2, 17.2, 15.7, 13.3$ for the simulation S2 from \\cite{RicottiGnedinShull02b}. The rendering shows several tens of small size \\HII~ regions around the first galaxies (there are a few hundreds of galaxies in this volume). A movie of the same simulation shows that the \\HII~ regions are short lived: they form, expand to a size comparable to the large scale filamentary structure of the IGM and recombine, promoting the formation of molecular hydrogen inside the relic \\HII~ regions.} \\label{fig:pfr2} \\end{figure*} \\subsection{Definition of ``pre-reionization fossils'' } Throughout this paper we define ``pre-reionization fossils'' as the dwarfs hosted in halos with a maximum circular velocity remaining below 20~km~s$^{-1}$ at all times during their evolution: $v_{\\rm max}(t)<20$~km~s$^{-1}$. It will become clear in this paper that this definition is {\\it not} directly related to the ability of fossils to retain gas and form stars after reionization: in \\S~\\ref{sec:infall} we describe a mechanism in which small mass halos with $v_{\\rm max}(t)< 20$~km~s$^{-1}$ are able to have a late phase of gas accretion and possibly star formation. Our definition of fossil reflects the {\\it special cooling mechanisms and feedback processes} that regulate star formation and the number of luminous halos with $v_{\\rm max}(t)<20$~km~s$^{-1}$, before and {\\it after} reionization. In proto-fossil galaxies -- even adopting the most conservative assumption of maximum efficiency of shock heating of the gas during virialization -- the gas is heated to a temperature below $T \\sim 10,000$~K. Thus the gas cannot cool by Lyman-alpha emission, a very efficient coolant. The cooling of the gas is dominated either by H$_2$ roto-vibrational line emission or by metal cooling, important if the metallicity exceeds $Z \\sim 10^{-3}$~Z$_\\odot$ \\citep[\\eg,][]{BrommF:01, SantoroS:06, RicottiGS:08, Smithetal:09}. These coolants are much less efficient than Lyman-alpha emission. Moreover, H$_2$ abundance and cooling is modulated and often suppressed by the FUV and EUV radiation fields. The FUV radiation in the H$_2$ Lyman-Werner bands and hard ultraviolet radiation have large mean free paths with respect to the typical distances between galaxies, thus their feedback is global in nature. Qualitatively, this explains why the first galaxies have low luminosities and low surface-brightness, similar to dwarf spheroidal (dSph) galaxies in the Local Group \\citep[RG05,][]{BovillR:09, SalvadoriF:09}. Simulations also show that stars in the the first galaxies do not form in a disk but in a spheroid \\citep[RG05,][]{RicottiGS:08}. A thin galactic disk is not formed because of the high merger rates and the low masses of dark halos in the early universe. Roughly, pre-reionization fossils have a mass at virialization $M_{\\rm tot}<10^8$~M$_\\odot$, assuming they form at $z_{\\rm vir} \\sim 10$, but their mass may increase by up to one order of magnitude by $z=0$ due to secondary infall \\citep[RGS02a, b,][]{ RicottiGnedinShull08}. Secondary infall does not affect $v_{\\rm max}$, which remains roughly constant after virialization. \\subsection{Pre-reionization fossils and reionization} The critical value of $v_{\\rm max, crit}$ for which dwarf galaxy formation is suppressed by reionization feedback is close to the 20~km~s$^{-1}$ value that defines a fossil, but it is {\\it not necessarily the same value}. Indeed, it can be significantly larger than 20~km~s$^{-1}$ if the IGM is heated to $T \\gg 10,000$~K \\citep{RicottiGS:00}. Thus, we expect that the virialization of new ``pre-reionization fossils'' is strongly suppressed after reionization due to IGM reheating (\\ie, they mostly form before reionization). However, pre-reionization fossils and dark halos with $v_{\\rm max}<20$~km~s$^{-1}$ that virialized before reionization may accrete gas and, in certain cases, form new stars after reionization at redshifts $z<1-2$ \\citep{Ricotti:09}. Unfortunately, the value of $v_{\\rm max, crit}$ is uncertain due to our poor understanding of the thermal history of the IGM \\citep{RicottiGS:00}. The uncertainty surrounding the IGM equation of state may partially explain the differences found in literature on the values for $v_{\\rm max, crit}$ and the different levels of suppression of star formation as a function of the halo mass after reionization \\citep[\\eg,][]{Weinberg:97, Gnedin:00, Hoeft:06, OkamotoGT08}. Regardless of assumptions for the reionization feedback model, one should bear in mind that no halo with $v_{\\rm max} <20$~km~s$^{-1}$ can form stars after reionization unless the gas in those halos has been significantly pre-enriched with metals. For instance, the model by \\cite{Koposov:09} assumes star formation after reionization in halos as small as $v_{\\rm max}\\sim 10$~km~s$^{-1}$. With this assumption they find that their model is consistent with observations of ultra-faint dwarfs but claim that fossils are not needed to explain the data. However, star formation in such small halos can only take place in a gas that was pre-enriched with metals, suggesting the existence of older populations of stars in those halos. Indeed, according to our definition, the smallest post-reionization dwarfs with $10~{\\rm km~s}^{-1}< v_{\\rm max} < 20$~km~s$^{-1}$ in the \\cite{Koposov:09} model are ``fossils''. As stated above, fossils may also be able to form stars after reionization due to a late phase of cold gas accretion from the IGM \\citep{Ricotti:09}. \\subsection{Identification of pre-reionization fossils in observations} Pre-reionization fossils are not easily identifiable because $v_{\\rm max}$ cannot be measured directly from observations. Understanding the star formation history of dwarf galaxies may help in this respect, as fossils likely show some degree of suppression of their star formation rate occurring about 12.5 Gyr ago due to reionization. However, their identification based on their star formation history may be complicated because some pre-reionization fossils in the last 10~Gyr may have had a late phase of gas accretion and star formation. The caveat is that star formation histories cannot be measured with accuracy better than to within 1-2~Gyr and the accuracy becomes increasingly poorer for old stellar populations. Thus, it is impossible to prove whether an old population of stars formed before reionization (which happened about 1~Gyr after the Big Bang) or at $z \\sim 3$, when the Milky Way was assembled. Nevertheless, ultra-faint dwarfs that show some degree of bimodality in their star formation history are candidates for being pre-reionization fossils. According to results by RG05 and \\cite{BovillR:09}, Willman~1, Bootes~II, Segue~1 and Segue~2 do not lie on the luminosity-surface brightness relationship of simulated pre-reionization fossils. This result is based on the assumption that fossil properties are not modified by tides. Their surface brightness is larger than the model predictions for objects with such low-luminosity. An as yet undiscovered population of ultra-faints with lower surface brightness is instead predicted by our simulations. It is likely that the properties of the lowest luminosity ultra-faints may have been modified by tidal forces due to their proximity to the Milky Way disk. Although it is difficult to identify individual fossils, statistical arguments suggest that at least some ultra-faint dwarf galaxies are pre-reionization fossils. This is because the number of satellites from N-body simulations with $v_{\\rm max}(t)>20$~km~s$^{-1}$ is substantially smaller than the estimated number of observed satellites after completeness corrections. Admittedly the current theoretical and observational uncertainties on the number of satellites are still large. However, if the estimated number (after completeness corrections) of ultra-faint dwarfs increases further, the existence of pre-reionization fossils will be inescapably proven. This is especially the case if a population of ultra-faint dwarfs with luminosities similar to Willman~1, Bootes~II, Segue~1 and Segue~2 but surface brightness below the current sensitivity limit of the SDSS -- as predicted by our simulations -- is discovered. The possibility of identifying the fossils of the first galaxies in our own backyard is very exciting. It would greatly improve our understanding of the physics involved in self-regulating the formation of the first galaxies before reionization. Clearly, even the launch of the James Webb Space Telescope ({\\it JWST}), would not yield the wealth of observational data on the formation of the first galaxies that could be obtained by studying ultra-faint galaxies in the Local Group. The rest of the paper is organized as follows. In \\S~\\ref{sec:obs} we briefly review and discuss observational data on Galactic satellites, in \\S~\\ref{sec:theo} we summarize the results of simulations of the formation of the first galaxies in a cosmological volume and the effect of reionization feedback on galaxy formation. In \\S~\\ref{sec:infall} we discuss a recently proposed model for ``late gas accretion'' from the IGM onto small mass halos. In \\S~\\ref{sec:comp} we compare the theoretical properties of simulated pre-reionization fossils to observations. In \\S~\\ref{sec:disc} we compare different ideas for the origin of classical and ultra-faint dwarf spheroidals. We present our conclusions in \\S~\\ref{sec:conc}. ", "conclusions": "\\label{sec:conc} We have summarized our work on the formation of the first galaxies before reionization (\\ie, pre-reionization dwarfs) and the quest to identify the fossils of these first galaxies in the Local Group. The definition of a pre-reionization fossil is not directly related to the suppression of star formation experienced by these galaxies due to reionization feedback. Indeed, we discussed how pre-reionization fossils may experience a late phase of gas accretion and possibly star formation at redshift $z<1-2$. Most importantly, fossils are a population of dwarf galaxies whose formation (\\ie, the fraction of halos that are luminous) is self-regulated on cosmological distance scales by radiative processes. Their existence is not certain due to a possible strong negative feedback that may prevent the majority of these halos from ever forming stars. In addition, if negative feedback heavily suppresses the number and luminosity of these first galaxies, more massive halos with $v_{\\rm max}>20$~km~s$^{-1}$ will evolve differently because of the lower level of metal pre-enrichment of the IGM. To summarize, the critical circular velocity $v_{\\rm max} \\sim 20$~km~s$^{-1}$ that we adopt to define a fossil is primarily motivated by fundamental differences in cooling and feedback processes that regulate star formation in these halos in the early universe. However, it is also close to the critical value for continued gas accretion after IGM reionization \\citep{Gnedin-filteringmass00, Hoeft:06, OkamotoGT08}. The number of Milky Way and M31 satellites provides an indirect test of galaxy formation and the importance of positive and negative feedback in the early universe. This test, although the uncertainties are large, supports the idea that a fraction of the new ultra-faint dwarfs are fossils. The good agreement of the SDSS and new M31 ultra-faint dwarf properties with predictions of our simulations (RG05, GK06, Bovill \\& Ricotti 2009) does not prove the primordial origin of the new ultra-faint dwarfs, but it supports this possibility. More theoretical work and more observational data are needed to prove that some dwarfs in the Local Group are true fossils of the first galaxies. Future theoretical work should focus on improving the accuracy of predictions on the properties of dwarf galaxies formed before reionization and their evolution to the present day. Modeling the evolution of the baryonic component after reionization in dwarf satellites and in the Milky Way -- Andromeda system may be necessary to make robust predictions. More observational data will certainly be available in the near future. A large number of surveys, both at optical and radio wavelengths will be online in the near future (\\eg, {\\it Pan-STARRS, LSST, ALMA, EVLA, JWST, SKA} to mention a few). Different survey strategies may be used to find and characterize fossil dwarf galaxies. A deep pencil beam survey would be useful to find the faintest dwarf satellites of the Milky Way and determine more precisely their Galactocentric distribution. A shallower all sky survey could be used to quantify the degree of anisotropy in the distribution of satellites around the Milky Way. The star formation history of the dwarf galaxies is not strongly discriminatory because fossil galaxies may have a late phase of gas accretion and star formation during the last $9-10$~Gyrs \\citep{Ricotti:09}. The distinction between fossils and non-fossil galaxies may be quite elusive but it is nevertheless important to understand galaxy formation and feedback in the early universe. Arguments based on counting the number of dwarfs in the Local universe are among the more solid arguments that could be used to prove the existence of fossil galaxies (see Table~\\ref{tab:count}). Future tests may be provided by deep surveys looking for ultra-faint galaxies in the local voids or looking for gas in dark galaxies (\\ie, dark halos that have been able to accrete gas from the IGM at $z<1-2$). Ultra-faint dwarfs should be present in the voids if dwarf galaxies formed in large numbers before reionization (Bovill \\& Ricotti, in preparation). If pre-reionization dwarfs never formed due to dominant negative feedback in the early universe, it is possible that a faint (in H$\\alpha$ and 21~cm emission) population of dark galaxies exists in the outer parts of the Local Group. Hence, another way to detect fossil galaxies in the outer parts of the Milky Way or outside the super-galactic plane would be to search for neutral or ionized gas that they may have accreted from the IGM. Future radio telescopes (\\eg, {\\it ALMA, EVLA, SKA}) may be able to detect neutral hydrogen in dark galaxies or in ultra-faint dwarfs. Ionized gas in the outer parts of dark halos may be observed in absorption along the line of sight of distant quasars (for instance in \\mbox{\\ion{O}{6}} or \\mbox{\\ion{O}{4}} with {\\it COS} on the {\\it HST}). However, the probability that a line of sight toward a quasar intersects the ionized gas collected from the IGM by dark or fossil galaxies might be small. Additional theoretical work is required to address these issues." }, "0911/0911.2271_arXiv.txt": { "abstract": "Protoplanetary disks are composed primarily of gas (99\\% of the mass). Nevertheless, relatively few observational constraints exist for the gas in disks. In this review, I discuss several observational diagnostics in the UV, optical, near-IR, mid-IR, and (sub)-mm wavelengths that have been employed to study the gas in the disks of young stellar objects. I concentrate in diagnostics that probe the inner 20 AU of the disk, the region where planets are expected to form. I discuss the potential and limitations of each gas tracer and present prospects for future research. ", "introduction": "\\label{intro} At the time when giant planets form, the mass (99\\%) of the protoplanetary disk is dominated by gas. Dust is a minor constituent of the mass of the disk. Still, it dominates the opacity; consequently, it is much easier to observe. Therefore, most of the observational constraints of disks have been deduced from the study of dust emission (e.g., see chapter by Henning et al.). In contrast, direct observational constraints of the gas in the disk are relatively scarce. Nonetheless, to obtain direct information from the gas content of the disk is crucial for answering fundamental questions in planet formation such as: how long does the protoplanetary disk last?, how much material is available for forming giant planets?, how do the density and the temperature of the disk vary as a function of the radius?, what are the dynamics of the disk?. Although we have learned important insights from disks from dust observations (see for example the reviews by Henning et al. 2006 and Natta et al. 2007), dust presents several limitations: (i) dust spectral features are broad; consequently, dust emission does not provide kinematical information; (ii) dust properties are expected to change during the planet formation process; therefore, quantities such as the gas-to-dust ratio (needed to derive the disk mass from dust continuum emission in the (sub)-mm) are expected to strongly vary with respect to the conditions of the Interstellar Medium (ISM). In addition, dust signatures are strongly related to dust size. In particular, as soon a dust particle reaches a size close to a decimeter it becomes practically \"invisible\". For example, a crucial quantity such as the disk dissipation time scale is normally deduced from the decline of the fraction of the sources presenting near-IR (JHKL) excess in clusters of increasing ages (e.g., Haisch et al. 2001). In reality, the time scale that we want to constrain is not the time scale in which the small warm dust particles disappear from the surface layers in the inner disk, but the time scale in which the gas in the disk disappears. In summary, independent information of the gas is crucial to understand protoplanetary disk structure and {\\it combined studies of gas and dust in the disk are needed}. In particular, we are interested in deriving constraints for the inner disk, the region where the planets are expected to form (R$<$20 AU). The observational study of the gas in the inner disk is a relatively new emerging topic. Only with the advent of ground-based high-spectral resolution infrared spectrographs and space-born infrared spectrographs has this research become possible. \\begin{figure} \\includegraphics[width=0.5\\textwidth]{disk_emission.jpg} \\caption{Line shape, inclination and the emitting region of a disk in Keplerian rotation. The double peaked profile indicates emission from a rotating disk. The line width and double peak separation depend on the inclination of the disk. The line profile wings depend on the innermost radius of the emission. In this toy model, it is assumed that the intensity is constant as a function of the radius and a spectral resolution of 1 km/s.} \\label{fig:1} % \\end{figure} The main tool used to study the gas in the disk is molecular spectroscopy. The gas in the disk is heated by collisions with dust, shocks or UV or X-rays from the central star. The heated molecules populate different rotational and vibrational levels and when they de-excite emission lines are produced. The key is that these emission lines have the imprint of the physical conditions of the gas where they originated. Important constraints can be derived from emission lines: (i) line ratios of different transitions constrain the excitation mechanism (shocks, UV or X-rays) and the temperature of the gas responsible of the emission; (ii) if the emitting gas is optically thin, then the measured line fluxes are proportionally related to the amount of molecules present; therefore, line fluxes can be used to derive column densities and gas masses; (iii) line shapes and spatial extended provide constraints on the size and geometry of emitting region as well as the dynamics of the emitting gas. Figure 1 displays a pedagogical example of how a line profile changes as a function of the inclination and size of the emitting region in the disk. The line width and double peak separation depend on the inclination . The line profile wings depend on the innermost radius of the emission. For the interested reader, recent reviews by Najita et al. (2000 \\& 2007a), Blake (2003) and Carr (2005) cover the subject of observational studies of gas in disks. \\begin{figure} \\includegraphics[width=0.5\\textwidth]{disk_structure.jpg} \\label{fig:2} \\caption{Cartoon of the structure of an optically thick disk. The disk has a hot inner region (R$<$20 AU) and a cold outer region (R$>$20 AU). Near-IR and mid-IR diagnostics probe the inner disk, (sub)-mm diagnostics probe the outer disk. The disk has a vertical structure: a dense mid-plane and a hotter less dense surface layer. Near-IR and mid-IR gas and dust emission features originate in the optically thin surface layer of the inner disk. Near-IR and mid-IR continuum originates from the optically thick interior layer. At (sub)-mm wavelengths we observe line emission from optically thick CO gas and optically thin emission of cold dust located in the outer disk.} \\end{figure} ", "conclusions": "In this contribution I have discussed several observational diagnostics that allow to study the gas in protoplanetary disks highlighting the strengths and limitation of each diagnostic. Here, I summarize important general points that one should bear in mind about observational constraints of gas in disks. \\begin{itemize} \\item Disks have a radial temperature structure; therefore, different gas diagnostics are employed for probing different regions of the disk. UV probes the hottest gas in the innermost disk (R$<<$1 AU, T$>$2000 K), NIR and MIR probes hot and warm gas in the inner disk (R$\\sim$0.1-20 AU, T$\\sim$100-2000 K), (sub)-mm probes cold gas in the outer disk (R$>20$ AU, T$\\sim$20-100 K). Note that disks have typical sizes of several hundreds of AU. \\item To observe a line from gas (in emission or absorption) the dust in the medium where the line is produced must be optically thin at the observed wavelength, or the gas should be at a higher temperature than the dust. \\item An optically thick disk has a vertical temperature structure. The surface layer is hotter than the interior mid-plane layer. The dust in the surface layer is optically thin, the dust in the interior layer (of the inner disk) is optically thick. In optically thick disks, hot and warm gas emission lines from the UV to the IR are produced in the surface layer. Note that the amount of mass in the surface layer is much smaller than the amount of mass in the interior layer. Gas lines in the IR probe only a limited amount of material. H$_2$ lines in the IR are optically thin. CO lines in the IR can be optically thick. \\item Sub-mm lines trace only the cold outer regions of the disk at R$>$20 AU. Planets usually are not expected to form at this distances. Still, most of the mass of the disk is at R$>$20 AU. \\item CO in the (sub)-mm is not a reliable disk gas mass tracer because (i) CO freezes onto dust grain surfaces, (ii) the CO/H$_2$ conversion factor is unknown in disks (iii) the emission is likely optically thick. \\item Be aware that so far we do not measure directly the mass of gas in disks. We deduce the disk mass from dust continuum emission in the (sub)-mm from cold dust in the outer disk (R$>$20 AU). The deduced mass depends on the dust opacity, dust temperature and the dust-to-gas ratio assumed. The disk mass observed is from material located at R$>$50 AU, and disk evolutionary models are typically of disks of R$<$50 AU. The mass and surface density profiles of gas at R$<$20 AU are poorly constrained from observations. \\item Modeling is required to interprete line observations. Conventional assumptions are that the observed gas is at LTE and is optically thin. The rotational diagrams employed for deducing the temperature of the observed gas make use those two assumptions. To deduce the inner most radius of the observed emission (i.e., R$_{\\rm in}$) from line profile modeling, it is required to assume the disk inclination and the dependence of the intensity as a function of radius ($I(R)$). Inclinations are typically deduced from (sub)-mm imaging and extended to the inner disk. On the other hand, if we want to deduce independently the inclination (e.g., to search for evidence of warps) we need to assume (or calculate) R$_{\\rm in}$ and $I(R)$. \\item Most of the gas tracers (e.g., H$_2$, [Ne II]) can be produced by shocked emission from outflows as well. High spectral/spatial resolution is required for determine whether the emission is observed at the velocity of the central star and whether it is spatially extended or not. \\item Gas emission could be excited by collisions with dust grains, shocks and UV or X-rays. Line ratios are generally employed to discriminate between the different excitation mechanisms. \\item Gas could be present in a disk even when dust emission is weak or not present. Gas emission lines have been observed in the inner regions of transitional disks and in several WTTS. \\item The study of gas in disks acquired momentum with the advent of high-resolution ground IR spectrographs and high-sensitivity IR spectrographs in space. \\item A new chapter in gas studies was recently written with the detection of emission of simple molecules in the inner disks. \\end{itemize} At the beginning of the paper I highlighted some fundamental science questions that required direct observational constraints from the gas in the disk. To conclude, I would like to address them again from the actual state of the observational study of gas. {\\it (i) How long does the protoplanetary disk last? } In general terms, disk gas diagnostics are not detected when the signatures of the dust in the disk have disappeared. Thus, as a first approximation the gas in the disk disappears almost simultaneously with the signatures from small dust particles. Nonetheless, we should be aware of a few details: (a) with the exception of H$\\alpha$ emission, there are many sources for which we know that they have a disk, but given sensitivity issues, we are not able to detect any gas tracer; (b) gas has been observed in several WTTS stars (e.g., H$_2$ in the UV and NIR), and in transitional disks (e.g., H$_2$ in the near-IR, CO at 4.7 $\\mu$m, [Ne II] at 12 $\\mu$m); (c) unbiased surveys for hot/warm gas in a large sample of WTTS have yet not been performed; (d) surveys of H$\\alpha$ at high spectral resolution need to be done in larger samples of WTTS in star-forming regions of different ages. If the accretion rate and spectral resolution are low, one can easily miss objects accreting gas at low accretion rates. In summary, an estimate of disk gas lifetime independent of dust is still required. {\\it (ii) How much material is available for forming giant planets? } On this aspect our hopes are fainter. Several surveys for H$_2$ emission in the MIR showed that the emission is observed only in a handful of objects, and that in these objects the gas is heated by an additional (to dust collisions) heating mechanism (UV, X-ray excitation). Although these results confirm our two-layer picture of the disk, in practice these results mean that we are blind to the disk's mid-plane and that we are unable to know directly how much mass is in the giant planet forming region of the disk. The good news are two fold: (a) in objects with data at several wavelengths, an educated guess can be found from modeling the surface density (see next point); (b) at some point the disk should become optically thin -by dust evolution- then we will be able to measure the gas mass. {\\it (iii) How do the density and temperature of the disk vary as a function of the radius? } There is a growing number of sources (e.g., TW Hya, AB Aur) for which we start to have information from several gas tracers at wavelengths from the UV to the (sub)-mm. In such sources, we can attempt to constrain the disk structure employing a disk model aimed to fit all the available data. Therefore, at least for a handful of objects, by modeling multiwavelength spectroscopy, we have the potential to constrain the density and temperature as a function of radius. However, we should always keep in mind that we observe emission from the disk's surface layer and that the mid-plane temperature and density will be deduced from the model. Hence, the final gas mass deduced will be model dependent. {\\it (iv) What are the dynamics of the disk? } In this aspect we had good news. To detect gas lines in the IR we commonly use the highest spectral resolution available. At the present, the resolution is sufficiently high to spectrally resolve the lines. This allows the modeling of the line shape; therefore, constraints in the dynamics of the emitting gas can be derived. Recently, with the combination of AO and high-resolution spectroscopy, it has been possible to spatially resolve gas disk emission directly or to spatially resolve the spectroastrometric signal with spectra taken at different slit orientations. So far the observed gas is consistent with gas in Keplerian rotation. This kind of studies should be extended in the future to a larger number of systems." }, "0911/0911.5451_arXiv.txt": { "abstract": "We present the results of a search for new members of the Taurus star-forming region using data from the {\\it Spitzer Space Telescope} and the {\\it XMM-Newton Observatory}. We have obtained optical and near-infrared spectra of 44 sources that exhibit red {\\it Spitzer} colors that are indicative of stars with circumstellar disks and 51 candidate young stars that were identified by Scelsi and coworkers using {\\it XMM-Newton}. We also performed spectroscopy on four possible companions to members of Taurus that were reported by Kraus and Hillenbrand. Through these spectra, we have demonstrated the youth and membership of 41 sources, 10 of which were independently confirmed as young stars by Scelsi and coworkers. Five of the new Taurus members are likely to be brown dwarfs based on their late spectral types ($>$M6). One of the brown dwarfs has a spectral type of L0, making it the first known L-type member of Taurus and the least massive known member of the region ($M\\sim4$-7~$M_{\\rm Jup}$). Another brown dwarf exhibits a flat infrared spectral energy distribution, which indicates that it could be in the protostellar class~I stage (star+disk+envelope). Upon inspection of archival images from various observatories, we find that one of the new young stars has a large edge-on disk ($r=2\\farcs5=350$~AU). The scattered light from this disk has undergone significant variability on a time scale of days in optical images from the Canada-France-Hawaii Telescope. Using the updated census of Taurus, we have measured the initial mass function for the fields observed by {\\it XMM-Newton}. The resulting mass function is similar to previous ones that we have reported for Taurus, showing a surplus of stars at spectral types of K7-M1 (0.6-0.8~$M_\\odot$) relative to other nearby star-forming regions like IC~348, Chamaeleon~I, and the Orion Nebula Cluster. ", "introduction": "\\label{sec:intro} The Taurus complex of dark clouds is one of the best sites for studying the formation of stars in a quiescent, relatively isolated environment. It is among the nearest star-forming regions ($d=140$~pc) and exhibits a very low stellar density ($n\\sim1$-10~pc$^{-3}$). Although the individual clouds are sparsely populated, the cloud complex as a whole contains more than 300 known members. Working toward a complete census of Taurus is important for the identification of rare objects (e.g., edge-on disks, transitional disks, protostars) as well as the statistical characterization of the stellar population (e.g., disk fraction, initial mass function, spatial distribution). A variety of methods have been employed in surveys for new members of Taurus \\citep{ken08}. Two of these techniques, mid-infrared (IR) imaging and X-ray imaging, are highly complementary. Mid-IR observations can identify stars that have circumstellar disks and can penetrate the high levels of extinction that surround stars at the earliest evolutionary stages while X-ray data can uncover the diskless members of a young stellar population. Because of their excellent sensitivities and large fields of view, the {\\it Spitzer Space Telescope} \\citep{wer04} and the {\\it XMM-Newton Observatory} \\citep{jan01} are the best available telescopes for wide-field imaging surveys at mid-IR and X-ray wavelengths, respectively. The unique capabilities of these facilities have been applied to the Taurus star-forming region through the Taurus {\\it Spitzer} Legacy Survey (D. Padgett, in preparation) and the {\\it XMM-Newton} Extended Survey of the Taurus Molecular Cloud \\citep[XEST,][]{gud07}. \\citet{luh06tau2,luh09fu} and \\citet{sce07,sce08} have used the data from these surveys to search for new members of Taurus. We have continued those efforts by performing spectroscopy on IR sources that we have identified in the {\\it Spitzer} images and X-ray sources that were reported by \\citet{sce07}. In this paper, we describe the selection of these candidate members of Taurus (\\S~\\ref{sec:select}) and measure their spectral types with optical and IR spectra (\\S~\\ref{sec:spec}). We then characterize the stellar parameters of the new members and discuss other notable properties of these objects (\\S~\\ref{sec:prop}). Finally, we use our updated census of the stellar population in Taurus to measure the initial mass function (IMF) within the fields observed by XEST (\\S~\\ref{sec:imf}). ", "conclusions": "We have presented a survey for new members of the Taurus star-forming region in which we obtained spectra of candidate members appearing in images from the {\\it Spitzer Space Telescope} (46~deg$^{2}$) and the {\\it XMM-Newton Observatory} (5~deg$^{2}$). Using the mid-IR data from {\\it Spitzer}, we identified 44 sources that could be young stars with disks, 24 of which were confirmed as members by our spectroscopy. We also performed spectroscopy on 51 candidates detected in X-rays by the XEST program \\citep{gud07,sce07}, demonstrating the youth and membership of 16 sources. Ten of these new X-ray members were independently confirmed through spectroscopy by \\citet{sce08}. In addition, to the sources from {\\it Spitzer} and {\\it XMM-Newton}, we observed four candidate companions to known members of Taurus that were found by \\citet{kra07} through analysis of 2MASS data, one of which we have classified as a young star. Our survey has uncovered several rare types of sources that are valuable for studies of various aspects of star and planet formation. They consist of a wide binary brown dwarf that is forming in isolation \\citep{luh09fu}, the first known L-type member of Taurus ($M\\sim4$-7~$M_{\\rm Jup}$), a highly reddened brown dwarf that may be in the class~I stage (star+disk+envelope), a disk that appears to have an inner hole (i.e., transitional disk), and a large edge-on disk ($r=2\\farcs5=350$~AU). The companion identified by \\citet{kra07} also may comprise the primary in wide, low-mass binary system (M4.5+M8.5, $a=12\\arcsec=1700$~AU). Meanwhile, the {\\it Spitzer} and {\\it XMM-Newton} data in conjunction with previous optical and near-IR surveys provide relatively well-defined completeness limits for the current census of Taurus, enabling a better characterization of the stellar population. For instance, we have estimated the IMF within the fields observed by XEST, arriving at a distribution that reaches a maximum near 0.8~$M_\\odot$, which agrees with our previous measurements for Taurus. Thus, the IMF in Taurus continues to appear anomalous compared to other nearby star-forming clusters, which peak at 0.1-0.2~$M_\\odot$. The disk fraction in the XEST fields and the spatial distribution of the SED classes are investigated by \\citet{luh09tau}. The completeness of the census of Taurus remains poorly determined among class~I sources at low masses and class~III sources outside of the XEST fields, which focused on the denser stellar aggregates. Future surveys can address these shortcomings through spectroscopy of red, faint sources detected by {\\it Spitzer} and measurements of variability and proper motions with wide-field, multi-epoch imaging (e.g., Panoramic Survey Telescope and Rapid Response System, Large Synoptic Survey Telescope)." }, "0911/0911.3041_arXiv.txt": { "abstract": "Semi-empirical method of calculation of quenching factors for scintillators is described. It is based on classical Birks formula with the total stopping powers for electrons and ions which are calculated with the ESTAR and SRIM codes, respectively. Method has only one fitting parameter (the Birks factor $kB$) which can have different values for the same material in different conditions of measurements and data treatment. A hypothesis is used that, once the $kB$ value is obtained by fitting data for particles of one kind and in some energy region (e.g. for a few MeV $\\alpha$ particles from internal contamination of a detector), it can be applied to calculate quenching factors for particles of another kind and for another energies (e.g. for low energy nuclear recoils) if all data are measured in the same experimental conditions and are treated in the same way. Applicability of the method is demonstrated on many examples including materials with different mechanisms of scintillation: organic scintillators (solid C$_8$H$_8$, and liquid C$_{16}$H$_{18}$, C$_9$H$_{12}$); crystal scintillators (pure CdWO$_4$, PbWO$_4$, ZnWO$_4$, CaWO$_4$, CeF$_3$, and doped CaF$_2$(Eu), CsI(Tl), CsI(Na), NaI(Tl)); liquid noble gases (LXe). Estimations of quenching factors for nuclear recoils are also given for some scintillators where experimental data are absent (CdWO$_4$, PbWO$_4$, CeF$_3$, Bi$_4$Ge$_3$O$_{12}$, LiF, ZnSe). ", "introduction": "In accordance with our current understanding of astronomical observations, usual matter constitutes only $\\simeq4$\\% of the Universe; the main components are dark matter ($\\simeq23$\\%) and dark energy ($\\simeq73$\\%) \\cite{Ber05}. Various extensions of the Standard Model propose many candidates on the role of dark matter (DM) particles \\cite{Ste09} which are neutral and only weakly interact with matter (Weakly Interacting Massive Particles, WIMPs). One of the approaches to discover these particles is to detect scattering of WIMPs on atomic nuclei in sensitive detectors placed deep underground and measured in extra low background conditions \\cite{Spo07}. Taking into account likely mass range and velocities of WIMPs, energies of nuclear recoils are expected below $\\simeq100$ keV with character interaction rates of $1-10^{-6}$ events kg$^{-1}$ d$^{-1}$. Many searches of WIMPs with semiconductor, scintillator and bolometer detectors to-date gave only negative results (see \\cite{Spo07} and references therein); instead positive evidence for DM particles (WIMPs are a subclass; other candidates and other kinds of interactions are also available) in the galactic halo has been pointed out by DAMA experiments by exploiting the DM annual modulation signature with NaI(Tl) scintillators during more than 10 years long measurements \\cite{Ber08a}. For a long time it is known that amount of light produced in scintillating material by highly ionizing particles is lower than that produced by electrons of the same energy \\cite{Bir64}. Thus, in a scintillator calibrated with electron and/or $\\gamma$ sources (which is an usual practice), signals from ions will be seen at lower energies (sometimes up to $\\simeq 40$ times) than their real values. Evidently knowledge of these transformation coefficients -- quenching factors -- is extremely important in searches for WIMPs and in predictions where the WIMPs signal should be expected. Many experimental efforts were devoted to measurements, sometimes very sophisticated, of quenching factors at low energies in different detectors (see e.g. recent works \\cite{Bav08,Cal08,Cha08,Cor08,Kob08,Nik08,Sor09} and further references). Quenching factors are also needed in measurements and interpretation of signals from $\\alpha$ particles in scintillators. As examples, we can mention here recent experiments on searches (and first observations) of extremely rare $\\alpha$ decays ($T_{1/2}=10^{18}-10^{19}$ yr): $^{180}$W in CdWO$_4$ \\cite{Dan03} and CaWO$_4$ \\cite{Zde05} crystal scintillators, and $^{151}$Eu in CaF$_2$(Eu) \\cite{Bel07}. While few approaches in calculation of quenching factors are known \\cite{Bir51,Mur61,Lin63,Hit05}, satisfactory theory able to exactly predict (and very often even to describe already measured) quenching factors for all detectors and particles still is absent. For example, in the Lindhard's approach \\cite{Lin63}\\footnote{Good description is given in more accessible source \\cite{Mei08}.} it is possible to calculate quenching of ions with atomic number $Z$ in scintillator only with the same $Z$ number; in addition, this theory predicts decrease of quenching factors at low energies, very often in contradiction with experimental data. Hitachi's model \\cite{Hit05} gives better description and for wider data range, however it is not easy to reproduce these calculations independently. Below we describe rather simple method of calculation of quenching factors for different ions (from protons to heavy recoils), based on semi-empirical approach of Birks \\cite{Bir51} and using available in Internet software for calculation of stopping powers for electrons and ions (ESTAR \\cite{ESTAR} and SRIM \\cite{SRIM} codes, respectively). It employs only one parameter ($kB$ Birks factor) which could be found by fitting experimental data measured for particles of one kind in some energy region (e.g. for $\\alpha$ particles from external sources or internal contamination of a detector by U/Th chains, $^{147}$Sm, $^{190}$Pt, etc.) but afterwards can be used to calculate quenching factors for other particles and in other energy regions (e.g. for nuclear recoils at low energies). Summary of the method is given in section 2. Calculations with this method are demonstrated in section 3 for number of scintillators: organic scintillators (solid C$_8$H$_8$, and liquid C$_{16}$H$_{18}$, C$_9$H$_{12}$); crystal scintillators (pure CdWO$_4$, PbWO$_4$, ZnWO$_4$, CaWO$_4$, CeF$_3$, and doped CaF$_2$(Eu), CsI(Tl), CsI(Na), NaI(Tl)); liquid noble gases (LXe). Estimations of quenching factors for nuclear recoils are also given for some scintillators where experimental data are absent (CdWO$_4$, PbWO$_4$, CeF$_3$, Bi$_4$Ge$_3$O$_{12}$, LiF, ZnSe). Section 4 gives conclusions. ", "conclusions": "Semi-empirical and quite simple in realization method of calculation of quenching factors for scintillators was described in this work. It is based on the classical Birks formula with the {\\em total} stopping powers for electrons and ions, and has only one parameter: the Birks factor $kB$. Value of this factor for a given scintillating material can be different in different conditions of measurements and data treatment. However, if experimental conditions and treatment of data are fixed, hypothesis that $kB$ has the same value for particles of different kinds gives reliable results. Once the $kB$ is found by fitting quenching factors for particles of one kind and in some range of energies (e.g. for $\\alpha$ particles from internal contamination of a detector by U/Th chains and/or by $^{147}$Sm, $^{190}$Pt with energies of a few MeV), it can be used to calculate quenching factors for particles of another kinds and for another energies of interest (e.g. for low energy nuclear recoils). Many examples were given for materials which, furthermore, have different mechanisms of scintillation: organic scintillators (solid C$_8$H$_8$, and liquid C$_{16}$H$_{18}$, C$_9$H$_{12}$); crystal scintillators (pure CdWO$_4$, PbWO$_4$, ZnWO$_4$, CaWO$_4$, CeF$_3$, and doped CaF$_2$(Eu), CsI(Tl), CsI(Na), NaI(Tl)); and liquid noble gases (LXe). It was demonstrated for many cases that the method allows not only to {\\em describe} measured data for ions of one kind in a reliable way but also to {\\em predict} behaviour of quenching factors for other particles which sometimes is immediately confirmed by already existing experimental data -- sometimes worse, sometimes better, and sometimes very good, but at least in a rough agreement. Some predictions (e.g. for LNe, LiF and others) could be checked in near future. Stopping powers for electrons and ions are calculated with the ESTAR and SRIM codes, respectively, which in fact present to-date state-of-art software in this field. It is easy to use these programs and they are publicly available; this makes $Q_i$ calculations quite simple. Calculations with the SRIM package have some tendency to overestimate quenching factors for $\\alpha$ particles at energies around $\\simeq2$ MeV and underestimate them at high energies ($>8$ MeV) as can be seen in Fig. 4a for CdWO$_4$, Fig. 5a for CaF$_2$(Eu), Fig. 8 for CaWO$_4$, and Fig. 15 for CeF$_3$. At the same time, calculation of the stopping powers for $\\alpha$ particles with the ASTAR package gave better description of $Q_\\alpha$ in CaF$_2$(Eu) scintillator at lower energies (see Fig. 5a). For some other materials difference between ASTAR and SRIM calculations was not big (see Fig. 2 for CsI(Tl) and Fig. 3a for C$_8$H$_8$). Evidently $Q_i$ values will depend on how one calculates stopping powers for ions and electrons, and it is a pity that stopping powers could not be computed in framework of the same package for any particle (SRIM calculates SP for ions in any substance but does not calculate SP for electrons; and STAR gives SP for electrons in any material but SP for ions are possible only for protons and $\\alpha$ particles and for a limited list of materials). Quenching factors calculated in the presented approach in general increase at low energies, and this encourages experimental searches for dark matter particles. Estimations of quenching factors for nuclear recoils are given for some scintillators where experimental data are absent (CdWO$_4$, PbWO$_4$, CeF$_3$, Bi$_4$Ge$_3$O$_{12}$, LiF, ZnSe)." }, "0911/0911.4457_arXiv.txt": { "abstract": "{Supernova remnants are believed to be a major source of energetic particles (\\emph{cosmic rays}) on the Galactic scale. Since their progenitors, namely the most massive stars, are commonly found clustered in \\emph{OB~associations}, one has to consider the possibility of collective effects in the acceleration process.} {We investigate the shape of the spectrum of high-energy protons produced inside the \\emph{superbubbles} blown around clusters of massive stars.} {We embed simple semi-analytical models of particle acceleration and transport inside Monte Carlo simulations of OB~associations timelines. We consider regular acceleration (Fermi~1 process) at the shock front of supernova remnants, as well as stochastic reacceleration (Fermi~2 process) and escape (controlled by magnetic turbulence) occurring between the shocks. In this first attempt, we limit ourselves to linear acceleration by strong shocks and neglect proton energy losses.} {We observe that particle spectra, although highly variable, have a distinctive shape because of the competition between acceleration and escape: they are harder at the lowest energies (index $s<4$) and softer at the highest energies ($s>4$). The momentum at which this spectral break occurs depends on the various bubble parameters, but all their effects can be summarized by a single dimensionless parameter, which we evaluate for a selection of massive star regions in the Galaxy and the LMC.} {The behaviour of a superbubble in terms of particle acceleration critically depends on the magnetic turbulence: if~B is low then the superbubble is simply the host of a collection of individual supernovae shocks, but if~B is high enough (and the turbulence index is not too high), then the superbubble acts as a global accelerator, producing distinctive spectra, that are potentially very hard over a wide range of energies, which has important implications on the high-energy emission from these objects.} ", "introduction": "\\label{sec:introduction} Superbubbles are hot and tenuous large structures that are formed around OB~associations by the powerful winds and the explosions of massive stars \\citep{Higdon2005a}. They are the major hosts of supernovae in the Galaxy, and thus major candidates for the production of energetic particles (e.g., \\citealt{Montmerle1979a}, \\citealt{Bykov2001b}, \\citealt{Butt2009a}, and references therein). Supernovae are indeed believed to be the main contributors of Galactic cosmic rays (along with pulsars and micro-quasars), by means of the \\emph{diffusive shock acceleration} process (a 1st-order, regular Fermi process) occurring at the remnant's blast wave as it goes through the interstellar medium \\citep{Drury1983a,Malkov2001c}. Supernovae in superbubbles are correlated in space and time, hence the need to investigate acceleration by multiple shocks \\citep{Parizot2004a}. \\citet{Klepach2000a} developed a semi-analytical model of test-particle acceleration by multiple spherical shocks (either wind termination shocks, or supernova shocks plus wind external shocks), based on the limiting assumption of small shocks filling factors. \\citet{Ferrand2008a} performed direct numerical simulations of repeated acceleration by successive planar shocks in the non-linear regime (that is, taking into account the back-reaction of energetic particles on the shocks). However, to ascertain the particle spectrum produced inside the superbubble as a whole, one must also consider important physics occurring \\emph{between} the shocks. Since the bubble interior is probably magnetized and turbulent, we need to evaluate gains and losses caused by the acceleration by waves (a 2nd-order, stochastic Fermi process) and escape from the bubble. In this study, we combine the effects of regular acceleration (occurring quite discreetly, at shock fronts) and stochastic acceleration and escape (occurring continuously, between shocks), to determine the typical spectra that we can expect inside superbubbles over the lifetime of an OB~cluster. We choose to treat regular acceleration as simply as we can, and concentrate on modeling the relevant scales of stochastic acceleration and escape inside superbubbles. We present our model in Sect.~\\ref{sec:model}, give our general results in Sect.~\\ref{sec:results}, and present specific applications in Sect.~\\ref{sec:application}. Finally we discuss the limitations of our approach in Sect.~\\ref{sec:limitations} and provide our conclusions in Sect.~\\ref{sec:conclusions}. ", "conclusions": "\\label{sec:conclusions} Our main conclusions are as follows: \\begin{enumerate} \\item Cosmic-ray spectra inside superbubbles are highly variable: at a given time they depend on the particular history of a given cluster. \\item Nevertheless, spectra follow a distinctive overall trend, produced by a competition between (re-)acceleration by regular and stochastic Fermi processes and escape: they are harder at lower energies ($s<4$) and softer at higher energies ($s>4$), shapes that are in agreement with the results of~\\citet{Bykov2001b} based on different assumptions\\footnote{ \\cite{Bykov2001b} considers acceleration of particles by large-scale motions of the magnetized plasma inside the superbubble, which depends on the ratio $D_u/D_x$ where $D_x$ is the space diffusion coefficient, controlled by magnetic fluctuations at small scales, and $D_u=UL$ describes the effect of large scale turbulence, where $U$ is the average turbulent speed and $L$ is the average size between turbulence sources.}. \\item The momentum at which this spectral break occurs critically depends on the bubble parameters: it increases when the magnetic field value and acceleration region size increase, and decreases when the density and the turbulence external scale increase, all these effects being summarized by the single dimensionless parameter $\\theta^{\\star}$ defined by Eq.~(\\ref{eq:Theta_def}). \\item For reasonable values of superbubble parameters, very hard spectra ($s<3$) can be obtained over a wide range of energies, provided that superbubbles are highly magnetized and turbulent (which is a debated issue). \\end{enumerate} These results have important implications for the chemistry inside superbubbles and the high-energy emission from these objects. For instance, in the superbubble Perseus~OB2 there is observational evidence of intense spallation activity \\citep{Knauth2000a} attributed to a high density of low-energy cosmic rays, but EGRET has not detected $\\pi^{0}$-decay radiation, which places strong limits on the density of high-energy cosmic rays. This is consistent with the shape of the spectra obtained in this work. We are thus looking forward to seeing how new instruments such as Fermi and AGILE will perform on extended sources such as massive star forming regions, which have recently been established as very high-energy sources. In that respect, we make a final comment that the high intermittency of predicted spectra might explain the puzzling fact that some objects are detected while others remain unseen." }, "0911/0911.1791_arXiv.txt": { "abstract": "We present results from the first systematic search for outlying HII regions, as part of a sample of 96 emission-line point sources (referred to as ELdots - emission-line dots) derived from the NOAO Survey for Ionization in Neutral Gas Galaxies (SINGG). Our automated ELdot-finder searches SINGG narrow-band and continuum images for high equivalent width point sources outside the optical radius of the target galaxy ($>$ 2 $\\times$ r$_{25}$ in the R-band). Follow-up longslit spectroscopy and deep GALEX images (exposure time $>$ 1000 s) distinguish outlying HII regions from background galaxies whose strong emission lines ([OIII], H$\\beta$ or [OII]) have been redshifted into the SINGG bandpass. We find that these deep GALEX images can serve as a substitute for spectroscopic follow-up because outlying HII regions separate cleanly from background galaxies in color-color space. We identify seven SINGG systems with outlying massive star formation that span a large range in H$\\alpha$ luminosities corresponding to a few O stars in the most nearby cases, and unresolved dwarf satellite companion galaxies in the most distant cases. Six of these seven systems feature galaxies with nearby companions or interacting galaxies. Furthermore, our results indicate that some outlying HII regions are linked to the extended-UV disks discovered by GALEX, representing emission from the most massive O stars among a more abundant population of lower mass (or older) star clusters. The overall frequency of outlying HII regions in this sample of gas-rich galaxies is 8 - 11\\% when we correct for background emission-line galaxy contamination ($\\sim75$\\% of ELdots). ", "introduction": "\\label{intro} In the last 10 years, deep H$\\alpha$ imaging has confirmed the occurrence of recent star formation far beyond the main optical bodies of galaxies, in both discrete knots with no underlying continuum emission and faint outer spiral arms (eg. \\nocite{ferguson98, emma,bfq97} Bland-Hawthorn, Freeman, \\& Quinn 1997; Ferguson et al. 1998; Ryan-Weber et al. 2004). More recently, the {\\it{Galaxy Evolution Explorer}} (GALEX) satellite has broadened our picture of outer-galaxy star formation by revealing that over $\\sim30$\\% of spiral galaxies possess UV-bright extensions of their optical disks, implying that low-intensity star formation in the outskirts of galaxies is not particularly rare \\citep{Thilker07}. This mode of star formation in the outer disks of galaxies may be a natural extension of inside-out disk formation, in which galactic stellar disks grow gradually with time following the condensation and assembly of their gaseous disks (\\nocite{whiteandfrenk91, bouwens97} e.g. White \\& Frenk 1991; Bouwens, Cayon, \\& Silk 1997). Moreover, extended-UV (XUV) emission and outlying HII regions sometimes are associated with previous or ongoing galaxy interactions \\citep{Thilker07, werk08}. Probing the nature and frequency of outer-galaxy star formation can facilitate our understanding of the disk building process, and the ways in which new stellar populations emerge via galaxy interactions. Star formation in galactic outskirts also has relevance to discussions involving star formation gas density thresholds, as it often occurs in a low-density environment (e.g. \\nocite{gerhard,mendes04, emma} Gerhard et al. 2002; Mendes de Oliveira et al. 2004; Ryan-Weber et al. 2004). One way to increase our understanding of star formation ``thresholds\" is to assemble a sample of massive stars forming in low column-density gas outside the influence of the existing stellar population in a galactic disk. The \\ha~ and UV radiation in the extended disks trace different ranges of stellar mass: UV radiation follows primarily the O and B stars, while \\ha~ emission traces just the O stars. Studies of spiral galaxies with XUV disks \\citep{thilker05, gildepaz05} have indicated that the spatial extent of star formation in outer disks is underestimated by looking for HII regions alone (via \\ha~ imaging), as \\ha~emission tends to be even less widespread than predicted by typical stellar population synthesis models that incorporate a power-law IMF with a slope of $\\alpha\\sim2.35$. The reasons for this relative lack of \\ha~emission at large galactocentric radii could be many, including but not limited to stochastic fluctuations in the IMF at low stellar luminosities \\citep{boissier06}, a top-light IMF in the remotest reaches of galaxies \\citep{meurer09}, or a large fraction of escaping ionizing photons from the less dense outer-galactic regions. The extent of this underestimation, specifically at large projected galactocentric distances, remains unquantified, as no comprehensive systematic study of \\ha~emission in the outer reaches of galaxies has yet emerged. Here, we return to \\ha~imaging to find star formation in the outskirts of galaxies. As a star formation tracer that is sensitive to only the highest mass stars, \\ha~provides an important complement to the recent GALEX results. We perform an automated search to find outlying compact sources of net line emission using the imaging data from the Survey for Ionization in Neutral Gas Galaxies (SINGG; Meurer et al 2006; hereafter M06\\nocite{SINGG}). The large angular area outside the optical radius of galaxies in each 14.7\\'~SINGG field presents the opportunity to search for outlying HII regions at large radii and perform a blind search for background emission-line sources. We initially refer to both types of sources as ``ELdots\" for their appearance as emission line dots in the images. In all cases, ELdots exhibit strong emission lines in the SINGG narrowband filter, and are point sources well outside the broadband optical emission of nearby galaxies (r$>$2 $\\times$ r$_{25}$). ELdots can be outlying \\ha-emitting HII regions at a similar velocity to the HIPASS source galaxy or background galaxies emitting a different line ([OIII], [OII], or H$\\beta$) that is redshifted into the narrow filter passband used for the observations. To distinguish between these options, we present follow-up spectroscopy and deep archival GALEX images for a subsample of ELdots. In this study, we are primarily interested in the \\ha-emitting ELdots, although we tabulate the properties of the background galaxies as well. While a number of previous works have discovered ``intergalactic\" HII regions, this is the first systematic search of an unbiased sample of gas rich galaxies for such emission. Furthermore, although many studies have chosen ``intergalactic\" as a modifier (see \\nocite{boquien09} Boquien et al. 2009 for a detailed description of the terminology), we opt instead for ``outlying.\" These HII regions are distinct from the central galactic star formation, and more sparse, yet they often lie in extended neutral gas and/or an extended UV component associated with the galaxy. Outlying HII regions (abbreviated outer-HIIs) provide a unique laboratory to understand star formation under relatively extreme conditions (e.g. low neutral gas column density, weak galaxy potential), and may shed light on the full extent of stellar disks \\citep{blandhawthorn05, irwin05}. The paper proceeds as follows: in Section 2, we describe the SINGG sample; in Section 3, we describe our ELdot sample selection; in Section 4, we present spectroscopic observations of a subsample of ELdots; in Section 5 we examine the ELdots in deep GALEX images, where available from the archives; in Section 6, we discuss the properties of our ELdot sample; in Section 7, we discuss extended star formation in SINGG; and in Section 8 we summarize and present the key findings of this study. ", "conclusions": "\\label{conc} ELdots are emission-line point sources (``dots\") well outside the main optical R-band emission of a galaxy, defined as 2 $\\times$ r$_{25}$ in this work. Using an automated finder, we catalogue a total of 96 ELdots in 50 of 89 SINGG systems. We classify ELdots as either outlying HII regions (outer-HIIs) or background galaxies by spectroscopy and GALEX morphology and colors. This study highlights four key results: \\begin{itemize} \\item{Follow-up GALEX data, combined with SINGG photometric properties, can help distinguish between outer-HIIs and background galaxy ELdots, nearly eliminating the need for spectroscopic follow-up. The distribution of ELdot FUV$-$R colors is bimodal, revealing that outer-HIIs are considerably bluer than higher-redshift emission-line background galaxies. The very blue outer-HII colors are due primarily to the extremely low R-band emission from these objects, indicating new stars forming where there is little or no underlying older stellar population. Outer-HIIs also tend to have higher emission-line EWs than background galaxy ELdots and are likely to be found closer to a potential host galaxy than the background galaxies, which are uniformly distributed across the SINGG search area.} \\item{Through our systematic search of a sample of gas-rich galaxies, we have confirmed that massive star formation in the far outskirts of galaxies tends to be associated with galaxy interactions and nearby companions. This result agrees with other recent findings of star formation outside the main optical bodies of galaxies. } \\item{ In the cases where long-exposure GALEX data is available, we find that outer-HIIs appear to be associated with XUV emission, specifically of Type-1 morphology. } \\item{ We find that the frequency of massive star formation in the far outskirts of galaxies, as traced by \\ha~emission in the full SR1 sample, is between 8 and 11\\%. Outer-HII regions in this full sample span a wide range of H$\\alpha$ luminosities, from 10$^{36.6}$ ergs s$^{-1}$ to 10$^{38.5}$ ergs s$^{-1}$, and have physical sizes from 70 pc to 500 pc, representing OB associations ionized by a few O stars in the most nearby systems, and dwarf galaxies with hundreds of O stars in the most distant systems. When we isolate our sample of outer-HIIs to a nearby sample with D $<$ 30 Mpc, we can compare our frequency statistics more directly with those of XUV galaxies. These outer-HIIs are more homogeneous in nature, similar in size and luminosity to Galactic OB associations. Where \\cite{Thilker07} find XUV emission to be present in $\\sim30$\\% of spiral galaxies in the local universe, we find the frequency of outlying HII regions in gas-rich galaxies more nearby than 30 Mpc to be between 6 and 10\\%. } \\end{itemize}" }, "0911/0911.4382_arXiv.txt": { "abstract": "We present a direct $N$-body simulation modeling the evolution of the old (7 Gyr) open cluster NGC 188. This is the first $N$-body open cluster simulation whose initial binary population is directly defined by observations of a specific open cluster: M35 (150 Myr). We compare the simulated color-magnitude diagram at 7 Gyr to that of NGC 188, and discuss the blue stragglers produced in the simulation. We compare the solar-type main sequence binary period and eccentricity distributions of the simulation to detailed observations of similar binaries in NGC 188. These results demonstrate the importance of detailed observations in guiding $N$-body open cluster simulations. Finally, we discuss the implications of a few discrepancies between the NGC 188 model and observations and suggest a few methods for bringing $N$-body open cluster simulations into better agreement with observations. ", "introduction": "\\label{intro} Recently, sophisticated $N$-body simulations have incorporated stellar dynamics and stellar evolution self consistently, gaining the ability to accurately model open cluster sized populations ($N \\sim 10^4$) of single and binary stars through many billions of years. (e.g. \\texttt{NBODY6}, \\citealt{aarseth:03}, with stellar and binary evolution included by \\citealt{hurley:00,hurley:02}). Concurrently, large observational surveys of open clusters, like the WIYN Open Cluster Study \\citep[WOCS;][]{mathieu:00}, are coming to maturity, providing comprehensive databases of open cluster characteristics, including detailed information on their binary populations. These observations, and specifically those of the binaries, provide important tests and guidance for $N$-body simulations. In this paper, we describe our first step in utilizing this wealth of data to help guide an $N$-body simulation with the aim of creating a realistic model of NGC 188. ", "conclusions": "\\label{disc} By utilizing the observed M35 binaries as a proxy for our primordial binary population, we have managed to nearly reproduce the NGC 188 binaries within our $N$-body simulation at 7 Gyr. However, as noted in Section~\\ref{comp188} the current NGC 188 model is deficient in BSs. Interestingly, the \\citet{hurley:05} model of M67 was able to produce the large number of observed BSs (though not their binary frequency), but was unable to reproduce the observed binary population. The quantity of BSs produced in the M67 model was largely dependant on the initial binary population. Hurley et al.~found that a large fraction of short-period binaries was necessary to create the observed numbers of BSs in M67 (and NGC 188). These short-period binaries not only have a higher likelihood of creating BSs in isolation (e.g., through mass transfer resulting from Roche lobe overflow), but dynamical interactions involving short-period binaries have a higher probability of resulting in collisions and mergers \\citep{fregeau:04}. Observations of M35 binaries are very different from the flat period distribution and large binary frequency used in the M67 model (e.g., Figure~\\ref{M35fig}). The important next step will be to match both the binary population and the BSs population within one cohesive simulation. We are currently exploring a number of possible methods for addressing this issue, the most promising of which appears to be the inclusion of primordial triples (e.g., Kozai induced mergers; \\citealt{perets:09}, larger cross section for dynamical interactions, etc.). \\begin{figure}% \\begin{center} \\includegraphics[width=0.8\\linewidth]{Geller_f4.eps} \\caption{\\label{NGC188fig3} Eccentricity plotted against the logarithm of the period ($e$ - log $P$) at 7 Gyr for the observed (top) and simulated (bottom) solar-type MS NGC 188 binaries. The solid gray lines represent the best-fit circularization functions from \\citet{meibom:05} for each sample, respectively. The tidal processes in the simulation result in a circularization period that is significantly lower than in NGC 188. Also, the simulation produces a population of circular binaries with $P >> P_{circ}$ that is not observed in NGC 188 (or the Galactic field). Most of these circular binaries in the model have previously gone through a stage of mass transfer.} \\end{center} \\end{figure} The tidal circularization period is one of our best observational tools to study the effects of tides on a binary population. In Figure~\\ref{NGC188fig3} and Section~\\ref{comp188} we show that the circularization period at 7 Gyr in the simulation is significantly lower than what is observed in NGC 188. This suggests that the tidal energy dissipation rate may be \\textit{underestimated} in the $N$-body model (at least for solar-type MS binaries). We may also (or alternatively) require a form of pre-MS tidal circularization \\citep[e.g.][]{kroupa:95}. As pointed out by \\citet{mathieu:08}, tidal processes have a significant impact on dynamical interactions between single and binary stars, and the energy dissipation rates can be the deciding factor between a close fly-by, tidal capture, or even a merger or collision that could result in a BS. WOCS data \\citep[e.g.,][]{meibom:05} provide the ideal foundation from which we will improve the tidal physics within the $N$-body model. Finally, the simulation has an excess of circular binaries with periods beyond the circularization period. The majority of these binaries have undergone some form of mass transfer, which has quickly circularized the orbit within the simulation. However, similar populations of circular binaries at $P >> P_{circ}$ are not observed in open clusters or in the Galactic field \\citep[e.g.,][]{meibom:05,duquennoy:91}. This may suggest that mass transfer, especially in initially eccentric binaries, does not necessarily lead to circular binaries \\citep[e.g.,][]{sepinsky:07,bonacic:08}, or that these cases of mass transfer happen much more frequently within the simulation than in reality. Correctly modeling the products of mass transfer and common-envelope evolution is critical for accurately reproducing BSs as well as the binary population within an $N$-body simulation. We will present further details about this NGC 188 model in future papers. Additionally, we will explore the various methods mentioned above for bringing the NGC 188 model, and $N$-body simulations in general, in better agreement with observations. This synergy between observations and simulations will continue to refine the $N$-body method and reveal further insights into the origins of BSs, the evolution of a binary population, and the dynamical evolution of star clusters. \\vspace{1em} \\footnotesize \\flushleft This work was funded in part by a National Science Foundation (NSF) East Asia and Pacific Summer Institutes (EAPSI) fellowship, the Wisconsin Space Grant Consortium, and NSF grant AST-0406615. \\vspace{-1em}" }, "0911/0911.0381_arXiv.txt": { "abstract": "{We report the results of an optical campaign carried out by the XMM-Newton Survey Science Centre with the specific goal of identifying the brightest X-ray sources in the XMM-Newton Galactic Plane Survey of Hands et al. (2004). In addition to photometric and spectroscopic observations obtained at the ESO-VLT and ESO-3.6m, we used cross-correlations with the 2XMMi, USNO-B1.0, 2MASS and GLIMPSE catalogues to progress the identification process. Active coronae account for 16 of the 30 positively or tentatively identified X-ray sources and exhibit the softest X-ray spectra. Many of the identified hard X-ray sources are associated with massive stars, possibly in binary systems and emitting at intermediate X-ray luminosities of 10$^{32-34}$\\ergs. Among these are a very absorbed likely hyper-luminous star with X-ray/optical spectra and luminosities comparable with those of $\\eta$ Carina, a new X-ray selected WN8 Wolf-Rayet star in which most of the X-ray emission probably arises from wind collision in a binary, a new Be/X-ray star belonging to the growing class of $\\gamma$-Cas analogs and a possible supergiant X-ray binary of the kind discovered recently by INTEGRAL. One of the sources, XGPS-25, has a counterpart of moderate optical luminosity which exhibits He{\\sc II} $\\lambda$4686 and Bowen C{\\sc III}-N{\\sc III} emission lines suggesting that this may be a quiescent or X-ray shielded Low Mass X-ray Binary, although its X-ray properties might also be consistent with a rare kind of cataclysmic variable (CV). We also report the discovery of three new CVs, one of which is a likely magnetic system displaying strong X-ray variability. The soft (0.4--2.0\\,keV) band \\lnls\\ curve is completely dominated by active stars in the flux range of 1$\\times$10$^{-13}$ to 1$\\times$10$^{-14}$\\ergscm . Several active coronae are also detected above 2\\,keV suggesting that the population of RS CVn binaries significantly contributes to the hard X-ray source population. In total, we are able to identify a large fraction of the hard (2--10\\,keV) X-ray sources in the flux range of 1$\\times$10$^{-12}$ to 1$\\times$10$^{-13}$\\ergscm\\ with Galactic objects at a rate consistent with that expected for the Galactic contribution only.} ", "introduction": "The brightest X-ray sources discovered by the early collimator based experiments in operation in the 1970s were very luminous ($\\sim$ 10$^{38}$\\ergs ) accretion-powered X-ray binaries located in our Galaxy \\citep[see e.g. the fourth Uhuru catalogue of X-ray sources;][] {forman1978}. The dramatic improvement in sensitivity and spatial resolution afforded by the next generation of instruments exploiting X-ray imaging revealed a Galactic landscape with much larger numbers of lower luminosity systems powered by a wide range of physical processes. For example, observations first with Einstein, and later with ROSAT, showed that low Galactic latitude regions are crowded with a large number of low-luminosity sources in the X-ray bands explored by these two missions (0.5--4.5\\,keV and 0.2--2.4\\,keV for Einstein and ROSAT respectively). These soft sources are predominantly identified with active stellar coronae \\citep[e.g.][]{hertz1984,motch1997}. The first imaging hard (2--10\\,keV) X-ray survey of the Galactic Plane was performed a few years later with ASCA \\citep{sugizaki2001}. ASCA revealed the presence of a genuinely Galactic population of low-luminosity non-coronal X-ray sources which emit hard X-rays and are therefore detectable through large columns of interstellar absorption. However, the still relatively large positional errors affecting ASCA sources combined with the effects of the strong interstellar extinction in the Galactic Plane prevented their optical identification in most cases. The prospects for building a detailed census of the low to intermediate X-ray luminosity sources has now greatly improved with the launch of the Chandra and XMM-Newton observatories. Thanks to their unprecedented sensitivity and spatial resolution, these observatories are able to study the properties of the faint X-ray source populations throughout the Galaxy and most notably those present in large numbers in the central parts of the Galaxy \\citep{wang2002,muno2003}. The fields-of-view of the Chandra and XMM-Newton X-ray cameras are such that many serendipitous sources are seen in low-latitude fields in addition to the primary targets and the preliminary results of such surveys have already emerged in the context of the Champlane and XMM-Newton SSC surveys \\citep{grindlay2003,motch2003,hands2004,grindlay2005,motch2006}. A number of dedicated Galactic surveys involving either single deep observations or a mosaic of shallower exposures have also been carried out or are in progress, \\citep[e..g. the Galactic Centre region,][]{muno2003,wijnands2006,koenig2008,revnivtsev2009,muno2009}. The nature of the intermediate- to low-luminosity Galactic hard X-ray sources discovered in the Galactic Plane surveys quoted above is only partially known. Cataclysmic variables (CVs) and quiescent X-ray binaries (both low- and high-mass types) account for a significant fraction, however, the exact nature of the majority of the sources detected in the central regions of the Galaxy remains doubtful in the absence of optical or infrared identifications, which in most cases are hindered by the large interstellar absorption. The observation of intermediate- to low-luminosity Galactic X-ray sources can address many issues related to binary-star evolution and to accretion physics at low rates. In addition surveys which extend the search boundaries have the potential to improve our knowledge of source population statistics and to unveil rare types of compact X-ray-emitting systems which have a low space density in the local Galactic neighbourhood. For instance, the determination of the space density and scale height of CVs is relevant to estimates of the Galactic novae rate and connects to the origin of low-mass X-ray binaries and type Ia supernovae. Low X-ray luminosity but long lived evolutionary stages of classical low and high mass X-ray binaries could also be found in relatively large numbers if the predictions of some evolutionary scenarios are correct \\citep[see e.g.][]{pfahl2002,willems2003}. Massive X-ray binaries are good proxies of relatively recent star-bursts and if detected in sufficient numbers might give insight into the recent star-formation history of specific regions of our Galaxy. Moreover, no X-ray telescope planned in the foreseeable future will have the necessary spatial resolution and sensitivity to study in detail the distribution of these low- and intermediate-luminosity sources in external galaxies. Consequently, X-ray surveys of our own Galaxy and its Magellanic satellites will remain for a long time the only means to study how the various X-ray emitting populations relate to the main Galactic structures such as the thin disc, the thick disc and the bulge. However, such studies presuppose the knowledge of the nature of the bulk of the X-ray sources surveyed, a prerequisite which can only be fulfilled through multi-wavelength spectroscopic identifications. Nevertheless, the statistically controlled cross-correlation with large catalogues such as the 2MASS catalogue and/or with the recently made available source lists from the GLIMPSE surveys, provide useful information on the intensity and shape of the part of the spectral energy distribution in the observing windows least affected by line-of-sight absorption in the interstellar medium. The work presented here was carried out in the framework of a Galactic Plane Survey programme instigated by the XMM-Newton Survey Science Centre (SSC). The long term objective of this project is to gather a representative sample of identified low-latitude X-ray sources, which eventually could be used as a template for identifying and classifying in a statistical manner the entire catalogue of serendipitous XMM-Newton sources detected in the Milky Way region. Companion programs are being carried out at high Galactic latitudes using three separate samples of sources selected at faint \\citep{furusawa2008}, medium \\citep{barcons2002} and bright \\citep{dellacecca2004} fluxes. In the first part of this paper, we recall the main properties of the \\xmm Galactic Plane Survey (XGPS) of \\cite{hands2004} and describe the way a sub-sample of X-ray sources were selected for optical follow-up at the telescope. The following section explains the method applied to cross-correlate the XGPS source list with the USNO-B1.0 and 2MASS catalogues and how identification probabilities were assigned to each optical or near infrared candidate. After a description of the observing procedures, the next section presents in detail the results of our optical spectroscopic identification campaign. The last sections discuss the \\lnls\\ curves and the statistics of the various kinds of X-ray emitters identified in the soft and hard bands as well as their contributions to the genuine Galactic population of low to medium X-ray luminosity sources. ", "conclusions": "We have successfully identified at the telescope a large number of the brightest X-ray sources detected in the XGPS survey. In addition, we were able to increase the number of identifications by cross-correlating X-ray positions with the USNO-B1.0 and 2MASS catalogues thanks to a carefully calibrated statistical method. The bulk of these statistical identifications with relatively bright optical or near infrared objects are likely stellar coronae based on their soft X-ray spectra. However, our optical campaign has demonstrated that some of them, in particular those exhibiting the hardest X-ray energy distributions, might be massive stars exhibiting particularly hard X-ray emission. Our optical spectroscopic follow-up observations have also confirmed the variety of the astrophysical objects contributing to the low to medium X-ray luminosity population of Galactic X-ray sources. We spectroscopically identify a total of 16 solar-type active stars, mostly among the softest X-ray sources with a high fraction of late Me stars. Furthermore, two early B type stars appear as the probable counterparts of some of our bright XGPS sources. We have discovered a total of three new cataclysmic variables with optical magnitudes in the range of $\\sim$ 23-22 and 0.5--10\\,keV fluxes from 0.7 to 4.7$\\times$10$^{-13}$\\ergscm . Four sources can qualify as massive X-ray binary candidates on the basis of hard X-ray emission and of the presence in the error circle of an early type star exhibiting \\Halp\\ emission. The optical counterpart of XGPS-3 is likely to be an hyper-luminous star possibly as bright as $\\eta$ Carinae. We also report the discovery of an X-ray selected Wolf-Rayet star. The brightest source of our sample, XGPS-1 is a known transient Low-Mass X-ray Binary exhibiting a bursting behaviour. The bright optical magnitude and unusual Bowen C{\\sc III}-N{\\sc III} emission of XGPS-25 might be the signature of a Low-Lass X-ray Binary in quiescence, although the observational picture could also be consistent with a rare type of CV. The main findings of our work can be summarized as follows: \\begin{enumerate} \\item In the 2--10\\,keV flux range from 10$^{-12}$ to 10$^{-13}$\\ergscm, we positively identify a large fraction of the hard X-ray sources with Galactic objects at a rate consistent with that expected for the Galactic contribution only. We note, however, that the diversity of the observed X-ray spectra makes the count to flux conversion somewhat uncertain. Moreover, different normalisations of the extragalactic \\lnls\\ curves are proposed in the literature. Taken together, these two issues do not allow a precise estimate to be made of the fraction of Galactic sources. However, based on hardness ratio considerations, we find that about half of the hard XGPS sources are likely Galactic objects, in rough agreement with the number estimated by subtracting the extragalactic ASCA \\lnls\\ curve from the relation obtained using the whole XGPS sample. At the faintest hard X-ray flux considered in our study of \\Fx\\ = 5$\\times$10$^{-14}$\\ergscm , we still identify spectroscopically or statistically between 30\\% and 50\\% of the 'Galactic fraction' of the hard X-ray sources. Since most of the identifications at the faint flux end rely on coincidence in position with a bright optical or infrared candidate, the nature of this optically bright population remains somewhat uncertain. However, drawing on the spectroscopic identifications of the optically bright X-ray brightest objects, we expect them for the most part to be made of particularly active late type stars and of early type stars. \\item The soft X-ray band is completely dominated by coronally emitting stars with CVs contributing only a small percentage of the identifications. We identify all sources down to \\Fx$\\sim$3$\\times$10$^{-14}$\\ergscm . With a mean integrated Galactic \\nh\\ of $\\sim$ 9$\\times$10$ ^{22}$ cm$^{-2}$ the background of extragalactic sources is completely screened out by line-of-sight absorption. Although the small number statistics does not allow us to draw robust conclusions, it seems that a larger number of faint Me stars is present in the XGPS survey than was identified in the ROSAT Galactic Plane Survey. Such an evolution is in rough agreement with the prediction of X-ray stellar population models. \\item A few relatively bright active coronae do contribute significantly above 2\\,keV. This probably indicates unusually high temperatures for the hottest of the two thin thermal components generally needed to represent the X-ray spectra of active stars \\citep[see a recent review in][]{guedel2009}. The observation of open clusters and field stars of different ages has established clear relations between stellar age, X-ray luminosity and temperature. For instance, the young stars in Orion (age $<$ 1\\,Myr) require on average $kT_1\\,\\sim$ 0.8\\,keV and $kT_2\\,\\sim$ 2.9\\,keV, while those of the $\\sim$ 160\\,Myr old stars in NGC 2516 have already dropped to $kT_1\\,\\sim$ 0.5\\,keV and $kT_2\\,\\sim$ 1.7\\,keV \\citep[see][ and references therein]{sung2008}. In contrast, the population of X-ray bright stars identified in the high galactic latitude XMM-Newton Bright Serendipitous Survey by \\cite{xbss2007} is well characterized by $kT_1\\,\\sim$ 0.3\\,keV and $kT_2\\,\\sim$ 1.0\\,keV, indicating a significantly older age on average. The hard X-ray coronae unveiled in the XGPS are thus likely to be very young stars in agreement with the predictions of X-ray stellar populations models. However, a fraction of these hard X-ray emitting stars may also be active binaries, mainly of the RS CVn type. Close binaries in which rotation and coronal activity is maintained by tidal synchronisation do amount to about one third of the ROSAT/Tycho-2 sample of stars \\citep{guillout2009}. \\item The hard X-ray source population has a strong component made of early-type stars exhibiting enhanced emission from a hard X-ray component. The physical mechanism responsible for the hard X-ray excess is unclear but could be related to wind collisions in the case of the Wolf-Rayet star counterpart of XGPS-14 and for the likely hyper-luminous star associated with XGPS-3, which shares several similarities with $\\eta$ Carinae. XGPS-36 is another example of an early Be star containing a well developed circumstellar disc and exhibiting thin thermal X-ray emission with typical temperatures of $\\sim$ 10\\,keV, much hotter than usually measured for normal OB stars in which shocks in the high-velocity radiation-pressure driven wind are responsible for the X-ray emission. The origin of the relatively modest X-ray emission (\\Lx\\ $\\sim$ 10$^{32-33}$\\ergs ) remains a matter of lively debate. Two distinct mechanisms have been proposed, mainly in the context of $\\gamma$-Cas, namely accretion on a companion star, most probably a white dwarf, or magnetic interaction between the early-type star and its decretion disc. XGPS-36 is the eighth identified member of the growing class of $\\gamma$-Cas analogs, and the fourth found so far in XMM-Newton observations of the Galactic Plane. \\end{enumerate} Our optical spectroscopic campaign has successfully unveiled the nature of the Galactic sources in the hard X-ray flux interval between 10$^{-12}$ and 10$^{-13}$\\ergscm. This \\Fx\\ range corresponds to the faintest flux limit covered by the ASCA Galactic Plane Survey but still remains between one and two orders of magnitudes above the sensitivity reached by the deepest Chandra pointings. Our identified sample of hard X-ray sources mainly consists of massive stars, possible X-ray binary candidates and CVs with X-ray luminosities in the range of \\Lx $\\sim$ 10$^{32}$ to 10$^{34}$\\ergs\\ and located within a few kpc from the Sun. At first glance, the nature of the ``local'' population of hard X-ray sources does not differ from that inferred from infrared follow-up imaging of deep Chandra pointings in the direction of the Galactic Centre \\citep[see e.g.][and references therein]{mauerhan2009}. However, the relative numbers of the various classes might well be quite different. Our work also illustrates the difficulty of optically identifying sources in this \\Fx\\ range, especially the cataclysmic variables as well as the importance of having arcsecond positions so as to alleviate the confusion issue." }, "0911/0911.2498_arXiv.txt": { "abstract": "In this work we present the first non-linear, non-Gaussian full Bayesian large scale structure analysis of the cosmic density field conducted so far. The density inference is based on the Sloan Digital Sky Survey data release 7, which covers the northern galactic cap. We employ a novel Bayesian sampling algorithm, which enables us to explore the extremely high dimensional non-Gaussian, non-linear log-normal Poissonian posterior of the three dimensional density field conditional on the data. These techniques are efficiently implemented in the HADES computer algorithm and permit the precise recovery of poorly sampled objects and non-linear density fields. The non-linear density inference is performed on a 750 Mpc cube with roughly 3 Mpc grid-resolution, while accounting for systematic effects, introduced by survey geometry and selection function of the SDSS, and the correct treatment of a Poissonian shot noise contribution. Our high resolution results represent remarkably well the cosmic web structure of the cosmic density field. Filaments, voids and clusters are clearly visible. Further, we also conduct a dynamical web classification, and estimated the web type posterior distribution conditional on the SDSS data. ", "introduction": "Observations of the large scale structure have always attracted enormous interest, since they contain a wealth of information on the origin and evolution of our Universe. The details of structure formation are very complicated and involve many different physical disciplines ranging from quantum field theory, general relativity or modified gravity to the dynamics of collisionless matter and the behavior of the baryonic sector. Throughout cosmic history the interplay of these different physical phenomena therefore has left its imprints in the matter distribution surrounding us. Probes of the large scale structure, such as large galaxy surveys, hence enable us to test current physical and cosmological theories and will generally further our understanding of the Universe. Especially a cosmographical description of the matter distribution will permit us to study details of structure formation mechanisms and the clustering behavior of galaxies as well as it will provide information on the initial fluctuations and large scale cosmic flows. For this reason, several different methods to recover the three dimensional density or velocity field from galaxy observations have been developed and applied to existing galaxy surveys \\citep[][]{1993PhRvE..47..704E,HOFFMAN1994,1994ASPC...67..171L,1994ApJ...423L..93L,1995A&AS..109...71Z,1995MNRAS.272..885F,1995ApJ...449..446Z,1997MNRAS.287..425W,1999ApJ...520..413Z,2001misk.conf..268V,2006MNRAS.373...45E,ERDOGDU2004,KITAURA2009B}. In particular, recently \\citet{KITAURA2009} presented a high resolution three dimensional Wiener reconstruction of the Sloan Digital Sky Survey data release 6 data, which demonstrated the feasibility of high precision density field inference from galaxy redshift surveys. These three dimensional density maps are interesting for a variety of different scientific applications, such as studying the dependence of galaxy properties on their cosmic environment, increasing the detectability of the integrated Sachs-Wolfe effect in the CMB or performing constrained simulations \\citep[see e.g.][]{BISTOLAS1998,LEE2008,LEELI2008,FROMMERT2008,KLYPIN2003,LIBESKIND2009,2009MNRAS.397.2070M}. However, modern precision cosmology demands an increasing control of observational systematic and statistical uncertainties, and the means to propagate them to any finally inferred quantity in order not to draw wrong conclusion on the theoretical model to be tested. For this reason, here we present the first application of the new Bayesian large scale structure inference computer algorithm HADES (HAmiltonian Density Estimation and Sampling) to data \\citep[see][for a description of the algorithm]{JASCHE2009B}. HADES performs a full scale non-linear, non-Gaussian Markov Chain Monte Carlo analysis by drawing samples from the lognormal Poissonian posterior of the three dimensional density field conditional on the data. This extremely high dimensional posterior distribution, consisting of \\(\\sim 10^6\\) or more free parameters, is explored via a numerically efficient Hamiltonian sampling scheme which suppresses the random walk behavior of conventional Metropolis Hastings algorithms by following persistent trajectories through the parameter space \\citep[][]{DUANE1987,NEAL1993,NEAL1996}. The advantages of this method are manyfold. Beside correcting observational systematics introduced by survey geometry and selection effects, the exact treatment of the non-Gaussian behavior and structure of the Poissonian shot noise contribution of discrete galaxy distributions, permits very accurate recovery of poorly sampled objects, such as voids and filaments. In addition, the lognormal prior has been demonstrated to be an adequate statistical description for the present density field and hence enables us to infer the cosmic density field deep into the non-linear regime \\citep[see e.g.][]{HUBBLE1934,PEEBLES1980,COLES1991,GAZTANAGA1993,KAYO2001}. The important thing to remark about HADES is, that it does not only yield a single estimate, such as a mean, mode or variance, in fact it provides a sampled representation of the full non-Gaussian density posterior. This posterior encodes the full non-linear and non-Gaussian observational uncertainties, which can easily be propagated to any finally inferred quantity. The application of HADES to Sloan Digital Sky Survey (SDSS) data therefore is the first non-linear, non-Gaussian full Bayesian large scale structure analysis conducted so far \\citep[SDSS;][]{YORK2000}. In particular, we applied our method to the recent SDSS data release 7 (DR7) data \\citep[DR7;][]{SDSS7}, and produced about 3TB of valuable scientific information in the form of \\(40000\\) high resolution non-linear density samples. The density inference is conducted on an equidistant cubic grid with side length \\(750\\) Mpc consisting of \\(256^3\\) volume elements. The recovered density field clearly reveals the cosmic web structure, consisting of voids, filaments and clusters, of the large scale structure surrounding us. These results provide the basis for forthcoming investigations on the clustering behavior of galaxies in relation to their large-scale environment. Such analyses yield valuable information about the formation and evolution of galaxies. In example, it has been known since long that physical properties such as morphological type, color, luminosity, spin parameter, star formation rate, concentration parameter, etc., are functions of the cosmic environment \\citep[see e.g.][]{DRESSLER1980,POSTMAN1984,WHITMORE1993,LEWIS2002,GOMEZ2003,GOTO2003,ROJAS2005,KUEHN2005,BLANTON2005,BERNARDI2006,CHOI2007,PARK2007,LEE2008,LEELI2008}. In this work we already conduct a preliminary examination of the dependence of stellar mass \\(M_{\\star}\\) and \\(g-r\\) color of galaxies on their large-scale environment. However, more thorough investigations will be presented in following works. Analyzing galaxy properties in the large-scale environment also requires to classify the large scale structure into different cosmic web types. We do so by following the dynamic cosmic web type classification procedure as proposed by \\citet{HAHN2007} with the extension of \\citet{FORERO2009}. The application of this procedure to our results yields the cosmic web type posterior, which provides the probability of finding a certain web type (void, sheet, filament, halo) at a given position in the volume conditional on the SDSS data. This permits simple propagation of all observational uncertainties to the final analysis of galaxy properties. The paper is structured as follows. We start by a brief review of the methodology in section \\ref{Methodology}, particularly describing the lognormal Poissonian posterior and the Bayesian computer algorithm HADES. Additionally, here we describe the dynamic web classification procedure as mentioned above. In section \\ref{DATA} we give a description of the SDSS DR7 data and present necessary data preparation steps required to apply the analysis procedure. Specifically, we describe the preparation of the linear observation response operator and the creation of the three dimensional data cube. In the following section \\ref{RESULTS} we present the results obtained from the non-linear, non-Gaussian sampling procedure. We start by analyzing the convergence behavior of the Markov chain via a Gelman \\& Rubin diagnostic, followed by a discussion of the properties of individual Hamiltonian samples. Here we also provide estimates for the ensemble mean density field and according variance. These fields depict remarkable well the properties of the cosmic web consisting of voids, filaments and halos. Based on these results we perform a simple cosmic web classification in section \\ref{web_classification_data}. In section \\ref{GALAXY_PROPERTIES}, we present a preliminary examination on the correlation between the large-scale environment of galaxies and their physical properties. In particular, here we study the stellar mass and \\(g-r\\) color of galaxies in relation with the density contrast \\(\\delta\\). We conclude the paper in section \\ref{SUMMARYANDCONCLUSION} by summarizing and discussing the results. ", "conclusions": "\\label{SUMMARYANDCONCLUSION} In this work we present the first application of the non-linear, non-Gaussian Bayesian large scale structure inference algorithm HADES to SDSS DR7 data. HADES is a numerically efficient implementation of a Hamiltonian Markov Chain sampler, which performs sampling in extremely high parameter spaces usually consisting of \\(\\sim 10^7\\) or more free parameters. In particular, HADES explores the lognormal Poissonian density posterior, which permits precision recovery of poorly sampled objects and density field inference deep into the non-linear regime \\citep[][]{JASCHE2009}. The large scale structure inference was conducted on a cubic equidistant grid with sidelength of \\(750\\) Mpc consisting of \\(256^3\\) voxels, yielding a grid resolution of about \\(3\\) Mpc. The large scale structure inference procedure correctly accounts for the survey geometry, completeness and radial selection effects as well as for the correct treatment of Poissonian noise. The analysis yielded about 3 TB of valuable scientific information in the form of full three dimensional density samples of the lognormal Poissonian density posterior. This set of density samples is thus a sampled representation of the full non-Gaussian density posterior distribution and therefore encodes all observational systematics and statistical uncertainties. Hence, all uncertainties and systematics can seamlessly be propagated to any finally inferred quantity, by simply applying the according inference procedure to the set of samples. In this fashion, the results permit us to make precise and quantitative statements about the large scale density field and any derived quantity. We stress that our Hamiltonian samples are not the result of a filtering procedure. A filter generally suppresses the power of the signal in low signal-to-noise regions and therefore does not yield a physical meaningful density, since it lacks power in poorly or unobserved regions. However, each Hamiltonian density sample represents a complete physical matter field realization conditional on the observations, in the sense that it possesses correct physical power throughout the entire volume. Already visual inspection of these density samples shows a homogeneous distribution of power throughout the entire volume. This fact was emphasized by the demonstration of power-spectra measured from these density samples, which show no sign of being affected by lack of power or artificial mode coupling nor do they show any sign of being affected by an adaptive smoothing kernel as would be expected for filter applications. It should be noted that this fact marks the crucial difference of our method to previous filter based density estimation procedures. In section \\ref{MEANANDVARIANCE}, we estimated the ensemble mean and the according variance from the \\(40000\\) density samples. The estimated ensemble mean nicely depicts the cosmic web consisting of filaments, voids and clusters extracted from the SDSS data. It is clear, that the ensemble mean represents the mean estimated from the lognormal Poissonian posterior conditional on the SDSS data. Therefore, it encodes the observational uncertainties and systematics. This can be seen by the fact, that the ensemble mean approaches cosmic mean density in poorly or not observed regions. Further, we plotted the according variance, which demonstrates that the non-Gaussian behavior and structure of the Poissonian shot noise was correctly taken into account in our analysis. Especially, the expected correlation between high mean density and high variance regions was clearly visible. We also estimated the cumulative probabilities for the density amplitude at each volume element, and demonstrated that the recovered density fields truly cover the broad range from linear to non-linear density amplitudes. To characterize the environment of our galaxy sample, but also to demonstrate the advantages of the Hamiltonian samples, we performed an example cosmic web type classification in section \\ref{web_classification_data}. In particular, we followed the dynamical cosmic web classification approach of \\citet{HAHN2007} with the extensions proposed by \\citet{FORERO2009}. This procedure involves the calculation of the cosmic deformation tensor and its eigenvalues. We demonstrated that this procedure can easily be applied to the set of samples, since they represent full physical matter field realizations. As a byproduct of this procedure we estimated the ensemble mean for the three eigenvalues of the cosmic deformation tensor. Further, we classified the individual volume elements as one of the four different web types void, sheet, filament and halo. The classification into four discrete web types enabled us to explicitely estimate the cosmic web posterior, which provides the probability of finding a specific web type at a given point in the volume conditional on the SDSS data. This result is especially appealing from a Bayesian point of view, since it emphasizes the fact, that the result of a Bayesian method is a complete probability distribution rather than just a single estimate. Here we saw, that especially voids are a sensitive measure for the cosmic web. Of course, it is possible to repeat the cosmic web classification in a similar manner with any other classification procedure. In the following section \\ref{GALAXY_PROPERTIES}, we presented a preliminary examination of the correlation between the large-scale environment and physical properties of galaxies. In particular, we considered the stellar mass and $g-r$ color of galaxies in relation to the density contrast \\(\\delta\\). A qualitative analysis revealed that there exist correlation between these galaxy properties and the large scale structure. In particular, massive galaxies are more likely to be found in massive structures, while low mass galaxies reside in void like structures. The plots also demonstrate the different clustering behavior of red and blue galaxies. Also note, that these observed trends are consistent with previous works \\citep[][]{LEE2008,LEELI2008,LI2006}. However, more work is required in order to provide quantitative statements. This will be done in a forthcoming publication. The results presented in this work will be valuable for many subsequent scientific analyses of the dependence of galaxy properties on their cosmic environment. Herefore, particularly the Hamiltonian samples allow for a more intuitive handling of observational data, since they can be understood as full matter field realizations or different multiverses consistent with our data of the Universe we live in. Beside providing quantitative characterizations of the large scale structure, the results also give us an intuitive understanding of the three dimensional matter distribution in our cosmic neighborhood. We intend to make our data publically available to the community. Future applications will also take into account non-linear bias models and peculiar velocity sampling procedures, to provide even more accurate density analyses. We hope that this work demonstrates the potential of Bayesian large scale structure inference and its contribution to current and future precision analyses of our Universe." }, "0911/0911.2983_arXiv.txt": { "abstract": "We study the \\Mbh{}--\\Mhost{} relation as a function of Cosmic Time in a sample of 96 quasars from $z=3$ to the present epoch. In this paper we describe the sample, the data sources and the new spectroscopic observations. We then illustrate how we derive \\Mbh{} from single-epoch spectra, pointing out the uncertainties in the procedure. In a companion paper, we address the dependence of the ratio between the black hole mass and the host galaxy luminosity and mass on Cosmic Time. ", "introduction": "The discovery of tight relations between the mass of massive black holes, \\Mbh{}, and the large scale properties of the galaxies where they reside \\citep[see][for a review]{ferrarese06} is one of the most intriguing results in astrophysics of the last decade, given the consequences in the frame of galaxy formation and evolution. When and how these relations are set are still open questions. Measuring \\Mbh{} of quiescent massive black holes is extremely challenging even in nearby galaxies, and has been done successfully only in few tens of cases \\citep{ferrarese06,pastorini07}. By contrast, an estimate of \\Mbh{} in Type-I Active Galactic Nuclei (AGN) is possible by assuming that the gas emitting broad lines is in virial equilibrium \\citep[][but see also Marconi et al. 2008 and references therein for possible effects of non-virial components]{peterson00}, and that the line width traces the black hole potential well. Based on reverberation mapping studies \\citep{blandford82}, \\citet{kaspi00} found a correlation between the characteristic size of the broad line region (hereafter, BLR) and the continuum luminosity of the AGN. This allows an estimate of the black hole mass from single epoch low-resolution spectra. In order to sample the black hole -- host galaxy relations through a wide range of Cosmic Ages, one has to focus on the brightest AGN. Quasars have been detected up to $z\\gsim6$ \\citep[][]{fan04}. Large field surveys such as the SDSS allowed a detailed spectroscopic study of $\\sim60,000$ quasars with $z<4$ \\citep{shen08}. The drawback is that the typical nuclear-to-host luminosity ratio in quasars is such that the light from the host galaxy is outshone by the nuclear component. This usually prevents the detection of stellar features in the spectra of bright quasars. Only through the excellent resolution of the HST, together with state-of-art observing techniques in the NIR at ground-based telescopes, the detection of the extended emission of the host galaxies of few hundreds of quasars up to $z\\lsim3$ became possible \\citep[see][and references therein]{kotilainen09}. In this project we focus on quasars in the redshift range $00.5$) quasars imaged in the NIR through ground-based observations under optimal seeing conditions performed by our group \\citep[50 objects; see][]{kotilainen98,kotilainen00, falomo04,falomo05,hyvonen07a,hyvonen07b,kotilainen07,falomo08,kotilainen09, decarli09a} or from HST-based compilations \\citep[10 sources from][]{kukula01,ridgway01}. We note that all these studies have very high host galaxy detection rates ($>$$85$ per cent at $z<2$, $>$$60$ per cent beyond $z=2$). Optical spectra were collected at the Nordic Optical Telescope and the ESO/3.6m telescope (see section \\ref{sec_observations}). \\end{itemize} The 5 unresolved quasars in \\citet{kotilainen09} were spectroscopically observed when the analysis of the imaging data was not complete yet. They are not included in the study of the evolution of the \\Mbh{}--\\Mhost{} relation. Other two objects were dropped, as we could not estimate \\Mbh{} from our spectra (see Appendix \\ref{sec_appendix}). Therefore the final sample consists of 96 quasars. According to the \\citet{vcv06} catalogue, 48 of them are radio loud quasars (RLQs) and 48 are radio quiet (RQQs). We remark that our sample is approximately twice as large as those of \\citet{peng06a,peng06b} and \\citet{mclure06} and represents the largest dataset ever considered in the study of the evolution of the \\Mbh{}--\\Mhost{} relation. The distribution of our targets in the ($z$,$M_V$) plane is shown in Figure \\ref{fig_distr_z}. Table \\ref{tab_sample} lists the main properties of each quasar in our sample. \\begin{figure} \\begin{center} \\includegraphics[width=0.49\\textwidth]{fig_zMabs.ps}\\\\ \\caption{The distribution of the quasars in our sample in the ($z$,$M_V$) plane. $M_V$ is the total rest-frame $V$-band absolute magnitude of the quasars, estimated from the $V$-band apparent magnitudes available in the \\citet{vcv06} catalogue and $k$-corrected assuming the quasar template by \\citet{francis91}. Filled symbols refer to the mid- and high-$z$ data for which we present optical spectroscopy in this paper. Empty symbols mark the $z\\lsim0.5$ data from \\citet{labita06} and \\citet{decarli08a}. Circles are radio loud quasars, triangles are radio quiet. The broad emission lines observed in the various redshift windows are also labelled. We note that the $-27M_V>M_*(z)-1$ \\citep[where $M_*(z)$ is the characteristic luminosity of the quasar luminosity function by][plotted in short-dashed lines]{boyle00} are well sampled at any redshift bin. }\\label{fig_distr_z} \\end{center} \\end{figure} \\section[]{New observations, data reduction}\\label{sec_observations} Spectra of the $z>0.5$ quasars were collected in 5 observing runs at the European Southern Observatory (ESO) $3.6$m telescope in La Silla (Chile) and at the $2.56$m Nordic Optical Telescope (NOT) in La Palma (Spain). Table \\ref{tab_runs} lists the dates of the observations and the number of spectra collected in each run, while table \\ref{tab_journal} summarizes the journal of observations. The ESO Faint Object Spectrograph and Camera (v.2, EFOSC2; see Buzzoni et al. 1984) and its twin NOT instrument, the Andalucia Faint Object Spectrograph and Camera (ALFOSC), were mounted in long-slit spectroscopy configuration. EFOSC2 observations were carried out with grism \\#4, yielding a spectral resolution of $R\\sim 400$ ($1.2\"$ slit) in the spectral range $4100$--$7500$ \\AA{} ($\\Delta\\lambda/$pxl=$3.36$ \\AA/pxl). For NOT observations, we used ALFOSC grisms \\#6 and \\#7, which allow the observation of the $3500$--$5530$\\footnote{The nominal observed range of ALFOSC grism \\#6 is larger blue-wards, but the sensitivity is so low that we decided to drop the observed spectra at wavelengths below $3500$ \\AA.} and $3800$--$6840$ \\AA{} windows with spectral resolutions $R\\sim490$ and $650$ with the $1.0\"$ slit ($\\Delta\\lambda/$pxl$\\approx1.5$ \\AA/pxl). At the central wavelength, the instrumental resolutions are $12.6$ \\AA{} (EFOSC2+grism \\#4) and $8.2$ \\AA{} (ALFOSC+grism \\#6 and \\#7). \\begin{table} \\begin{center} \\caption{List of the observing runs. (1) Run ID. (2) Telescope (3) Dates of observations. (4) Number of observed objects per run. } \\label{tab_runs} \\begin{tabular}{cccc} \\hline Run & Telescope & Nights & N.obj.\\\\ (1) & (2) & (3) & (4) \\\\ \\hline E77 & ESO/$3.6$m & Sep.30-Oct.1 2005 & 6 \\\\ E78 & ESO/$3.6$m & Mar.23-25 2007 & 12 \\\\ E79 & ESO/$3.6$m & Sep.8-12 2007 & 22 \\\\ N35 & NOT & Apr.9-10 2007 & 2 \\\\ N36 & NOT & Oct.17-19 2007 & 18 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\begin{figure*} \\begin{center} \\includegraphics[angle=-90, width=0.9\\textwidth]{fig_example.ps}\\\\ \\caption{The combined spectrum of the quasars considered in the present study, normalized to the continuum value at 2250 \\AA{}. The median spectrum is plotted with a solid line. Main emission lines are labelled. The shaded regions mark the intervals used in the fit of the continuum. The resulting fit is plotted as a dash line. }\\label{fig_cont} \\end{center} \\end{figure*} Standard IRAF tools were used in the data reduction. The \\texttt{ccdred} package was employed to perform bias subtraction, flat field correction, image alignment and combination. Cosmic rays were eliminated by combining 3 or more exposures of the same objects, and applying \\texttt{crreject} algorithm while averaging. When only one or two bidimensional spectra were available, we applied the \\texttt{cosmicrays} task in the \\texttt{crutils} package. In these cases, in order to prevent the task from altering the narrow component of emission lines, we masked the central region of our bidimensional spectra. The spectra extraction, the background subtraction and the calibrations both in wavelength and in flux were performed with \\texttt{doslit} task in \\texttt{specred} package, using He-Ar, Th-Ar and He-Ne lamps and spectrophotometric standard stars as reference. Wavelength calibration residuals are around $0.5$, $0.35$ and $0.03$ \\AA{} in the three adopted setups (sub-pixel), thus implying a negligible ($<1$ per cent) error on redshift estimates. Absolute flux calibration of spectra was corrected through the photometry of field stars, as described in \\citet{decarli08a}. This procedure yields uncertainties in the flux calibration as low as $0.05$ mag \\citep[see, e.g., ][]{kotilainen09}, and commonly around $0.1$ mag. Photometric accuracy of each target is reported in table \\ref{tab_journal}. Galactic extinction was accounted for according to \\citet{schlegel98}, assuming $R_V = 3.1$. We shifted the spectra to the rest frame, according to the catalogue $z$, remembering that quasar lines with different ionization potentials may present slightly different shifts \\citep[e.g.][]{bonning07}. Average signal to noise ratio of our spectra is $\\sim$30. The composite spectrum, obtained by median averaging rest frame individual observations of these new data, is presented in Figure \\ref{fig_cont}. The whole dataset is available electronically at \\texttt{www.dfm.uninsubria.it/astro/caqos/}. \\section[]{Data analysis}\\label{sec_analysis} We focus our attention on the analysis required to estimate \\Mbh{} from single epoch observations of the rest-frame UV spectra of quasars. Applied to the gas in the BLR of Type-1 AGN, the virial paradigm yields: \\begin{equation}\\label{eq_virial} {\\cal M}_{\\rm BH} = G^{-1} R_{\\rm BLR} v_{\\rm BLR}^2 \\end{equation} where $R_{\\rm BLR}$ is the characteristic radius of the broad line emission, and $v_{\\rm BLR}$ is the velocity of the emitting clouds at $R_{\\rm BLR}$. The cloud velocity can be inferred from the width of the broad emission lines, e.g.: \\begin{equation}\\label{eq_def_f} v_{\\rm BLR} = f \\cdot {\\rm FWHM} \\end{equation} where $f$ is a geometrical factor around unity which accounts for the de-projection of $v_{\\rm BLR}$ from the line-of-sight, and FWHM is the Full Width at Half Maximum of the line profile (see McGill et al. 2008 and Decarli et al. 2008 for discussions on other line width parametrizations). On the other hand, the BLR size cannot be directly measured from single epoch spectra. Our estimates of the broad line region size rely on the discovery that $R_{\\rm BLR}$ scales with a certain power of the continuum luminosity of the AGN, $\\lambda L_\\lambda$ \\citep[see][]{kaspi00}, as expected from simple photoionization models. \\subsection[]{The continuum luminosity and the BLR size} Quasar UV--optical spectra are characterized by the superposition of the following components: \\begin{itemize} \\item[-] A power-law-like continuum from the nucleus; \\item[-] Broad lines emitted within the BH influence radius; \\item[-] Narrow emission lines from the quasar host galaxy and the AGN Narrow Line Region; \\item[-] The star light continuum from the host galaxy; \\item[-] A pseudo-continuum due to the blending of several \\FeII{} and \\FeIII{} multiplets. \\end{itemize} In order to infer the BLR radius, we have to isolate the first component from the others. Our spectra cover the rest-frame UV range of bright AGN, where the flux from the host galaxy star light is always negligible. The contamination of both broad and narrow emission lines is usually avoided simply by fitting the power-law continuum to the observed spectra in a number of wavelength windows free of strong features. Here we adopted the intervals: 1351--1362, 1452--1480, 1680--1710, 1796--1834, 1970--2010, 2188--2243, 2950--2990, 3046--3090 \\AA{} (see Figure \\ref{fig_cont}). The fitted parameters, together with the derived monochromatic specific fluxes and luminosities are reported in table \\ref{tab_fits}. We note that the luminosity estimates obtained through the fit of the power-law component and those obtained from a direct measure of the continuum at 1350 and 3000 \\AA{} are practically equivalent in our datasets, the differences being $<10$ per cent on average. Concerning luminosity--radius relations, simple photoionization models with a constant ionization parameter predict that a ionizing source emitting isotropically affects a region with a characteristic radius scaling as the square root of the source luminosity. This dependence has been confirmed in several reverberation mapping studies focused on \\Hb{}. Time-lag data of the rest-frame UV lines are still limited in number, therefore any available relation for \\Mgii{} and \\Civ{} is affected by poor statistics \\citep[e.g., see figure 6 in][]{kaspi07}. For consistency with our low-redshift studies, we will refer to the the relations provided by \\citet{mclure02} for \\Mgii{}\\footnote{\\citet{mclure02} set their relation by comparing the \\Hb{} time lags to the continuum luminosity around 3000 \\AA{}, given the fact that \\Mgii{} and \\Hb{} have similar ionizing potentials, hence they are emitted in the same region. This assumption allows us to adopt the same geometrical factor for the \\Mgii{} BLR as the one derived in \\citet{decarli08a} for \\Hb{}. Note that \\citet{mclure02} fitted the line profile of \\Mgii{} into a broad and a narrow component, and used only the former to infer \\Mbh{} \\citep[although there is no evidence of significant narrow components in \\Mgii{} lines: see, e.g.,][]{shen08}. As a consequence, the values of \\Mbh{} reported in their study are systematically larger than ours, but our conclusions are unchanged as we adopt our own internal calibration of $f$.}: \\begin{equation} \\frac{R_{\\rm BLR}({\\rm MgII})}{10~ {\\rm lt-days}}=(2.52\\pm0.3) \\left[\\frac{\\lambda L_{\\lambda}(3000 \\AA)}{10^{44} {\\rm erg/s}}\\right]^{0.47\\pm0.05} \\end{equation} and by \\citet{kaspi07} for \\Civ{}: \\begin{equation} \\frac{R_{\\rm BLR}({\\rm CIV})}{10~ {\\rm lt-days}}=(0.24\\pm0.06) \\left[\\frac{\\lambda L_{\\lambda}(1350 \\AA)}{10^{43} {\\rm erg/s}}\\right]^{0.55\\pm0.04} \\end{equation} All these relations are based on low-redshift ($z<0.3$) objects, with the only exception of S5 0836+71 \\citep[$z=2.17$; see][]{kaspi07}. Throughout this work, we assume that no significant Cosmic evolution of the radius--luminosity relations occurs in the sampled redshift range. The \\Civ{} relation is poorly constrained, as mentioned above, especially in the bright side where most of our objects reside. Assuming a different slope of the luminosity--time lag relation, e.g., $0.5$, affects our radius estimates at \\lLl $\\sim$ $10^{47}$ erg/s up to $0.2$ dex (i.e., a factor $\\approx1.7$). Nevertheless, we stress here that in terms of \\Mbh{} estimates such a difference would result in a global offset of all values and in a re-definition of the geometrical factor $f$, with little effect on the evolution of the \\Mbh{}--\\Mhost{} relation. The scatter around the luminosity--radius relations dominates the uncertainties in the radius estimates, contributing up to a factor $\\sim2$, while the uncertainties in the luminosity estimates never exceed 10 per cent. \\begin{figure} \\begin{center} \\includegraphics[width=0.49\\textwidth]{fig_fit2.ps}\\\\ \\caption{Examples of the line fit, for \\Civ{} in quasar Q2225-403 A (\\emph{upper panel}) and for \\Mgii{} in quasar Q2225-403 B (\\emph{lower panel}). The rest-frame observed spectra of the two quasars are shown in dotted lines. Dotted vertical lines mark the regions used for the underlying continuum estimates and the fitted wavelength range. The thin, solid lines show the continua underlying broad emission lines. The bold, solid lines are the broad emission line fits. In the bottom side of each panel, we plot the fitted \\FeII{} template and the broad line model (dotted lines) together with the fit residual (dashed line). }\\label{fig_fit} \\end{center} \\end{figure} \\subsection[]{Emission line widths and cloud velocities} Measuring the width of broad lines consists of three key steps: the definition of the local continuum, the removal of contaminating spectral features (in particular the \\FeII{} multiplets), and the fit with a certain analytical function. We define underlying continua by matching the fluxes in the windows 2230--2250 and 3020--3050 \\AA{} for \\Mgii{} and 1465--1485 and 1685--1705 \\AA{} for \\Civ{}, where no significant feature is observed (see Figure \\ref{fig_fit}). To evaluate the effects of \\FeII{} contamination, we first fitted the \\Mgii{} and \\Civ{} lines in the 2720--2880 and 1490--1570 \\AA{} windows, where the \\FeII{} contribution is less relevant. Then, we compared these estimates of the line widths with those obtained by fitting the 2450--3020 and 1490--1620 \\AA{} regions with a superposition of a broad line model plus a template reproducing the \\FeII{} emission \\citep{vestergaard01}. In the latter approach the \\FeII{} emission observed in narrow line Type-1 AGN (usually, IZw001) is taken as representative for all quasars, after being properly broadened. The validity of this assumption has been widely discussed in the literature, both from a theoretical and an observational point of view \\citep[e.g.,][]{phillips78,boroson92,marziani03,tsuzuki06,mcgill08}. Since most \\FeII{} lines are blended with each others, their width is hardly constrained. We therefore fixed their broadening to the width of \\Civ{} and \\Mgii{} lines. Figure \\ref{fig_fit} shows examples of fits on \\Civ{} and \\Mgii{}. Line width estimates obtained with or without the \\FeII{} modelling are usually consistent within 20 per cent, but \\Mgii{} shows few deviations larger than 30 per cent, especially when the \\FeII{} emission is strong with respect to the line flux. Hereafter, we will refer to the line models obtained by template fitting the \\FeII{} emission. \\begin{figure*} \\begin{center} \\includegraphics[width=0.4\\textwidth]{fig_lf_hb.ps} \\includegraphics[width=0.4\\textwidth]{fig_lf_c4_lz.ps}\\\\ \\includegraphics[width=0.4\\textwidth]{fig_lf_m2.ps} \\includegraphics[width=0.4\\textwidth]{fig_lf_c4_hz.ps}\\\\ \\caption{The distribution of our data in the (\\lLl,FWHM) plane, for \\Hb{}, \\Civ{} at low-$z$ (upper panels), \\Mgii{} and \\Civ{} at high-$z$ (bottom panels). The (normalized) projections along each axis are also shown. The loci with constant \\Mbh{} (dot-dashed lines) and \\Edd{} (dashed lines) are plotted. Concerning the luminosities, low-$z$ \\Civ{} data show on average The values of \\Hb{} FWHM show a wider distribution with respect to \\Mgii{}. data are log-normally distributed both in the \\lLl{} and in the FWHM values. \\Civ{} data show less spread in the sampled \\lLl{} space, and a flatter distribution in the line width with respect to the \\Mgii{}. }\\label{fig_labita} \\end{center} \\end{figure*} Concerning the fitted function, a single gaussian usually does not provide satisfactory fits, especially for \\Civ{} \\citep[see][]{decarli08a,shen08,bon09}. We performed our analysis both with the sum of 2 gaussian profiles with the same peak (hereafter, G2) and the Gauss-Hermite series \\citep[GH;][]{vandermarel93}, truncated at the fourth order. Reduced-$\\chi^2$ values obtained from these functions are similarly good, suggesting that both functions can describe line profiles reasonably well. No significant offset is observed among the line width estimates based on the two functions, the residuals lying within 20 per cent \\citep[see also Figure A1 in][]{decarli08a}. We thus conclude that the two functions are equivalent to our purposes. We note that, since G2 fits are necessarily symmetric, while GH fits are not, the consistency between the two estimates of the line width suggests that the FWHM is poorly sensitive to line profile asymmetries. Throughout the paper, we will refer to the FWHM estimates based on GH fits. Summarizing, we estimate that the typical uncertainties in the line width estimates due to the adopted fit procedures lie within 20 per cent, and usually even lower (if some modelling of the \\FeII{} emission is adopted). Reduced-$\\chi^2$ maps suggest that, given the high signal to noise ratio of our data (exceeding 20 in all but few cases), formal uncertainties in the parameter estimates contribute to $\\lsim 10$ per cent of the FWHM value. Therefore, we conclude that typical uncertainties in the FWHM estimates are around 20 per cent. \\section[]{Virial estimates of BH masses}\\label{sec_results} In this section, we match the new estimates of the continuum luminosity, FWHM and \\Mbh{} with those from our previous low-$z$ studies \\citep{labita06,decarli08a}. In Figure \\ref{fig_labita}, we plot the distribution of our data in the (\\lLl,FWHM) plane \\citep[see a similar approach in][]{fine08,labita09a}. This allows us to monitor how observable quantities (here, the continuum luminosity of the quasar and the line widths) affect our estimates of \\Mbh{} and the Eddington ratio, \\Edd, derived \\emph{assuming} equation \\ref{eq_virial}, $f({\\rm H\\beta,MgII})=1.6$, $f({\\rm CIV})=2.4$ \\citep[as defined in equation \\ref{eq_def_f} and in][]{decarli08a} and the bolometric corrections given in \\citet{richards06}: $L_{\\rm bol}/$\\lLl{}=$9.26$, $5.15$ and $3.81$ for \\Hb{}, \\Mgii{} and \\Civ{} respectively. Average and rms values of FWHM, \\lLl{}, \\Mbh{} and \\Edd{} are provided in Table \\ref{tab_ave_results}, while data of individual quasars are given in table \\ref{tab_fits}. The monochromatic luminosity at 5100 \\AA{} is on average $3.5$ times fainter than at 1350 \\AA{} in the same redshift bin. This is consistent with an average power-law index of the quasar continuum $\\alpha$ close to 2 (defined so that $F_\\lambda\\propto\\lambda^{-\\alpha}$). The 3000 \\AA{} luminosities sampled in our study range from few times $10^{44}$ to $10^{47}$ erg/s, the bulk being around $10^{46}$ erg/s. The \\Civ{} data at high-$z$ are more clustered around approximately the same luminosity as \\Mgii{} data, $\\approx 6\\cdot 10^{45}$ erg/s. This thinner distribution is due to the combination of two selection effects: lowest luminosity quasars are missed due to the Malmquist bias (the \\Civ{} line falls in the optical bands at $z>1.6$, see Figure \\ref{fig_distr_z}); highest luminosity quasars were rejected in the studies of the host galaxy luminosities, in order to make the detection of the extend emission around the nuclear source more feasible \\citep[see Figure 1 and ][]{kotilainen09}. Concerning the line widths, \\Hb{} values show a wider dispersion with respect to the \\Mgii{} line, notwithstanding the two lines have similar ionization potential, thus they are believed to originate in the same regions \\citep[see also][]{shen08,labita09a,labita09b}. Low-redshift data from the \\Civ{} line show a smaller average value with respect to both \\Hb{} and high-$z$ \\Civ{} data, and a significantly narrower distribution \\citep[see the discussion in][]{decarli08a}. With only few exceptions, all our data reside in the locus defined by $10^{8} \\gsim$ \\Mbh{}/\\Msun{} $\\gsim 10^{10}$ and $0.01 \\gsim$ \\Edd{} $\\gsim 1$. \\begin{table} \\begin{center} \\caption{Average and rms values of FWHM (2), \\lLl{} (3), \\Mbh{} (4) and \\Edd{} (5) for all the subsamples in this study. } \\label{tab_ave_results} \\begin{tabular}{ccccc} \\hline Line &$\\langle$log FWHM$\\rangle$&$\\langle$log \\lLl$\\rangle$&$\\langle$log \\Mbh{}$\\rangle$&$\\langle$log \\Edd{}$\\rangle$ \\\\ & [km/s] & [erg/s] & [\\Msun] & \\\\ (1) & (2) & (3) & (4) & (5) \\\\ \\hline \\multicolumn{5}{l}{{\\it Low-$z$}} \\\\ \\Hb{} & $3.66 \\pm 0.19$ & $45.01 \\pm 0.51$ & $9.01 \\pm 0.47$ & $-1.13 \\pm 0.45$ \\\\ \\Civ{} & $3.60 \\pm 0.11$ & $45.55 \\pm 0.56$ & $9.06 \\pm 0.38$ & $-1.03 \\pm 0.34$ \\\\ \\hline \\multicolumn{5}{l}{{\\it High-$z$}} \\\\ \\Mgii{} & $3.63 \\pm 0.19$ & $45.78 \\pm 0.54$ & $9.21 \\pm 0.49$ & $-0.82 \\pm 0.46$ \\\\ \\Civ{} & $3.66 \\pm 0.19$ & $45.74 \\pm 0.36$ & $9.28 \\pm 0.47$ & $-1.06 \\pm 0.41$ \\\\ \\hline \\end{tabular} \\end{center} \\end{table} In Figure \\ref{fig_shen} we compare the distributions of \\Mbh{} and \\Edd{} from our dataset with the huge quasar sample from \\citet[][S08]{shen08}\\footnote{From Shen et al. (2008) we take the values of $z$, FWHM and \\lLl{} of the objects in the same redshift ranges than our study. All the derived quantities (e.g., \\Mbh{} and \\Edd{}) are re-computed following the recipes discussed in this work.}. We dropped our low-$z$, \\Civ{}-based data from this comparison as they do not have any counter-part in the S08 data. The distributions of black hole masses and Eddington ratios computed from \\Hb{} in our study are in substantial agreement with those by S08. On the other hand, the black hole masses in our sample have increasingly smaller values than those from S08. We interpret this trend as the superposition of a number of effects: the luminosity cut adopted by \\citet{kotilainen09} for $z>2$ objects, the evolution of the luminosity and mass functions of active black holes through redshift and the occurrence of the Malmquist bias \\citep[on this topic, see][]{labita09a,labita09b}. On average, the Eddington ratios sampled in our study do not show significant differences with respect to those in S08, nor evolution from $z=0$ to $z=3$. \\begin{figure} \\begin{center} \\includegraphics[width=0.49\\textwidth]{fig_shen2b.ps}\\\\ \\caption{ The distributions of \\Mbh{} and \\Edd{} as derived from \\Civ{} (\\emph{upper panels}), \\Mgii{} (\\emph{central panels}) and \\Hb{} (\\emph{bottom panels}). Low-$z$ \\Civ{}-based estimates are not included, as no counter-part is available in the \\citet{shen08} sample. Shaded histograms refer to our objects, while empty histograms refer to the dataset of \\citet{shen08}. }\\label{fig_shen} \\end{center} \\end{figure} The relatively small statistics and the luminosity-based selection of the targets in our sample hinder the study of the dependence of \\Mbh{} on the redshift. We note that, on average, the higher is the redshift, the more massive are the black holes (see Figure \\ref{fig_mbh_z}). The linear best fit is: \\begin{equation} \\log {\\cal M}_{\\rm BH}/{\\rm M}_{\\odot} = (0.19 \\pm 0.06) z + (8.98 \\pm 0.06) \\end{equation} This trend is in qualitative agreement with the one found by \\citet{labita09b} with a procedure aimed to minimize the Malmquist bias. However, the occurrence of selection biases cannot be ruled out in the present work. \\begin{figure} \\begin{center} \\includegraphics[angle=-90, width=0.49\\textwidth]{fig_mbh_z.ps}\\\\ \\caption{The redshift dependence of \\Mbh{} for the quasars in our sample. Circles, triangles and squares refer to \\Civ{}, \\Mgii{} and \\Hb{}-based \\Mbh{} estimates. The best linear fit is shown as a solid line. The dashed line refers to the redshift evolution of the maximum \\Mbh{} as derived by \\citet{labita09b}. }\\label{fig_mbh_z} \\end{center} \\end{figure} ", "conclusions": "We start from a list of 96 quasars with known host galaxy luminosity. New, high-quality (S/N$\\sim$30) spectra of the mid- and high-redshift targets ($z>0.5$) are presented here and matched with those of low-$z$ quasars we collected in our previous studies. We analyse the continuum luminosity and profiles of broad emission lines in order to infer black hole masses. In particular, the \\Civ{} line is studied both at low and high redshift, thus avoiding a systematic error related to the adopted emission line in the estimate of \\Mbh{}. We found that high redshift quasars in our sample do have, on average, larger BH masses than local ones, but we note that this result is potentially affected by the luminosity selection and the Malmquist bias. In a accompanying Paper II, we study the ratio between \\Mbh{} and the luminosity and mass of the host galaxies of the quasars in our sample, and its evolution as a function of the redshift." }, "0911/0911.1949_arXiv.txt": { "abstract": "{} {In this paper, we explore the diagnostic power of the far-IR fine-structure lines of [O{\\sc i}] 63.2$\\,\\mu$m, 145.5$\\,\\mu$m, [C{\\sc ii}] 157.7$\\,\\mu$m, as well as the radio and sub-mm lines of CO J=1-0, 2-1 and 3-2 in application to disks around Herbig Ae stars. We aim at understanding where the lines originate from, how the line formation process is affected by density, temperature and chemical abundance in the disk, and to what extent non-LTE effects are important. The ultimate aim is to provide a robust way to determine the gas mass of protoplanetary disks from line observations.} {We use the recently developed disk code {{\\sc ProDiMo}} to calculate the physico-chemical structure of protoplanetary disks and apply the Monte-Carlo line radiative transfer code {{\\sc Ratran}} to predict observable line profiles and fluxes. We consider a series of Herbig Ae type disk models ranging from $10^{-6}$~M$_\\odot$ to $2.2\\!\\times\\!10^{-2}$~M$_\\odot$ (between 0.5 and 700\\,AU) to discuss the dependency of the line fluxes and ratios on disk mass for otherwise fixed disk parameters. This paper prepares for a more thorough multi-parameter analysis related to the Herschel open time key program {{\\sc Gasps}}.} {We find the [C{\\sc ii}]\\,157.7\\,$\\mu$m line to originate in LTE from the surface layers of the disk, where $\\Tg\\neq\\Td$. The total emission is dominated by surface area and hence depends strongly on disk outer radius. The [O{\\sc i}] lines can be very bright ($>10^{-16}$~W/m$^2$) and form in slightly deeper and closer regions under non-LTE conditions. For low-mass models, the [O{\\sc i}] lines come preferentially from the central regions of the disk, and the peak separation widens. The high-excitation [O{\\sc i}]\\,145.5\\,$\\mu$m line, which has a larger critical density, decreases more rapidly with disk mass than the 63.2\\,$\\mu$m line. Therefore, the [O{\\sc i}] 63.2\\,$\\mu$m/145.5\\,$\\mu$m ratio is a promising disk mass indicator, especially as it is independent of disk outer radius for $R_{\\rm out}>200$~AU. CO is abundant only in deeper layers $A_V\\!\\ga\\!0.05$. For too low disk masses ($M_{\\rm disk}\\!\\la\\!10^{-4}$~M$_\\odot$) the dust starts to become transparent, and CO is almost completely photo-dissociated. For masses larger than that the lines are an excellent independent tracer of disk outer radius and can break the outer radius degeneracy in the [O{\\sc i}]\\,63.2\\,$\\mu$m/[C\\,{\\sc ii}]157.7\\,$\\mu$m line ratio.} {The far-IR fine-structure lines of [C{\\sc ii}] and [O{\\sc i}] observable with Herschel provide a promising tool to measure the disk gas mass, although they are mainly generated in the atomic surface layers. In spatially unresolved observations, none of these lines carry much information about the inner, possibly hot regions $<\\!30\\,$AU.} ", "introduction": "Observations of gas in protoplanetary disks are intrinsically difficult to interpret as they reflect the interplay between a complex chemical and thermal disk structure, statistical equilibrium and optical depth effects. This is particularly true if non-thermal excitation such as fluorescence or photodissociation dominate the statistical equilibrium. The first studies of gas in protoplanetary disks concentrated on the rotational transitions of abundant molecules such as CO, HCN and HCO$^+$ \\citep[e.g.][]{Beckwith1986, Koerner1993, Dutrey1997, vanZadelhoff2001, Thi2004}. Those lines originate in the outer regions of disks, $r>100$~AU, where densities are at most $n\\sim10^7$~cm$^{-3}$. The interpretation of those lines was mainly based on tools and expertise developed for molecular clouds. Using the CO J=3-2 line, \\citet{Dent2005} inferred for a sample of Herbig Ae and Vega-type stars a trend of disk outer radius with age; on average, the outer disk radius in the 7-20~Myr range is three times smaller than that in the $<7$~Myr range. Also, the disk radii inferred from the dust spectral energy distribution (SED) are generally smaller than those derived from the gas line \\citep{Isella2007,Pietu2005}. \\citet{Hughes2008b} suggest a soft outer edge as a solution to the discrepancy. Comparison of CO J=3-2 maps of four disks to different types of disk models strongly supports a soft edge in favor of a sharp cutoff. \\citet{Pietu2007} use the CO and HCO$^+$ lines to probe the radial and vertical temperature profile of the disk. Simple power law disk models and LTE radiative transfer provides best matching results for radial temperature gradients around $r^{-0.5}$. The $^{12}$CO J=2-1 line, $^{13}$CO J=2-1 and $^{13}$CO J=1-0 lines used in their analysis probe subsequently deeper layers and reveal a vertical temperature gradient ranging from 50~K in the higher layers to below the freeze-out temperature of CO in the midplane. This confirms earlier findings by \\citet{Dartois2003}. However, disk masses derived from the CO lines are in general lower than disk masses derived from dust observations \\citep[e.g.][]{Zuckerman1995, Thi2001}. Possible explanations include CO ice formation in the cold midplane and photodissociation in the upper tenuous disk layers. Any single gas tracer alone can only provide gas masses of the species and volume from which it originates, the same way as dust masses derived from a single photometric measurement are only sensitive to grains of a particular size range, namely those grain sizes that dominate the emission at that photometric wavelength. Hence individual gas tracers are more valuable for probing the physical conditions of the volume where they arise than the total disk mass. A combination of suitable gas tracers can then allow us to characterize the gas properties in protoplanetary disks and study it during the planet formation process. This paper aims at exploring the diagnostic power of the fine structure lines of [O\\,{\\sc i}] and [C\\,{\\sc ii}] in the framework of upcoming Herschel observations. Earlier modeling of these lines indicated that they should be detectable down to disk masses of $10^{-5}$~M$_\\odot$ of gas, so also in the very gas-poor debris disks \\citep{Kamp2005}. More recent work by \\citet{Meijerink2008} presents fine structure lines from the inner 40~AU disk of an X-ray irradiated T Tauri disk; the models indicate that the [O\\,{\\sc i}] emission originates over a wide range of radii and depth and is sensitive to the X-ray luminosity. However, most of the C\\,{\\sc ii} and O\\,{\\sc i} line emission comes from larger radii where the gas temperature is dominated by UV heating processes. \\citet{Jonkheid2007} find from thermo-chemical models of UV dominated Herbig Ae disks that the [O\\,{\\sc i}] lines are generally a factor 10 stronger that the [C\\,{\\sc ii}] line. We started to explore the origin of the fine structure lines in \\citet{Woitke2009a} and find that the [C\\,{\\sc ii}]~157.7~$\\mu$m line probes the upper flared surface layers of the outer disk while the [O\\,{\\sc i}]~63.2~$\\mu$m line originates from the thermally decoupled surface layers inward of about 100~AU, above $A_V \\approx 0.1$. The latter line is very sensitive to the gas temperature and might be used to distinguish between hot ($T_{\\rm gas} \\approx 1000$~K) and cold ($T_{\\rm gas} = T_{\\rm dust}$) disk atmospheres. Since the fine structure lines generally originate from a wider radial and vertical range than for example the $^{12}$CO rotational lines, they are potentially more suitable gas mass tracers. We use in this paper the disk modeling code {{\\sc ProDiMo}} presented in \\citet{Woitke2009a} to study the gas line emission from disks around Herbig Ae stars. The main focus are the fine-structure lines of C\\,{\\sc ii} and O\\,{\\sc i} which will be observed for a large sample of disks during the {{\\sc Herschel}} open time Key Program {{\\sc Gasps}} (Gas evolution in protoplanetary systems: http://www.laeff.inta.es/projects/herschel). The disk parameters were chosen to resemble the disk around MWC480. According to previous work by \\citet{Mannings1997b}, \\citet{Thi2001} and \\citet{Pietu2007}, the disk around this star extends from $\\approx 0.5$~AU to 700~AU. We choose here a surface density profile $\\Sigma \\sim r^{-1.0}$. The central star is an A2e Herbig star with a mass of 2.2~M$_\\odot$ and an effective temperature of $8500$~K. Section~\\ref{prodimo} gives a short summary of the disk modeling approach. The line radiative transfer method, re-gridding and the atomic input data are described in Sect.~\\ref{LineRadTrans}. We then briefly discuss some basic properties of the Herbig Ae disk models (Sect.~\\ref{resultsdisks}) before we present the fine-structure lines (Sect.~\\ref{resultslines}) and conclude with a discussion of the diagnostic strength of fine structure line ratios and a comparison to previous ISO and submm observations (Sect.~\\ref{discussion}). ", "conclusions": "We used here a limited series of Herbig Ae disk models computed with the new thermo-chemical disk code called {\\sc ProDiMo} to study the origin and diagnostic value of the gas line tracers [C\\,{\\sc ii}], [O\\,{\\sc i}] and CO. We do not include in this study effects of X-ray irradiation, grain settling or mixing. Even though X-rays are generally of minor importance for Herbig stars, grain settling and large scale mixing processes could affect the conclusions. The effect of grain settling will be included in a forthcoming larger model grid, but mixing processes need to be addressed with dynamical time-dependant models. The Monte-Carlo radiative transfer code {\\sc Ratran} is used to compute line profiles and integrated emission from various gas lines. The main results are: \\begin{itemize} \\item The [C\\,{\\sc ii}] line originates in the disk surface layer where gas and dust temperatures are decoupled. The total line strength is dominated by emission from the disk outer radius. Thus the 157.7~$\\mu$m line probes mainly the disk extension and outer disk gas temperature. The line forms in LTE. \\item The [O\\,{\\sc i}] lines originate also in the disk surface layer even though somewhat deeper than the [C\\,{\\sc ii}] line. The main contribution comes from radii between 30 and 100~AU. The [O\\,{\\sc i}] lines form partially under NLTE conditions. Differences in line emission from escape probability and Monte Carlo techniques are smaller than 10\\%. \\item CO submm lines are optically thick down to very low disk masses of $<10^{-4}$~M$_\\odot$ and form mostly in LTE. $T_{\\rm gas}=T_{\\rm dust}$ is not a valid approximation for these lines. Differences in line emission from escape probability and Monte Carlo techniques are smaller than 3\\%, except in the case of very optically thin disk models ($10^{-5}$ and $10^{-6}$~M$_\\odot$). \\item The [O\\,{\\sc i}]\\,63/145 $\\mu$m and [O\\,{\\sc i}]\\,63/[C\\,{\\sc ii}]\\,158 $\\mu$m line ratios trace disk mass in the regime between $10^{-2}$ and $10^{-6}$~M$_\\odot$. Since the [C\\,{\\sc ii}]\\,158 $\\mu$m line is very sensitive to the outer disk radius, the [O\\,{\\sc i}]\\,63/[C\\,{\\sc ii}]\\,158 $\\mu$m is degenerate in that respect and its use requires additional constraints from ancilliary gas and/or dust observations. The sensitivity of these two line ratios to the dust grain sizes underlines the importance of using SED constraints along with the gas modeling to mitigate the uncertainty of dust properties. \\item A combination of the [O\\,{\\sc i}]\\,63/145 $\\mu$m and [O\\,{\\sc i}]\\,63/[C\\,{\\sc ii}]\\,158 $\\mu$m line ratios can be used to diminish the degeneracy caused by an unknown outer disk radius. \\item Neither total CO submm line fluxes nor line ratios can be used to measure the disk mass. However, the low rotational lines studied here provide an excellent tool to measure the disk outer radius and can thus help to mitigate the degeneracy between gas mass and outer radius found for the [O\\,{\\sc i}]\\,63/[C\\,{\\sc ii}]\\,158 $\\mu$m line ratio. \\end{itemize}" }, "0911/0911.4794_arXiv.txt": { "abstract": "{} {We seek the long-term variations close to the length of a solar cycle in the mean meridional motion of sunspot groups (a proxy of the meridional plasma flow).} {Using the largest set of available reliable sunspot group data, the combined Greenwich and Solar Optical Observation Network sunspot group data during the period 1879\\,--\\,2008, we determined variations in the mean meridional motion of the sunspot groups in the Sun's whole northern and southern hemispheres and also in different $10^\\circ$ latitude intervals. We determined the variations from the yearly data and for the sake of better statistics by binning the data into 3\\,--\\,4 year moving time intervals (MTIs) successively shifted by one year. We determined the periodicities in the mean meridional motion from the fast Fourier transform (FFT) power spectrum analysis. The values of the periodicities are determined from the maximum entropy method (MEM) and the temporal dependencies of the periodicities are determined from the Morlet-wavelet analyses.} {We find that the mean meridional motion of the spot groups varies considerably on a time scale of about 5\\,--\\,20 years. The maximum amplitude of the variation is about 10\\,--\\,15 m s$^{-1}$ in both the northern and the southern hemispheres. Variation in the mean motion is considerably different during different solar cycles. At the maximum epoch (year 2000) of the current cycle~23, the mean motion is relatively strong in the past 100 years and northbound in both the northern and the southern hemispheres. This abnormal behavior of the mean motion may be related to the low strength and the long duration of the current cycle, and also to the violation of the Gnevyshev and Ohl rule by the cycles pair 22,23. The power spectral analyses suggest the existence of $\\approx$ 3.2- and $\\approx$ 4.3-year periodicities in the mean motion of the spot groups in the southern hemisphere, whereas a 13\\,--\\,16 year periodicity is found to exist in the mean motion of the northern hemisphere. There is strong evidence for a latitude-time dependency in the periodicities of the mean motion. The north-south difference in the mean motion also varies by about 10 m s$^{-1}$. During the recent cycles, the north-south difference is negligibly small. Approximate 12- and 22-year periodicities are found to exist in the north-south difference. The implications of all these results are briefly discussed.} {} ", "introduction": "The study of the variations in the meridional flows is important for understanding the underlying mechanism of the solar cycle (Babcock 1961; Ulrich \\& Boyden 2005). Surface Doppler measurements and helioseismology measurements of the surface and the subsurface flows (e.g., Ulrich \\& Boyden 2005; Gonz\\'alez Hern\\'andez et al. 2008) suggest poleward flows of about 10\\,--\\,20 m s$^{-1}$ and a considerable north-south asymmetry in the meridional flow. Surface Doppler measurements are available for about 3 cycles (Ulrich \\& Boyden 2005). Doppler measurements suffer from errors caused by the scatter light and the B-angle influences (Beckers 2007). The motions of many magnetic tracers, particularly sunspots, have been used for a long time as a proxy of the fluid motions to study the solar rotational and the meridional flows (Schr\\\"oter 1985; Javaraiah \\& Gokhale 2002). The sunspot data have been available for more than 100 years. However, the derived rates of rotation and meridional flows depend on the method of the selection of the spots or the spot groups (e.g., Howard et al. 1984; Balthasar et al. 1986; Zappal\\'a \\& Zuccarello 1991; Zuccarello 1993; Javaraiah \\& Gokhale 1997a; Javaraiah 1999; Hiremath 2002; Sivaraman et al. 2003). Proper motions and evolutionary factors of the spot groups may also influence the derived rates of the flows to some extent. The proper motions are random in nature, hence their effect can be reduced with the use of a large data set. Recently, Ru\\v{z}djak et al. (2005) have found that evolutionary factors of the sunspot groups in the determination of the positions of the spot groups is small in the estimated mean meridional motion of the spot groups. A number of scientists studied the solar cycle variations of the mean meridional motion of the spot groups (see Javaraiah \\& Ulrich 2006). Since yearly data are inadequate for this purpose, particularly around a solar cycle minimum, some authors have used the superposed epoch analysis of the data during a few or all cycles for which the data were available and studied the mean solar cycle variation (Balthasar et al. 1986; Howard \\& Gilman 1986). Recently, by using the same method, Javaraiah \\& Ulrich (2006) analyzed the combined Greenwich and Solar Optical Observation Network sunspot group data during the period 1879\\,--\\,2004 and determined the mean solar cycle variation in the meridional motions of the sunspot groups. In that early paper, the spot group data during the cycles 12\\,--\\,20 were superposed according to the minima of these cycles. This yielded an average solar cycle variation of the meridional motion over about 5\\,--\\,9 cycles, which suggests that only around the end of a solar cycle the meridional motion is considerably significant from zero and it is poleward in both the northern and the southern hemispheres. However, during some individual cycles the variations may be considerably different from this average solar cycle variation. Javaraiah \\& Ulrich (2006) also determined the cycle-to-cycle variation of the mean (over the duration of a cycle) meridional motion of the spot groups during cycles 12\\,--\\,23 and found the existence of a weak long-period cycle (Gleissberg cycle) in the cycle-to-cycle variation of the mean motion. Since the meridional speed is negligibly small or zero in some phases of a cycle or even with opposite signs during different phases of some cycles, hence, the motions are washed out in the average over the cycle. In order to get rid of this problem in the study of long-term variations in the mean meridional motion, it is necessary to analyse the spot or the spot group data in the intervals considerably shorter than the length of a solar cycle. In the present paper we have analyzed the annual spot group data during 1879\\,--\\,2008 and determined the variations in the mean meridional motions of the spot groups in the northern and the southern hemispheres. As expected the statistics is poor in case of the results derived from the annual data, particularly at the cycles minima. We have taken some additional precautions, so it is possible to see the patterns of the variations when they are close to a solar cycle length, even in the annual time series. However, we have also determined the variation in the mean meridional motion of the spot groups by binning the spot group data into the moving-time intervals (MTIs) of lengths (3\\,--\\,4 years) considerably greater than a year, but reasonably smaller than the length of a cycle; $i.e.$, less than the half of the length of a cycle. In such a time-series which comprised the longer time intervals, it is relatively easy to detect the long-term variations near the length of a solar cycle (Javaraiah \\& Gokhale 1995, 1997b). In addition, the sizes of these series reasonably large. Hence, it enabled us to find the periodicities that approach the length of a solar cycle from the power spectral analysis. In the next section we describe the data and analysis. In Sect.~3 we present the variations in the mean meridional motion of the spot groups in the whole northern and the southern hemispheres, as well as in different $10^\\circ$ latitude intervals--and the corresponding differences between the whole northern and the southern hemispheres--during the period 1879\\,--\\,2008 and point out their important features. In the same section, we show the periodicities in the mean meridional motion of the spot groups from the traditional FFT analyses. From the MEM analyses, we determined the values of the periodicities, and from the Morlet-wavelet analyses we determined the temporal dependencies of the detected periodicities. In Sect.~4 we summarize the results and the conclusions, and briefly discuss the implications of these results for understanding the solar long-term variability. \\vspace{0.3cm} ", "conclusions": "From the analysis of the largest available reliable sunspot group data; $i.e.$, the combined Greenwich and SOON sunspot group data during the period 1879\\,--\\,2008, we find the following. \\begin{enumerate} \\item The mean meridional motion of the sunspot groups varies considerably on 5\\,--\\,20 year timescales during the period 1879\\,--\\,2008. The maximum amplitude of the variation is 10\\,--\\,15 m s$^{-1}$. \\item The pattern and amplitude of the solar cycle variation in the mean motion are significantly different during the different cycles. During the maximum epoch (year 2000) of the current cycle, the mean motion is relatively stronger than in the past $\\approx$100 years, and it is northbound in both the northern and the southern hemispheres. \\item The north-south difference (north-south asymmetry) in the mean meridional motion of the spot groups also varies with a maximum amplitude of about 10 m s$^{-1}$. The north-south difference was considerably larger during the early cycles (with a strong contribution from the high latitudes). It is negligible during the recent cycles. \\item Power spectral analyses suggest that $\\approx$ 3.2- and $\\approx$ 4.3-year periodicities exist in the mean meridional motion of the spot groups of the southern hemisphere, whereas a 13\\,--\\,16 year periodicity is found in the mean motion of the spot groups of the northern hemisphere. \\item The $\\approx$ 12- and $\\approx$ 22-year periodicities are found to exist in the north-south difference of the mean motion. \\item There is a considerable latitude-time dependence in the periodicities of the mean meridional motion of the spot groups. There is a strong suggestion that, in the $10^\\circ$\\,--\\,$20^\\circ$ latitude-interval of the northern hemisphere, a periodicity slowly evolved from $\\approx$ 16 year to $\\approx$ 10 year, over the period 1880\\,--\\,2007, and it evolved in the opposite way, $\\approx$ 10 year to $\\approx$ 16 year, in $20^\\circ$\\,--\\,$30^\\circ$ latitude interval. \\end{enumerate} The behavior of the mean motion of the spot groups in cycle~23 is similar to that of cycle~14, which is a low-amplitude (lowest in the last century) and considerably long-duration cycle. Cycle~23 is also a relatively low-amplitude and long-duration cycle, so that the result above (conclusion (2)) may be a part of a real long-term behavior in the mean meridional motion of the spot groups; $i.e.$, most probably it is not an artifact of the differences (if any) between the Greenwich and the SOON datasets, within the continuous time series of the combined dataset used here. Most of the helioseismic measurements of the meridional flows during the current sunspot cycle~23 suggest an increase in the amplitudes of the surface and the subsurface poleward meridional flows with a decrease in magnetic activity (Gonz\\'alez Hern\\'andez et al. 2008), whereas we find a strong northbound mean meridional motion of the spot groups during the maximum of cycle~23. A reason for this discrepancy may be that sunspot motions may not represent the Sun's plasma motions (D'Silva \\& Howard 1994), or the motions of the magnetic structures may represent the motions of the deeper layers of the Sun's convection zone where these structures are anchored (Javaraiah \\& Gokhale 1997a; Hiremath 2002; Sivaraman et al. 2003; Meunier 2005). In addition, the mean meridional motion of the sunspot groups may only represent the mean solar meridional plasma motion at low and middle latitudes, because sunspots data are confined to only these latitudes. The magnetic structures of the only large spot groups during their initial days might be anchored near the base of the convection zone (Javaraiah \\& Gokhale 1997a; Hiremath 2002; Sivaraman et al. 2003), hence might have largely equatorward motions (Javaraiah 1999). While rising through the convection zone, the magnetic structures of the large spot groups may be fragmented into the smaller structures (Javaraiah 2003; Sch\\\"ussler \\& Rempel 2005). The small structures may move mainly toward the poles (\\v{S}vanda et al. 2007). However, as can be seen in Figs.~1\\,--\\,3 there are also equatorward motions (may be due to an effect of the reverse meridional flows), mainly near minima of the cycles where a spot group is relatively small. Meridional flows can transport magnetic flux and cause cancellation/enhancement of magnetic flux, and it is believed that poleward meridional flows play a major role in the polarity reversals of the polar magnetic fields ($e.g.$, Wang 2004). The $\\approx$ 12-year and the $\\approx$ 22-year periodicities of the north-south difference in the mean meridional motion of the spot groups may have a close relationship with the 11-year solar activity (the emerging magnetic flux) cycle and the 22-year solar magnetic cycle, respectively. Many of the other periodicities found here also exist in several activity phenomena (Knaack et al. 2005; Song et al. 2009, and references therein) and solar differential rotation determined from sunspot data (Javaraiah \\& Gokhale 1995, 1997b; Javaraiah \\& Komm 1999; Braj\\v{s}a et al. 2006). They may be closely related to the Rossby type waves that were discussed by Ward (1965) and others (e.g., Knaack et al. 2005; Chowdhury et al. 2009). According to the well known Gnevyshev and Ohl rule (G-O rule), an odd cycle is stronger than its immediately preceding even cycle (Gnevyshev \\& Ohl 1948). Cycle pair~22,23 violated the G-O rule. The duration of the current cycle~23 is very long, and during the declining phase of this cycle, the activity in the southern hemisphere is considerably stronger than in the northern hemisphere. All these properties of the cycle~23 could be strongly related to the large and northbound mean meridional motion of the spot groups during this cycle. As already mentioned above, motions of magnetic structures such as sunspots mimic the motions of deeper layers of the Sun (see also Javaraiah \\& Gokhale 2002). Therefore, the magnetic flux cancellation/enhancement due to the mean meridional motion of sunspot groups may take place in the subsurface layers of the Sun. By considering the poleward meridional plasma flows detected by surface Doppler measurements and by helioseismology, the deeper counter-motion (suggested by the mean meridional motion of spot groups) might amplify the action of the near-surface dynamo in the southern hemisphere during cycle~23 for causing stronger magnetic activity on this hemisphere." }, "0911/0911.4107_arXiv.txt": { "abstract": "From high-resolution images of 23 Seyfert-1 galaxies at z=0.36 and z=0.57 obtained with the Near Infrared Camera and Multi-Object Spectrometer on board the {\\it Hubble Space Telescope} (HST), we determine host-galaxy morphology, nuclear luminosity, total host-galaxy luminosity and spheroid luminosity. Keck spectroscopy is used to estimate black hole mass (\\mbh). We study the cosmic evolution of the \\mbh-spheroid luminosity (\\ls) relation. In combination with our previous work, totaling 40 Seyfert-1 galaxies, the covered range in BH mass is substantially increased, allowing us to determine for the first time intrinsic scatter and correct evolutionary trends for selection effects. We re-analyze archival HST images of 19 local reverberation-mapped active galaxies to match the procedure adopted at intermediate redshift. Correcting spheroid luminosity for passive luminosity evolution and taking into account selection effects, we determine that at fixed present-day V-band spheroid luminosity, \\mbh/\\ls $\\propto$ $(1+z)^{2.8 \\pm 1.2}$. When including a sample of 44 quasars out to $z=4.5$ taken from the literature, with luminosity and BH mass corrected to a self-consistent calibration, we extend the BH mass range to over two orders of magnitude, resulting in \\mbh/\\ls $\\propto$ $(1+z)^{1.4 \\pm 0.2}$. The intrinsic scatter of the relation, assumed constant with redshift, is 0.3$\\pm$0.1 dex ($<$0.6 dex at 95\\% CL). The evolutionary trend suggests that BH growth precedes spheroid assembly. Interestingly, the \\mbh-{\\it total host-galaxy} luminosity relation is apparently non-evolving. It hints at either a more fundamental relation or that the spheroid grows by a redistribution of stars. However, the high-z sample does not follow this relation, indicating that major mergers may play the dominant role in growing spheroids above $z \\simeq 1$. ", "introduction": "\\label{sec:intro} Supermassive black holes (BHs) seem to be ubiquitous in the center of spheroids -- elliptical galaxies and classical bulges of spirals \\citep[e.g.,][]{kor95,fer05}. In the local Universe, tight empirical relations have been found between the mass of the BH (\\mbh) and the properties of the spheroid, i.e.~stellar velocity dispersion $\\sigma$ \\citep{fer00,geb00}, stellar mass \\citep[e.g.,][]{mar03}, and luminosity \\citep[e.g.,][]{har04}. The tightness of these relations is surprising, given the very different scales involved -- from accretion onto the BH ($\\mu$pc scale), the dynamical sphere of influence of the BH (pc scale) to the size of the spheroid (kpc scale) -- and poses a challenge to any theoretical model explaining their origin. In general, the correlations are believed to indicate a close connection between galaxy formation and evolution and the growth of the BH. A variety of theoretical models have been developed to explain the observed relations, involving galaxy mergers and nuclear feedback through quenching of star formation \\citep[e.g.,][]{kau00,vol03,cio07,hop07,dim08,hop09b}. Measuring the evolution with redshift of these correlations constrains theoretical interpretations and provides important insights into their origin \\citep[e.g.,][]{cro06,rob06,hop07}. For quiescent galaxies, the biggest challenge is to measure the BH mass, given the pc-scale sphere of influence of the BH which needs to be resolved spatially through either gas or stellar dynamics (see \\citealt{gue09} and \\citealt{gra08} for a recent compilation and references therein; for a review see \\citealt{fer05} and references therein) or from X-ray spectroscopy probing the existence of a central temperature peak of the interstellar medium \\citep[e.g.,][]{bri99,hum06,hum08}. With current technology, direct quiescent black hole mass measurements are thus limited to nearby galaxies. For active galaxies, for which nuclear luminosity is comparable to or larger than that of the host galaxy, the situation is virtually the opposite. Estimating BH masses within a factor of 2-3 is fairly straightforward through empirically calibrated relations based on spectroscopic data measuring the kinematics of the broad-line region (BLR) \\citep[e.g.,][]{wan99,woo02,ves02,ves06,mcg08}. Unfortunately, the active galactic nuclei (AGN) often outshines the host galaxy, making it difficult to disentangle nuclear and host-galaxy light for an accurate measurement of the spheroid luminosity. Also, measuring $\\sigma$ from stellar absorption lines is hampered by the contaminating AGN continuum and emission lines. Different groups have tackled these problems in distinct ways, e.g.~by using the [\\ion{O}{3}] emission line width as surrogate of $\\sigma$ \\citep[e.g.,][]{shi03}, or by using gravitational lensing to super-resolve the host galaxies of quasars \\citep[e.g.,][]{pen06a,pen06b}. Our group \\citep{tre04,woo06,tre07,woo08} has focused on Seyfert-1 galaxies - for which the nucleus is not as bright as for quasars - at moderate redshifts ($z=0.36$ and $z=0.57$, corresponding to look-back times of $\\sim$4-6 Gyrs). The non-negligible stellar light produces strong enough absorption lines to measure $\\sigma$ from unresolved spectra, as shown by \\citet{tre04} and \\citet[][hereafter Paper I \\& III]{woo06, woo08}. At the same time, high resolution {\\it Hubble Space Telescope} (HST) imaging allows for an accurate determination of the AGN luminosity (for an unbiased estimate of nuclear luminosity and hence \\mbh) and spheroid luminosity (to create the \\mbh-\\ls~relation; \\citealt[][hereafter Paper II]{tre07}). We are thus able to simultaneously study both the \\mbh-$\\sigma$ and \\mbh-\\ls~relations, allowing us to distinguish mechanisms causing evolution in $\\sigma$ (e.g., dissipational merger events) and \\ls~(e.g.~through passive evolution due to aging of the stellar population, or dissipationless mergers). Results presented in Paper I, II, and III suggest an offset with respect to the local relationships, which cannot be accounted for by known systematic uncertainties. At a given \\mbh, in the range 10$^8$-10$^9$ M$_{\\odot}$, spheroids had smaller velocity dispersion and spheroid mass 6 Gyrs ago ($z\\sim0.57)$, consistent with recent growth and evolution of intermediate-mass spheroids. Paper II concludes that the distant spheroids have to grow by $\\sim$60\\% in stellar mass ($\\Delta \\log M_{\\rm sph}$ = 0.20 $\\pm$ 0.14) at fixed black hole mass in the next 4 billion years to obey the local scaling relations if no significant BH growth is assumed, consistent with the relatively low Eddington ratios. Indeed, the HST images reveal a large fraction of merging or interacting systems, suggesting that gas rich mergers will be responsible for the spheroid growth. Although tantalizing, the results presented in our previous papers suffer from several limitations. Samples were small, and the local comparison sample of Seyferts measured in a self-consistent manner was even smaller than the distant sample, thus contributing substantially to the overall error budget. The limited range in black hole mass was insufficient to determine independently the offset of the scaling relation and its scatter, while taking into account selection effects. If the \\mbh-$\\sigma$ and \\mbh-\\ls~relations of active galaxies were not as tight as those for quiescent ones, selection effects could be mimicking evolutionary trends \\citep{tre07,lau07,pen07}. To overcome these limitations, we have now doubled the sample size (from 20 in Paper II to 40 total here) and expanded the covered range of BH masses to lower masses (from $\\log$ \\mbh/$M_{\\odot} = 8-8.8$ in Paper II to $\\log$ \\mbh/$M_{\\odot} = 7.5-8.8$ here). We focus on the resulting BH mass - spheroid luminosity relation. The BH mass - $\\sigma$ relation will be presented in a separate paper (Woo et al.\\ 2009, in preparation). We also analyze archival HST images of the sample of local Seyferts with reverberation-mapped (RM) \\mbh~in the same way as our intermediate-z objects, to eliminate possible systematic offsets. Finally, we combine our results with data compiled from the literature and treated in a self-consistent manner to extend the redshift range over which we study evolution. For conciseness the three samples will be referred to as ``intermediate-redshift'' sample, ``local'' sample, and ``high-redshift'' sample, respectively. The paper is organized as follows. We summarize the properties of our intermediate-redshift Seyfert sample, observations, data reduction, and analysis in \\S~\\ref{sec:sample},~\\ref{sec:obs}, and~\\ref{sec:surface}. \\S~\\ref{sec:quan} summarizes the derived quantities, including the derivation of \\mbh~from Keck spectra. In \\S~\\ref{sec:comp}, we describe the local comparison sample consisting of reverberation-mapped AGNs, re-analyzed here, as well as the high-redshift comparison sample taken from the literature, calibrated for consistency with the other samples. We present our results in \\S~\\ref{sec:res}, including host-galaxy morphology and merger rates, the evolution of the \\mbh-\\ls~relation, a full discussion and treatment of selection effects, and a relation between BH mass and host-galaxy luminosity. We discuss the possible implications of our findings for the origin and evolution of the BH mass scaling relation in \\S~\\ref{sec:disc}. A summary is given in \\S~\\ref{sec:sum}. In Appendix~\\ref{sec:mc}, we describe Monte Carlo simulations used to probe our analysis and determine errors. Appendix~\\ref{sec:sersic} discusses the choice of the S{\\'e}rsic index in the adapted 2D surface-brightness fitting procedure. Details on the re-analysis of the HST images of the local RM AGNs are given in Appendix~\\ref{sec:rm}. Throughout the paper, we assume a Hubble constant of $H_0$ = 70\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_{\\Lambda}$ = 0.7 and $\\Omega_{\\rm M}$ = 0.3. Magnitudes are given in the AB system \\citep{oke74}. ", "conclusions": "\\label{sec:disc} \\subsection{The role of mergers} \\label{ssec:mergerd} Theoretical studies generally invoke mergers to explain the observed scaling relations between BH mass and host-galaxy spheroid properties -- a promising way to grow both spheroid and BH. In a simple scenario, spheroids grow by (i) the merging of the progenitor bulges (assuming that both progenitors have a spheroidal component), (ii) merger-triggered starbursts in the cold galactic disk, and (iii) by transforming stellar disks into stellar spheroids \\citep[e.g.,][]{bar92,mih94,cox04}, thus increasing the spheroid luminosity and stellar velocity dispersion. The fueling of the BH, on the other hand, is triggered by the merger event as the gas loses angular momentum, spirals inward and eventually gets accreted onto the BH, giving rise to the bright AGN or 'quasar' period in the evolution of galaxies \\citep[e.g.,][]{kau00,dim05}. Eventually, if BHs are present in the center of both progenitor galaxies, they may coalesce. In such a simple scenario, an evolution in the BH mass - spheroid luminosity relation is not necessarily expected: Both spheroid and BH grow from the same gas reservoir, and bulge stars added to the final spheroid followed the BH mass - spheroid luminosity relation prior to merging, so the relation will be preserved when the BHs coalesce. However, while mergers provide a way to grow both spheroids and BHs, they may do so on very different timescales. Moreover, the merger history of galaxies varies, depending e.g.~on formation time and environment. Different types of merger, for example with a different relative role of dissipation \\citep[e.g.,][]{hop09a} have different effects on the growth of spheroid and BH: For a gas-rich major merger between an elliptical galaxy and a spiral galaxy - the latter without a (massive) BH -- the bulge grows more efficiently than the BH by the disruption of the stellar disk \\citep{cro06}. In general, our images of the intermediate-z Seyfert galaxies support the merger scenario (see Fig.~\\ref{nicmos} and \\S~\\ref{ssec:merger}). However, objects with evidence for merger/interaction do not form any particular outliers in the BH mass - spheroid luminosity relation (Fig.~\\ref{bhlv}). This may not be too surprising: For those objects for which we still see two separate galaxies in the process of merging, we fitted both separately and the bulge luminosity of the AGN host has not yet increased from the process of merging. Other objects with signs of interaction may be in a later evolutionary stage where the bulge luminosity has already increased and thus, the object falls closer to the local relation. Finally, mergers between similar objects would only move the system parallel to the local relation. In general, the effect of mergers on the measured bulge luminosity of an object depends on the type of the merger, the evolutionary stage of the merger, and the timescales involved to grow spheroid and BH. Such a detailed comparison of merger type and age with theoretical predictions is beyond the scope of this paper, given the small sample of merging objects and the limited information at hand. Note that the fraction of apparently disturbed systems we find is not higher than that of a comparison sample of inactive galaxies at the same redshift (\\S\\ref{ssec:merger}). Thus, from our images alone, we cannot infer a causal link between a merger/interaction event and the AGN activity we observe. Instead, ``normal'' galaxies may have the same merger history, and ongoing interactions are not necessarily predictive of AGN activity. The role of mergers for the fueling of AGNs is debated in the literature \\citep[e.g.][]{san88,hec84,hut88,dis95,bah97,mcl99,can01,dun03,flo04,can07,urr08,ben08,vei09,tal09}. While there is little doubt that mergers are helpful, they are certainly not a sufficient condition, considering the numerous inactive interacting galaxies.\\footnote{However, this might also be due to the timescales involved, with the signs of interaction outliving the AGN activity (see also the case of present-day Type II AGNs; \\citealt[e.g.][]{cho09}).} Also, mergers may be necessary for the high-luminosity QSOs only while for Seyfert galaxies, secular evolution through processes such as bar instabilities may be the dominant effect in the evolution of these galaxies. We will come back to this issue in \\S~\\ref{ssec:mbhtotr}. \\subsection{BH Mass - Spheroid Luminosity Relation} \\label{ssec:mbhlr} Combining results of low-z, intermediate-z and high-z AGNs, treated in a self-consistent manner, we can estimate the intrinsic scatter of the \\mbh-\\ls~scaling relation and correct evolutionary trends for selection effects. We discuss scatter and evolution in the next two subsections. \\subsubsection{Scatter of \\mbh-\\ls} \\label{ssec:scatter} The intrinsic scatter we find (0.3$\\pm$0.1 dex; $<$0.6 dex at 95\\% CL) is non-negligible. However, we assume the intrinsic scatter of the \\mbh-\\ls~relation to be non evolving. While it would be desirable to directly study the evolution of the scatter with redshift, this requires a larger sample than the one we have at hand. Actually, we might expect a larger intrinsic scatter at higher redshifts, given the different ways and timescales involved when growing spheroids and BHs through mergers. Indeed, for the local Universe, the observed tightness in the relations has been a challenge for theoretical studies. It has been explained by self-regulated models of BH growth \\citep{hop09b} in which the energetic feedback of the AGN eventually halts accretion, preventing the BH from further growth and quenches star formation \\citep[e.g.,][]{cio97,cio01,sil98,mur05,dim05,saz05,hop05,spr05,dim08}. Also, a significant fraction of the host galaxies of both our local RM AGN sample ($\\sim$9/19) and our intermediate-z sample ($>$15/40) are prominent late-type spirals of type Sa or later which have been found to have a larger intrinsic scatter than elliptical galaxies \\citep[e.g.,][for the \\mbh-$\\sigma$ relation: 0.44 dex when including all galaxies vs. 0.31 for elliptical galaxies only]{gue09}. As already discussed in paper II, the intermediate-z late-type spirals may eventually fall on the local relation later, through merging, in line with ``downsizing'' \\citep[e.g.~][]{cow96,bri00,kod04,bel05,noe07}: Less massive, blue galaxies merge at later times and arrive at the local relation by becoming larger, bulge-dominated red galaxies. Also, at least some spiral galaxies may not have classical bulges, but pseudobulges which are characterized by surface-brightness profiles closer to exponential profiles, ongoing star formation or starbursts, and nuclear bars or spirals. It is generally believed that they have evolved secularly through dissipative processes rather than being formed by mergers \\citep[see e.g. review in][]{kor04}. BHs have been found to reside in galaxies without classical bulges which may not follow the same scaling relations \\citep[e.g.,][]{gre08}. \\subsubsection{Evolution of \\mbh-\\ls~with redshift} \\label{ssec:evor} To generalize our results and to facilitate comparison with theoretical and observational works, it is useful to estimate the evolution of the \\mbh - spheroid {\\it stellar mass} relation. We can convert the observed evolution of \\mbh~- spheroid luminosity into that between \\mbh~and spheroid mass, if we assume that -- after correction for luminosity evolution -- the mass-to-light ratio does not change from sample to sample\\footnote{Unfortunately, spatially resolved color information for a more sophisticated estimation of the stellar mass of the bulge is currently not available.}. Under this assumption, an offset of $\\Delta$\\mbh~at fixed \\ls~equals that at fixed $M_{\\rm star}$ and thus, \\mbh/\\ms~$\\propto$$(1+z)^{1.4 \\pm 0.2}$. We are now in a position to make a broad range of comparisons. In the literature, the BH mass evolution is discussed quite controversially. \\citet{shi03} study the \\mbh-$\\sigma$ relation out to $z=3.3$, estimating \\mbh~from H$\\beta$ and $\\sigma$ from [\\ion{O}{3}] and find that the QSOs and their host galaxies follow the local relation. (Note, however, that using [\\ion{O}{3}] as a surrogate for $\\sigma$ can be problematic as [\\ion{O}{3}] is known to often have an outflow component; for a discussion see e.g. \\citet{gre05,kom07}.) A similar conclusion has been reached by \\citet{shen08} who study over 900 broad-line AGNs out to $z \\simeq 0.4$ from SDSS. \\cite{ade05} use the correlation length of 79 quasar hosts at $z \\sim 2-3$ to estimate the virial mass of the halo and the \\ion{C}{4} line width and UV flux at 1350\\AA~to estimate \\mbh. When comparing the resulting \\mbh~-$M_{\\rm halo}$ relation to the local one \\citep{fer02}, they find no evidence for evolution. In particular, they can rule out evolution of the form \\ms/\\ms~$\\propto$$(1+z)^{2.5}$ with $z=2.5$ at 90\\% CL, given their error bars. Other observational studies find the same trend in evolution as we do, i.e.~that BHs are too massive for a given bulge mass or velocity dispersion at higher redshifts \\citep{wal04,shi06,mcl06,pen06b,sal07,wei07,rie08,rie09,gu09}. \\citet{mcl06}, for example, study radio-loud AGN ($0 < z < 2$) and find \\mbh/\\ms$\\propto$$(1+z)^{2.07 \\pm 0.76}$. \\citet{pen06b}, whose data, treated in a consistent manner to match our data set, are included in this study, rule out pure luminosity evolution and find that the ratio between \\mbh~and $M_{\\rm sph}$ was $\\sim$four times larger at $z \\sim 2-3$ than today. For 89 broad-line AGNs between $1 1.3$ at more than 99\\% CL (with the possibility of a stronger evolution for the more massive BHs). Another study using fully cosmological hydrodynamic simulations of $\\Lambda$CDM following the growth of galaxies and supermassive BHs, as well as their associated feedback processes, finds only limited evolution in \\mbh~with a steepening at z=2-4 \\citep{dim08}. \\citet{mer04} expect a weak evolution of \\mbh/\\ms~$\\propto$ (1+z)$^{0.4-0.6}$, when fitting the total stellar mass and star formation rate density as a function of redshift and comparing that to the hard X-ray selected quasar luminosity function, assuming that BHs only grow through accretion. Such a slope is in agreement with work by \\citet{hop09a} who combine prior observational constraints in halo occupation models with libraries of high-resolution hydrodynamic simulations of galaxy mergers. Using semi-analytic models, \\citet{cro06} predicts an evolution of \\mbh/\\ms~$\\propto$ (1+z)$^{0.4-1.2}$. A more rapid evolution is predicted by \\citet{wyi03} who assume a self-regulated BH growth model and find \\mbh/\\ms~$\\propto$ (1+z)$^{1.5}$, similar to our observational result. However, the great advantage of the study presented here are the high-quality images at hand, allowing for a detailed bulge-to-disk decomposition of the host galaxy of the low- and intermediate-z Seyfert-1 galaxies. Combining data from a large sample of active galaxies, covering a redshift range from the local Universe out to z=4.5, all treated in a consistent manner, results in smaller error bars on the predicted evolution than previous studies. Moreover, it allows, for the first time, to correct evolutionary trends for selection effects. The evolution we find (\\mbh/\\ms$\\propto$$(1+z)^{1.4 \\pm 0.2}$) is indicative of BH growth preceding host-spheroid assembly. Still, we did not take into account that the evolution may depend on BH mass (see also Paper III). Indeed, there are theoretical predictions that objects with higher BH (or bulge) masses evolve faster \\citep[e.g.,][]{hop09a}. For example, \\citet{dim08} find that when restricting their fits to objects with $M_{\\star}\\geq 5\\times 10^{10}\\ M_{\\odot}$, the relation has a slope of $\\sim 1.9$ at z=3-4 and $\\sim 1.5$ at z=2. Unfortunately, our sample is too small to allow us to address this possibility. There may indeed be some evidence that the offset in BH mass is larger for objects with more massive BHs (Fig.~\\ref{bhlv}, upper left panel and Fig.~\\ref{offset}, right panel). \\subsection{BH Mass - Galaxy Luminosity Relation} \\label{ssec:mbhtotr} A different scenario seems to emerge when considering the relation between \\mbh~and {\\it total host-galaxy} luminosity (Fig.~\\ref{bhlv}, lower left panel). This relation is almost non-evolving within the last six billion years.\\footnote{Note that there is insufficient information to constrain the intrinsic scatter.} Recently, \\citet{jah09} found qualitatively similar results for a small sample of ten AGNs at redshifts between 1 $< z <$ 2: They derive host-galaxy masses from colors based on ACS and NICMOS imaging, finding that they lie on the \\mbh-$M_{\\star, bulge}$ relation in the local Universe \\citep{har04}. Such a non-evolving \\mbh-\\lh~relation can be interpreted twofold. (a) The amount by which some of the more distant objects have to grow their spheroid is already contained within the galaxy itself, and the growth can be achieved by the redistribution of stars, i.e.~transforming disk stars into bulge stars. Such a redistribution can be the result of mergers or secular evolution, e.g.~bar instabilities \\citep[e.g.,][]{com81,vanb98,avi05,deb06} and torque-driven accretion \\citep[see e.g. review in][]{kor04} which may coincidentally be also the triggering mechanism for the BH activity we observe \\citep[e.g.,][]{shl93,ath03,dum07,haa09}. While not every object in the intermediate-z sample will experience a major merger in the last 4-6 billion years, secular evolution is a promising alternative way to grow the spheroidal components in these objects. But even if they do experience a major merger (as indeed evidenced for at least some objects in our sample), the role of the merger depends on the merger type as discussed above (e.g.~a merger between similar objects will simply move the system along the local relation). There may again be a dependency on BH mass: For the low-mass objects, the offset becomes almost negative, indicating that in the low-mass range, either the BH is, at the same time, still growing by a non-negligible amount (consistent with the higher Eddington ratio in the low-mass regime\\footnote{Of course, we may be biased against low-mass objects with low Eddington ratios.}) or that not all of the stellar mass will end up in the spheroid component. Indeed, for local RM AGNs, at least 6/19 objects reside in late-type host galaxies (preferentially those with lower BH masses). (b) The relation between BH mass and host-galaxy luminosity (or mass) may be the more fundamental one. Indeed, this is predicted by \\citet{pen07}: In his thought experiment, he shows that a tight linear relation between \\mbh~and host-galaxy mass can evolve -- if the galaxy mass function declines with increasing mass -- due to ``a central-limit-like tendency for galaxy mergers, which is much stronger for major mergers than for minor mergers, and a convergence toward a linear relation that is due mainly to minor mergers''. Also, it is possible that BHs in late-type galaxies or galaxies without classical bulges, while not following the same \\mbh~scaling relations as spheroids (see discussion in \\S~\\ref{ssec:scatter}), they instead obey a more fundamental relation between BH mass and host-galaxy mass. However, the relation between host-galaxy luminosity and \\mbh~seems to exist only up to z$\\lesssim$1: The offset for the high-z comparison sample does not decrease as the luminosity given by \\citet{pen06b} is already the total host-galaxy luminosity \\citep[the same is true also for the results by][]{mer09,dec09}. Along the line of argument of (a) above, the growth of the spheroid above a redshift of $z \\gtrsim 1$ cannot simply be achieved through secular evolution (with quasars being predominantly hosted by ellipticals), but instead, major mergers are needed. A major merger is more likely to happen for the high-z sample given the longer time span. Or, following (b), a relation between BH mass and host galaxy is already at place at $z \\lesssim 1$, but still evolving at earlier times. However, we cannot exclude that part of the difference is due to the difference in BH mass between the samples, with the high-z objects generally having larger BH masses. In the end, the discussion boils down to the following question: What is the dominant mechanism that grows spheroids, and does it depend on spheroid mass and/or redshift? This is debated controversially in the theoretical literature. For example, based on their semi-analytic models, \\citet{par09} find that the majority of ellipticals and spirals never experience a major merger but rather, that they acquire their spheroid stellar mass through minor mergers or disc instabilities. \\citet{hop09c}, on the other hand, combine empirically constrained halo occupation distributions with high-resolution merger simulations, and find that major mergers dominate the formation of $\\sim$$L_{\\star}$ bulges and systems with higher B/T, but that lower-mass or lower B/T systems are preferentially formed by minor mergers. They predict that the major merger rate increases with redshift. Qualitatively, we can reconcile such a scenario with our results: Higher-mass objects and those at higher redshifts (i.e.~the majority of the high-z sample) form their spheroids preferentially through major mergers and are thus still evolving toward a \\mbh-\\lh~relation, while lower-mass and lower-z objects (i.e.~our intermediate-z sample) grow their spheroids through minor mergers or disk instabilities that redistribute the stars and thus, they fall on the \\mbh-\\lh~relation." }, "0911/0911.4325_arXiv.txt": { "abstract": "Just comparing with the scenario that the $(3+1)$-dimensional ``real world'' of the Calabi-Yau compactification has a tremendous landscape, we conjecture that a $(4+1)$-dimensional holographic theory may also hold a landscape of its vacua. Unlike the traditional studies of the AdS/CFT phenomenology where the vacua are always constructive, we discuss the proper holographic vacua and their flux compactification, starting from some general compact Einstein manifolds. The proper vacua should be restricted by (i) a consistent worldsheet theory that possesses the superconformal symmetry, (ii) some definite symmetries to keep/break the corresponding symmetries of the dual field theory, (iii) certain brane/flux configurations to cancel anomalies, and (iv) stabilities. We consider diverse fundamental parameters of the dual field theory, fixed by some special vacuum moduli. In an opposite way, if some field theory such as QCD holds an AdS dual, it may also possesses various fundamental parameters by an ``landscape'' of its vacuum. Different vacua may be adjacent with each other, and divided by domain walls. If the size of a single vacuum region is smaller than the visible universe, it may be testable. We discuss the consequences of this conjecture in the astrophysical environments, include but not limit to: (i) consistency with the critical energy density of the universe, (ii) the behaviors of cosmic rays, (iii) the stability and abundance of deuterons and other nuclei in the big-bang nucleosynthesis and the star burning scenarios, and (iv) the existence of strange/charm stars. \\vspace{0.5cm} \\noindent \\emph{PACS}: 11.25.Wx, 98.80.Cq, 11.25.Tq, 11.25.Mj ", "introduction": "} The anti-de Sitter/conformal field theory (AdS/CFT) correspondence, one of the most ambitious scenarios in string phenomenology, conjectures that a type IIB superstring theory on $\\mathrm{AdS}_5 \\times S^5$ is equivalent with a $\\mathcal{N} = 4$ $U(N_c)$ super Yang-Mills (SYM) theory in four-dimensions~\\cite{Maldacena:1997re}, or more generally a gravity theory on $\\mathrm{AdS}_{p+2} \\times \\mathcal{M}_q$ is dual to a $(p+1)$-dimensional boundary CFT~\\cite{Witten:1998qj,Aharony:1999ti}. The idea of ``holography''~\\cite{Susskind:1998dq} also pushes the applications of AdS/CFT to more realistic environments, such as QCD~\\cite{Peeters:2007ab,Mateos:2007ay}, or condensed matter systems~\\cite{Herzog:2009xv}. Nevertheless, the compact dimensions (thus the ``vacua'') and their stabilization in AdS/CFT models, has seldom been studied systematically. On the one hand, theoretical researches always study some specific vacuum by constructive methods. For example, they break boundary supersymmetry by quotient spaces $S^5 / \\Gamma$~\\cite{Kachru:1998ys,Lawrence:1998ja}, or the conifold construction~\\cite{Klebanov:1998hh,Acharya:1998db,Morrison:1998cs}. On the other hand, phenomenological models which aim to approach the ``real world'' physics, always neglect the discussions of the compact dimensions directly. However, although difficult, the study of AdS/CFT vacua has no alternative but within the framework of flux compactification~\\cite{Grana:2005jc,Douglas:2006es}. Some founding works in this direction can be found in~\\cite{Aharony:2008wz,Polchinski:2009ch}. The original studies of flux compactification, always aim to the Calabi-Yau threefolds. The main reason is that $\\mathrm{CY}_3$, which possesses a special holonomy of $SU(3) \\subset SO(6)$, can reduce the ten-dimension critical superstring theory to some four-dimensional effective field theory which possesses $\\mathcal{N} = 1$ supersymmetry. One of the properties of this scenario beyond one's expectation, is the tremendously abundant vacua, which mainly rise from the not-very-small Betti numbers $b_2$ and $b_3$ of $\\mathrm{CY}_3$, and the various possible fluxes wrapped on it; this set of vacua is always called a ``string landscape''~\\cite{Susskind:2003kw}. Different vacua in the landscape hold different fundamental parameters. It was argued that the number of consistent quasirealistic flux vacua may be greater than $10^{500}$~\\cite{Douglas:2003um}, and models has been constructed to solve the cosmological constant problem using this property~\\cite{Bousso:2000xa}. In this paper, we conjecture that as an analog, a holographic theory may also hold a landscape of vacua. We verify this hypothesis in two different ways, from top-down and from bottom-up. Along the first root, we discuss properties of the compact manifolds, and the restrictions of them from physical purposes. For the uncompactified dimensions to be AdS, the compact manifold should be Einstein; thus, most of our discussions are within the framework of Einstein manifolds~\\cite{Besse:1987}. After then, we consider the possibilities and phenomenologies of various AdS vacua, especially the properties of domain walls separate them. We also discuss the possibilities for a non-conformal boundary field theory to hold a landscape. Along the opposite root, we studied the consequences of our conjecture, if QCD (as a non-conformal boundary field theory) holds a landscape of vacua. In this case, different vacua of QCD should possess different fundamental parameters, such as the quark masses $m_q$, the running coupling constant $\\alpha_S$, or the CP violating phase $\\theta_\\mathrm{QCD}$. Another vacuum with parameters different from ours may be testable; and we estimate this possibility within several astrophysical environments. The organization of this paper is as follows. In Sec.~\\ref{sec:Einstein_manifold} we discuss several mathematical and physical issues that relate to the Einstein manifolds. We consider the symmetric conditions of the string worldsheet and the dual field theories, and the properties of wrapped branes and fluxes. We also consider the stability conditions of vacua topologies. In Sec.~\\ref{sec:holo_phenomenology}, we discuss the theoretical issues to approach a QCD landscape. We consider the possibilities to break CFT, the fundamental parameters a vacuum should determine, and the deduced parameters that may relate to applications/observations. In Sec.~\\ref{sec:astronomy}, we consider how a QCD landscape affects astrophysical observations. The applications are abundant, but the studies in our paper are only tentative. We summarize our results in Sec.~\\ref{sec:conclution}. Some mathematical supplements relatively independent to the main text are gathered in Appendix~\\ref{app:mathematics_addons}; and the validity of the orders of magnitudes estimations used in this paper, are reconsidered more carefully in Appendix~\\ref{app:order_estimations}. We gather these materials together, rather than write two separate papers from either the theoretical or the astrophysical aspects, because we think neither one alone is enough to make our conjecture reasonable; however, the two roots can in fact be read separately. We always denote the indices of the extended dimensions by $\\mu, \\nu$, of the compact dimensions by $m, n$, and of the entire target space by $M, N$. We set $\\hbar = k = c = 1$ for simplification throughout this paper. ", "conclusions": "} In this paper, we discussed the possibilities that whether QCD, or more generally some holographic $(3+1)$-dimensional gauge theories, can possess a ``landscape'' of its vacua. We first limited our discussions to some boundary CFTs, for which the compact manifolds are Einstein. An Einstein 5- or 6-manifold $\\mathcal{M}_q$, which needs not to be homogeneous, may have some not-very-small Betti numbers $b_m$. Therefore, if fluxes wrap different patterns on cycles of it, the moduli are also different; a landscape of the holographic vacua should rise for this reason. We examined several relevant issues for this conjecture from the theoretical viewpoint. The geometry of $\\mathrm{AdS}_5 \\times \\mathcal{M}_q$ should also possess some symmetric conditions, such as the worldsheet superconformal symmetry, or definite supersymmetries of the dual field theories. However, it seems that even the $\\mathrm{AdS}_5 \\times S^5$ violates the one-loop worldsheet conformal symmetry. The isometry group of $\\mathcal{M}_q$ decides the R-symmetry of the holographic theory, which is not imposed in our case. In addition, we considered the anomaly cancelation, the no-go theorem, the brane/flux configurations, and the vacua stabilities. We focused on the $(2+1)$-dimensional domain walls within $\\mathrm{Mink}_4$, and also the $(3+1)$-dimensional domain walls filling the radial $\\mathrm{AdS}_4$ inside $\\mathrm{AdS}_5$. We considered the possibilities to break the $\\mathrm{AdS}_5$ geometry, or equivalently the conformal symmetry of the boundary theory, and argued that the confinement-deconfinement phase transition may not affect the vacua configurations. After then, we applied our conjecture of the ``holographic landscape'' directly to QCD, for which the fundamental parameters such as $m_q$, $\\alpha_S$ or $\\theta_\\mathrm{QCD}$ should depend on the moduli of the compact dimensions. We studied the properties of the domain walls, such as their masses, thicknesses, and their dynamical properties; they may be both relativistic and non-relativistic. We discussed how this ``QCD landscape'' affects nuclear physics; they may alter the hadronic masses, the reaction rates, and the stabilities of the nuclei. In an opposite way, we studied how can a ``QCD landscape'' affects the astrophysical observations. As domain walls may be non-relativistic, another vacuum of QCD may be within the visible universe; if they are not, we can also consider the properties of the other multiverses of QCD, just as what is done in~\\cite{Bousso:2008bu,Bousso:2009gx}. We first considered whether the mass contribution of the domain walls exceeds the critical density of the universe. Most of the case, domain walls are really dangerous; however, it is not enough reasonable that they should be completely ruled out. We then discussed the properties of cosmic rays affecting by this landscape, include the GZK cutoff, the neutron lifetime, and an alternative explanation of the coincidence of the cosmic ray anisotropic spatial distribution with the supergalactic plane. The GZK cutoff depends on, but is not very sensitive to, a different QCD vacua; the variation of the neutron lifetime seems helpless to explain the observations. The alternative explanation may loose the constraints for the origins of the UHECRs; however, as too ambitious the scenario is, it needs to be studied more seriously. We also considered how the QCD landscape affects BBN and the star burning scenarios. As only the abundance of deuterium ${}^2\\mathrm{D}$ can be measured in far away part of the universe, only it can restrain the QCD landscape; the constraint is really loose. A different QCD vacuum can also affect the stellar evolution; for example, the time spent for the proton-proton chain reaction depends on the deuterium mass $m_\\mathrm{D}$ very sensitively. The Chandrasekhar mass limits of white dwarfs and neutron stars are also altered, but not very sensitive to the QCD landscape. In addition, whereas the ``strange quark matter'' may be the true matter ground state of our QCD~\\cite{Witten:1984rs}, charm or $(ud)$ QGP may be the ground states of other QCDs; thus, as an alternative, charm stars or $(ud)$ quark stars may exist in those QCDs. However, the mass-radius relationships and the surface properties of those objects seem really similar to our strange stars. Recently, Denef and Hartnoll discussed the landscape in condensed matter systems, which results a statistical distribution of critical superconducting temperatures~\\cite{Denef:2009tp}. In here, we compare briefly the original ``string landscape'', their ``atomic landscape'', and our ``QCD landscape''. First, whereas the relativistic quantum critical theories are conformal, our QCD landscape should break conformal symmetry by some mechanics. Second, the original Calabi-Yau compactification possesses a $\\mathcal{N} = 1$ supersymmetry, the ``atomic landscape'' in M-theory by the Sasaki-Einstein 7-manifolds holds a $\\mathcal{N} = 2$ supersymmetry; however, it seems that our ``QCD landscape'' is not restricted by supersymmetry conditions. Third, as the ``atomic landscape'' discussed in~\\cite{Denef:2009tp} is limited to the Freund-Rubin compactification, only the background electromagnetic four-flux contributes to the vacua field equations; thus their statistics of the landscape rises from the different topologies of the compact dimensions. However, the original Calabi-Yau vacua, or the QCD vacua discussed in this paper, are more abundant because of different flux configurations." }, "0911/0911.3396_arXiv.txt": { "abstract": "We discuss the universal relation between density and size of observed Dark Matter halos that was recently shown to hold on a wide range of scales, from dwarf galaxies to galaxy clusters. Predictions of $\\Lambda$CDM N-body simulations are consistent with this relation. We demonstrate that this property of $\\Lambda$CDM can be understood analytically in the secondary infall model. Qualitative understanding given by this model provides a new way to predict which deviations from $\\Lambda$CDM or large-scale modifications of gravity can affect universal behavior and, therefore, to constrain them observationally. ", "introduction": " ", "conclusions": "" }, "0911/0911.2997_arXiv.txt": { "abstract": "By means of a semiclassical analysis we show that the trace anomaly does not affect the cosmological constant. We calculate the evolution of the Hubble parameter in quasi de Sitter spacetime, where the Hubble parameter varies slowly in time, and in FLRW spacetimes. We show dynamically that a Universe consisting of matter with a constant equation of state, a cosmological constant and the quantum trace anomaly evolves either to the classical de Sitter attractor or to a quantum trace anomaly driven one. There is no dynamical effect that influences the effective value of the cosmological constant. ", "introduction": "T^{\\mu}_{\\phantom{\\mu}\\mu} = 0 \\,. \\end{equation} As an explicit example, consider a conformally coupled scalar field. In quantum field theory the stress-tensor is promoted to an operator. A careful renormalisation procedure renders its expectation value $\\langle \\hat{T}^{\\mu}_{\\phantom{\\mu}\\nu}\\rangle$ finite. However, inevitably, the renormalisation procedure results in general in a non-vanishing trace of the renormalised stress-energy tensor: \\begin{equation}\\label{intro2} \\langle \\hat{T}^{\\mu}_{\\phantom{\\mu}\\mu}\\rangle \\neq 0 \\,. \\end{equation} Classical conformal invariance cannot be preserved at the quantum level. In short, this is the trace anomaly. An important question that immediately arises is how this non-zero trace affects the evolution of the Universe through the Einstein field equations. In particular, one could wonder whether it influences the effective value of the cosmological constant. It has been argued (see \\cite{Antoniadis:2006wq} and references therein) that the trace anomaly could potentially provide us with a dynamical explanation of the cosmological constant problem. In brief, the line of reasoning is as follows. Firstly, a new conformal degree of freedom is introduced as $g_{\\mu\\nu}(x)=\\exp[2\\sigma(x)]\\overline{g}_{\\mu\\nu}(x)$. Furthermore, the trace anomaly stems from a non-local effective action that generates the conformal anomaly by variation with respect to the metric \\cite{Riegert:1984kt}. The authors of \\cite{Antoniadis:2006wq} then argue that the new conformal field should dynamically screen the cosmological constant, thus solving the cosmological constant problem. The proposal advocated in \\cite{Antoniadis:2006wq} is very appealing. Before studying the effect of this new conformal degree of freedom, we feel that firstly a proper complete analysis of the dynamics resulting from the effective action of the trace anomaly should be performed. This is what we pursue in this contribution. We argue in favour of a semiclassical approach to examining the connection between the cosmological constant and the trace anomaly: we take expectation values of inhomogeneous quantum fluctuations with respect to a certain state to study its effect on the background spacetime. Phase transitions aside, quantum fluctuations affect the background \\textit{homogenously}. Moreover, in accordance with the Cosmological Principle, inhomogeneous fluctuations of the metric tensor and in particular of the conformal part of the metric tensor are observed to be small at the largest scales, comparable to the Hubble radius (and expected to be small also in the early Universe). We do not consider a new, conformal degree of freedom. Hence, we do certainly not exclude any possible effect this (inhomogeneous) conformal degree of freedom might have on the cosmological constant. However, it is plausible that in order to address the link between the cosmological constant and the trace anomaly, a semiclassical analysis suffices. A final note is in order regarding an application of the conformal anomaly: trace anomaly induced inflation. In the absence of a cosmological constant, the trace anomaly could provide us with an effective cosmological constant \\cite{Starobinsky:1980te, Hawking:2000bb, Pelinson:2002ef,Shapiro:2002nz, Shapiro:2003gm, Pelinson:2003gn, Shapiro:2008sf}. If one includes a cosmological constant, the theory of anomaly induced inflation is plagued by instabilities, which we will also come to address. Improving on e.g. \\cite{Pelinson:2002ef, Brevik:2006nh}, we incorporate matter with a constant equation of state in the Einstein field equations. ", "conclusions": "We have studied the dynamics of the Hubble parameter both in quasi de Sitter and in FLRW spacetimes including matter, a cosmological term and the trace anomaly. We have seen that there is no dynamical effect that influences the effective value of the cosmological constant, i.e.: the classical de Sitter attractor. Based on our semiclassical analysis we thus conclude that the trace anomaly does \\textit{not} solve the cosmological constant problem. Ostrogradsky's theorem merits another remark. Clearly, including the higher derivative contributions in FLRW spacetimes modifies the dynamics of the Hubble parameter significantly: attractors, that were stable in the absence of higher derivatives, under certain conditions destabilise. We do not know which of the two approaches is correct. Discarding these higher derivatives and studying the trace anomaly in quasi de Sitter spacetime would seem plausible. Finally, one could wonder whether the quantum anomaly driven attractor is physical. The quantum attractor is of the order of the Planck mass $M_{\\mathrm{pl}}$, so only when the matter in the early Universe is sufficiently dense, $H \\simeq \\mathcal{O}( M_{\\mathrm{pl}})$. We then expect to evolve towards the quantum attractor. However, at these early times we also expect perturbative general relativity to break down. Hence, this attractor may be seriously affected by quantum fluctuations or it might even not be there." }, "0911/0911.0395_arXiv.txt": { "abstract": "{} {We report on the third phase of our study of the neutrino-cooled hyperaccreting torus around a black hole that powers the jet in Gamma Ray Bursts. We focus on the influence of the black hole spin on the properties of the torus.} {The structure of a stationary torus around the Kerr black hole is solved numerically. We take into account the detailed treatment of the microphysics in the nuclear equation of state that includes the neutrino trapping effect. } {We find, that in the case of rapidly rotating black holes, the thermal instability discussed in our previous work is enhanced and develops for much lower accretion rates. We also find the important role of the energy transfer from the rotating black hole to the torus, via the magnetic coupling. } {} ", "introduction": "Observed Gamma Ray bursts (GRBs) form at least two classes of events. Long GRBs ($T_{90} > $ 2 s) are the the result of a collapsar in the core of an exploding massive star, as supported by the detections of their afterglows. Shorter bursts ($T_{90} < $ 2 s) are likely to be due to the coalescence of a binary system containing two compact stars, as observationally supported by the localization of their afterglows at the outskirts of their host galaxies (see e.g. the review of Zhang 2007). Moreover, it is possible that a third class of events exists, namely a fraction of the shortest bursts ($T_{90} < $ 0.1 s), that can be produced via the evaporation of the primordial black holes (Cline et al. 2005). Both the two main classes of bursts involve a stage of a newly formed black hole surrounded by an accreting torus, responsible for the production of a collimated jet. This jet is a source of a highly beamed emission observed in the gamma ray band. The accreting torus is either fed by the material from the stellar envelope, which part falls back after the hypernova explosion, or consumes the material from the disrupted neutron star debris. The torus accretion proceeds with highly hyper-Eddington rates, on the order of a solar mass per second. Such hyper-accreting tori have been already discussed in a number of works (e.g. Popham et al. 1999; Di Matteo et al. 2002; Kohri et al. 2005; Chen \\& Beloborodov 2007; Janiuk et al. 2004; 2007; Lei et al. 2009). In the present paper we expand on our previous work and we further investigate the properties and evolution of such hot and dense accretion tori. At the extreme densities and temperatures, determined by the hyper-Eddington accretion rate, the torus is cooled mainly by neutrino emission produced primarily by electron and positron capture on nucleons ($\\beta$ reactions). The model we develop here was first described in Janiuk et al. (2004) for the case of a non-rotating black hole and the nuclear matter with a simplified equation of state. We further developed this model in Janiuk et al. (2007), considering the more elaborate equation of state to account for the microphysics of the accreting plasma. In the latter work, we solved for the disk structure and its time evolution by introducing the new EOS which includes photodisintegration of helium, the condition of $\\beta$ equilibrium, and neutrino opacities. We self-consistently calculated the chemical equilibrium in the gas consisting of helium, free protons, neutrons, and electron-positron pairs and compute the chemical potentials of the species, as well as the electron fraction throughout the disk. We found an important property of our solutions: for very large accretion rates, the torus becomes viscously and thermally unstable in a narrow range of radii near the central black hole. The instability results as the intrinsic property of the torus, when the helium is fully photodisintegrated in the neutrino opaque plasma. In the present work, we further expand our model to account for the black hole rotation. The rotating black hole should naturally be produced in the center of a collapsar or after the merger event, when the collapsing material possesses a large amount of angular momentum. As a consequence, the black hole spin is a plausible mechanism that helps to launch the jets emitted in a narrow cone along the rotation axis. As shown in Janiuk et al. (2008), especially for the long duration GRBs the BH spin is required to sustain the central engine activity and the jet production for a sufficiently long time. Because these long GRBs often exhibit very variable temporal structure (e.g. Beloborodov, Stern \\& Svensson 2000), it seems very important to study a possible instability mechanism that may account for the variability. We check here, for what black hole spin the instability may operate near the central black hole, depending on the accretion rate and viscosity in the disk. We also check what effects may be imposed on the torus, and its unstable strip, by the additional heating via the magnetic fields, threaded by the rotating black hole. The content of the article is as follows. First, we discuss the basic assumptions and present the main equations of the model. Second, we introduce the changes and corrections to the model that are the result of a black hole rotation with an arbitrary spin. Third, we introduce the energy extraction from the rotating black hole via the magnetic field, as an additional physical process that may operate in the gamma ray burst central engine. In Section \\ref{sec:results} we present the results of our calculations, and in Section \\ref{sec:diss} we discuss the results and conclude. ", "conclusions": "\\label{sec:diss} The importance of black hole spin in the accretion disk systems and the formation of jets, also in the presence of the magnetic fields, has been studied by many authors. It has been discussed e.g. in the review by Blandford (1999). In the standard accretion disk theory (Shakura \\& Sunyaev 1973) it is assumed that there is no coupling between the black hole and the disk. However, if the magnetic field is considered, such a coupling may be present. Contrary to the Blandford-Znajek process (in which the magnetic filed lines threading the black hole may close on a very remote load, far away from the black hole), the magnetic field lines coupled to the accretion disk are closed in the vicinity of the inner radius. In this process, the energy and angular momentum are extracted from the black hole and transferred to the disk, if the hole rotates faster than the disk (MacDonald \\& Thorne 1982; Li 2000). Such coupling may have much higher efficiency in extracting the black hole rotation energy than the Blandford-Znajek mechanism, as was first quantitatively estimated by Li (2002). Let us also remind here after Li (2002), that in this model the torque produced by the magnetic coupling with the black hole propagates only outwards, and no torque is imposed at the inner boundary of the disk. The non-zero torque at the inner boundary is possible if the disk is magnetically connected with the plasma in the transition region between the disk and the black hole (e.g. Krolik 1999), which we do not consider here. Another important assumption behind the present model is that about the quasi steady-state of the disk and magnetic field configuration. Under this assumption, the magnetic field lines are not tangled by the rotation and MHD instabilities of the Balbus \\& Hawley (1991) are not taken onto account. In our present work we hold this assumption, after Li (2002) and Wang et al. (2002; 2003a,b). The evolution of the coronal magnetic field threading the differentially rotating disk was studied however in the literature, e.g. by Lovelace et al. (2002) in the context of Poynting flux dominated jets in quasars, and in Uzdensky (2004; 2005) by numerically solving the Grad-Shafranov equation in the Schwarzschild and Kerr metric. In the context of Gamma Ray Bursts, the black hole spin interaction with the magnetized torus was studied by van Putten et al. (2004) These authors take into account the dissipation of energy in the torus not only via the thermal and neutrino emission, but also through the gravitational waves and winds. They estimate the lifetime of the central engine powered by rapidly spinning black hole, whose spin-down is governed by the surrounding magnetic field coupled to the inner torus. In the present model, we neglect the gravitational waves emission and we concentrate on the structure of the neutrino-cooled disk. We restrict ourselves to the steady-state solution, however we note that the lifetime of the long gamma ray burst central engine is defined by the period of the rapid spin of the black hole, which may decrease with time during the engine activity (see also Janiuk et al. 2008). In this article, we study the model of a stationary neutrino cooled disk around a rotating Kerr black hole. We confirmed the presence of the thermal-viscous instability in the region of the large neutrino opacity and efficient helium photodisintegration. The instability was first studied in Janiuk et al. (2007) in the time-dependent model for the Schwarzschild black hole. It was found to produce dramatic changes in the density and temperature profiles, as well as a rapid time variability of the disk emission. Here, we quantitatively described the role of the black hole spin and viscosity in the disk. We found that in the Kerr black hole disks the large spin parameter results in a larger size of an unstable strip, which appears even for a moderate accretion rate. The critical accretion rate is anticorrelated with the BH spin required for the onset of the instability. In the extreme BH case, the instability close to the inner edge may be effective even for $\\dot M=0.5 M_{\\odot}$ s$^{-1}$. On the other hand, the more viscous is the disk, the smaller in size is the unstable strip and the weaker should be the instability. The time dependent calculations were not performed in the present work, because they would require to compute time evolution of the BH spin and a moving radial grid, which is intended to be studied in the future work. However, we studied the influence of the black hole rotation on the disk structure considering our stationary model. This was done by means of the magnetic transfer of energy from the rotating black hole to the disk, and vice versa. The first important finding in this work is that thermal instability occurs for quite low accretion rates, on the order of a fraction of solar mass per second, provided the black hole rotates very fast. This can be relevant for the origin of the variable energy input to the GRB jets not only in case of the short GRBs, but also for long ones. The latter are presumably powered by the accretion of the fallback material from the massive star envelope onto the newly born black hole after the hypernova explosion (e.g. Mac Fadyen \\& Woosley 1999). In these models, the accretion rate is found to be substantially lower than in the merger of two neutron stars, and therefore the event can last for much longer time to power the GRB for several hundreds of seconds. On the other hand, the massive star that is a progenitor of a hypernova is a rotating star and therefore a natural consequence of the explosion should be the rapidly rotating black hole in the center. As was recently shown by Janiuk, Moderski \\& Proga (2008), the large black hole spin is a key ingredient for the occurrence of the longest duration GRBs, and is required for the GRB central engine to operate for a time of the order of hundreds of seconds. The present study leads us therefore to a general conclusion that the longest GRBs, that are powered first by the neutrino annihilation and later by the black hole spin and require $a>0.9$, should be very variable at the initial phase of the prompt emission. Later on, when the accretion rate drops, the variability may be suppressed and what is observed is a smooth tail in the gamma ray lightcurve. On the other hand, the period of the highly variable GRB emission may be somewhat extended if the black hole is being spun up by the accretion. The issue discussed in the literature was whether in such a case the spin equilibrium value can be achieved at $a=0.998$ (Thorne 1974), or a less value (Gammie, Shapiro \\& McKinney 2004). In our study, we also considered the heating of the innermost regions of the torus due to the magnetic coupling with the rotating black hole. We found that this process can have some influence on the thermal instability, because the size and position of the unstable region are somewhat changed. Also, the profiles of density and temperature in the unstable region are modified. The magnetic coupling enhances the change in density profile, however it weakens the change in temperature profile. The energy transfer to the disk from the rotating black hole does not suppress the thermal instability. This finding also supports our conclusion that the variable and long duration GRBs are powered by the rapidly spinning black holes." }, "0911/0911.2673_arXiv.txt": { "abstract": "UV radiation from massive stars is thought to be the dominant heating mechanism of the nuclear ISM in the late stages of evolution of starburst galaxies, creating large photodissociation regions (PDRs) and driving a very specific chemistry. We report the first detection of PDR molecular tracers, namely HOC$^+$, and CO$^+$, and confirm the detection of the also PDR tracer HCO towards the starburst galaxy NGC\\,253, claimed to be mainly dominated by shock heating and in an earlier stage of evolution than M\\,82, the prototypical extragalactic PDR. Our CO$^+$ detection suffers from significant blending to a group of transitions of $^{13}$CH$_3$OH, tentatively detected for the first time in the extragalactic interstellar medium. These species are efficiently formed in the highly UV irradiated outer layers of molecular clouds, as observed in the late stage nuclear starburst in M\\,82. The molecular abundance ratios we derive for these molecules are very similar to those found in M\\,82. This strongly supports the idea that these molecules are tracing the PDR component associated with the starburst in the nuclear region of NGC\\,253. The presence of large abundances of PDR molecules in the ISM of NGC\\,253, which is dominated by shock chemistry, clearly illustrates the potential of chemical complexity studies to establish the evolutionary state of starbursts in galaxies. A comparison with the predictions of chemical models for PDRs shows that the observed molecular ratios are tracing the outer layers of UV illuminated clouds up to two magnitudes of visual extinction. We combine the column densities of PDR tracers reported in this paper with those of easily photodissociated species, such as HNCO, to derive the fraction of material in the well shielded core relative to the UV pervaded envelopes. Chemical models, which include grain formation and photodissociation of HNCO, support the scenario of a photo-dominated chemistry as an explanation to the abundances of the observed species. From this comparison we conclude that the molecular clouds in NGC\\,253 are more massive and with larger column densities than those in M\\,82, as expected from the evolutionary stage of the starbursts in both galaxies. ", "introduction": "Intense UV radiation from massive stars is one of the main mechanisms responsible for the heating of the interestelar medium in the nuclear region of starburst galaxies. This mechanism is particularly important in the latest stages of starburst (SB) galaxies where the newly formed massive star clusters are responsible for creating large photodissociation regions (PDRs). This is the case for the prototypical SB galaxy M\\,82, where the large observed abundances of molecular species such as HCO, HOC$^+$, CO$^+$, and H$_3$O$^+$ are claimed to be probes of the high ionization rates in large PDRs formed as a consequence of its extended evolved nuclear starburst \\citep{Burillo02,Fuente06,VdTak08}. Observational evidences point to a significant enhancement in the abundance of HOC$^+$ in regions with large ionization fractions. The abundance ratio [HCO$^+$]/[HOC$^+$]$=270$ is found in the prototypical Galactic PDRs of the Orion Bar \\citep{Apponi99}. Similar or even lower abundance ratios are observed in the PDRs NGC\\,7023 \\citep[50-120,][]{Fuente03}, Sgr\\,B2(OH) and NGC\\,2024 \\citep[360-900,][]{Ziurys95,Apponi97}, and the Horsehead \\citep[75-200][]{Goico09}, as well as in diffuse clouds \\citep[70-120,][]{Liszt04}. This is in contrast with the much larger ratios of $\\gg 1000$ found in dense molecular clouds well shielded from the UV radiation. However, these low HCO$^+$/HOC$^+$ ratios are not found in other galactic PDRs. Large values of this ratio of $\\gtrsim2000$ are found in the PDRs M17-SW, S140, and NGC2023 \\citep{Apponi99,Savage04}. The HCO molecule has also been observed to be a particularly good tracer of the PDR interfaces. Low ratios of [HCO$^+$]/[HCO]$\\sim2.5-30$ are found in prototypical Galactic PDRs \\citep{Schene88,Schilke01}. The large HCO abundance ($>10^{-10}$) altogether with the low ratio [HCO$^+$]/[HCO]$\\sim1$ in the Horsehead PDR is claimed to be a diagnostic for an ongoing FUV-dominated photochemistry \\citep{Gerin09}. CO$^+$ is also claimed to be particularly prominent in the chemical modeling of PDRs and high abundances of this molecule appear to be correlated to similar enhancements of HOC$^+$ \\citep{Sternberg95,Savage04}. [CO$^+$]/[HOC$^+$] ratios in the range of 1-10 are observed in a number of PDRs \\citep{Savage04}, but only of $\\gtrsim 0.1.$ towards the Horsehead PDR \\citep{Goico09}. As mentioned above, this set of PDR probes has been extensively studied towards M\\,82. However, no such complete studies have been carried out towards other prototypical galaxies, but for the detection of HCO and HOC$^+$ towards NGC\\,1068 \\citep{Usero04} and H$_3$O$^+$ in Arp\\,220 \\citep{VdTak08}. M\\,82 and NGC\\,253 are the brightest prototypes of nearby SB galaxies, at a similar distance and showing very similar IR luminosities and star formation rates of about $\\sim3\\,M_\\odot \\rm yr^{-1}$ \\citep{Ott05,Minh07}.\\ However, both galaxies show very different chemical composition. The chemistry and to a large extend the heating in the central region of NGC\\,253 is believed to be dominated by large scale low velocity shocks \\citep{Martin06b}. The similar chemical composition found in the nuclear region of NGC\\,253 to that in Galactic star forming molecular complexes points to an earlier evolutionary stage of the starburst in this galaxy than that in M\\,82 \\citep{Martin03,Martin05,Martin06b}. Furthermore, our recent observations of the PDR component as traced by the easily photodissociated HNCO molecule towards a sample of galaxies \\citep{Martin08} showed the non-detection of HNCO in M\\,82, at a very low abundance limit. This low HNCO abundance supports the scenario that the PDR chemistry dominates the molecular composition of the ISM in this galaxy. However, from the HNCO measured abundance in NGC\\,253, it would be placed in an intermediate stage of evolution where photodissociation should be starting to play a significant role in driving a UV-dominated chemistry which has not been yet identified towards this galaxy. The presence of a significant PDR component in NGC\\,253 claimed from the HNCO abundance is also inferred from the similar intensity of the atomic fine structure line intensities from PDR tracers like CII and OI \\citep{Carral94,Lord96} observed in both M\\,82 and NGC\\,253. In this paper we present the first detection of PDR molecular tracers HOC$^+$ and CO$^+$, and confirm the detection HCO \\citep[tentatively detected by][]{Sage95} in the central region of NGC\\,253 which allows the evaluation of the influence of the photodissociation radiation in the nuclear ISM of this SB galaxy. The results presented here support the scenario of the presence of a significant PDR component and clearly show the potential of molecular complexity in estimating the contribution of the different heating mechanisms of the ISM in the nuclei of galaxies. ", "conclusions": "The comparison of model predictions with the observations presented show that the abundance of the species observed in this work towards NGC\\,253, namely HCO$^+$, CO$^+$, and HCO, are most efficiently formed in the outer region of the molecular clouds where the gas is highly irradiated by the incident UV photons from massive stars. The high molecular abundances derived for these species in NGC\\,253 suggest that the PDR component in this galaxy is similar to that found in M\\,82, claimed to be the prototype of extragalactic PDR. The abundance ratios found for this limited sample of galaxies are of the same order as those observed towards galactic PDRs, which stress the importance of photo-dominated chemistry in galaxy nuclei. Large amounts of molecular material are affected by photodissociation not only in NGC\\,253, but also towards the star forming regions around the Seyfert 2 nuclei in NGC\\,4945 and NGC\\,1068. This is consistent with the HNCO/CS ratio in these galaxies which suggest that a fraction of HNCO has been photodissociated in PDRs. The combination of the observations of HCO, HOC$^+$ and CO$^+$ with that of HNCO seems to confirm that their abundances reflect the evolutionary stage of the starbursts in these galaxies. Although photodissociation is the most likely scenario for the enhancement of the observed reactive ion in starburst environments, X-ray dominated chemistry has been claimed to be responsible for the high abundances observed around AGNs in circunnuclear disk of NGC\\,1068 \\citep[HOC$^+$][]{Usero04} and towards the ultra luminous infrared galaxy Arp\\,220 \\citep[$\\rm H_3O^+$][]{VdTak08}. Therefore, M\\,82 is still outstanding not only as a PDR dominated galaxy, but by the underabundance of complex molecules such as CH$_3$OH, HNCO or SiO \\citep{Mauers93,Martin06a,Martin06b}, evidence for the lack of large amounts of dense molecular material which would potentially fuel its nuclear starburst as compared to other starburst galaxies like NGC\\,253 \\citep{Martin09}. Our data in combination with the HNCO abundances \\citep{Martin09} indicate that the molecular clouds in M\\,82 are different from those in NGC253. Although having a similar overall PDR component, the clouds in NGC\\,253 have to be more massive and have larger column densities those in M\\,82." }, "0911/0911.0506_arXiv.txt": { "abstract": "Using archival data from the Hubble Space Telescope, we study the quantitative morphological evolution of spectroscopically confirmed bright galaxies in the core regions of nine clusters ranging in redshift from $z = 0.31$ to $z = 0.84$. We use morphological parameters derived from two dimensional bulge-disk decomposition to study the evolution. We find an increase in the mean bulge-to-total luminosity ratio $B/T$ as the Universe evolves. We also find a corresponding increase in the fraction of early type galaxies and in the mean S\\'ersic index. We discuss these results and their implications to physical mechanisms for evolution of galaxy morphology. ", "introduction": "Clusters serve as laboratories for investigating the dependence of galaxy morphology on the density of the environment. \\citet{dre80} found that the fraction of early type galaxies increases with local density of galaxies. He also found that around 80\\% of galaxies in nearby clusters are of the early type. In recent years, deep, high resolution imaging by the \\textit{Hubble Space Telescope (HST)} has helped to extend the study of galaxy morphology as a function of the environment to $z \\sim 1$. The observed fraction of different morphological types can be explained in terms of both 'nature' and 'nurture' scenarios. The former argues that a galaxy's morphological type is determined by initial conditions at formation \\citep[e.g. ][]{egg62}, while the latter depends on the influence of the environment and of secular evolution for determining the final morphological type \\citep[e.g. ][]{too77}. Numerical simulations play an important role in investigating the relative importance of the two scenarios under different physical conditions. Following Dressler's pioneering work, many observational studies have been done to measure and understand the morphology density relation (MDR), and the dependence of morphological type on distance of the galaxy from the cluster centre \\citep{dre97,got03,smi05,pos05,hol07,tre03,cap07}. \\citet{smi05} found that the early type fraction is constant in low density environments over the last 10 Gyr, but there is significant evolution in this fraction in higher density regions. According to \\citet{smi05}, this suggests that most of the ellipticals in clusters formed at high redshift, and the increase in the fraction of early type galaxies is because of the physical processes in dense regions which transform disk galaxies with ongoing star formation to early types. Dissipationless merging of cluster galaxies may also be responsible for this increase. There is increasing evidence to show that massive ellipticals formed by dissipationless (dry) merger of two or more systems. Such ellipicals must have formed at later times than their low luminosity counterparts \\citep{luc06}. \\citet{dok05} analysed tidal debris of elliptical galaxies and concluded that $\\sim70\\%$ of the bulge dominated galaxies have experienced a merger. The analysis of nearby bulge dominated galaxies has shown that the gas to stellar mass ratio is very small and these mergers are mostly 'dry'. Using a semi-analytic model for galaxy formation, \\citet{naa06} found that both the photometric and kinematic properties of massive elliptical galaxies are in agreement with the scenario where massive elliptical galaxies are produced by mergers of lower mass ellipticals. They suggested that the merger of two spiral galaxies alone cannot reproduce the observed properties, and that the large remnant mass ($> 6\\times 10^{11} M_\\odot$) implies that they must have undergone elliptical - elliptical mergers. They found that this process is independent of the environment and redshift, which means that dry mergers can occur at low redshift as well. By analysing spirals from cluster and group environments at intermediate redshift ($z\\sim 0.5$), \\citet{mor07} found that the Tully-Fisher relation shows larger scatter for cluster spirals than for those in the field. They also found that the central surface mass density of spirals in clusters is small beyond the cluster virial radius, and argued that these observations provide evidence for merger/harrasment. Mergers are not common in clusters, as the velocity dispersion of virialized clusters is large. So if ellipticals in clusters have formed by mergers, that most likely happened during the early stage of cluster collapse \\citep{roo82}. There is some observational evidence to support this idea; \\citet{dok99} showed that there is a large fraction of ongoing mergers in a cluster at $z = 0.83$. The observation of the unvirialized cluster $\\textrm{RX J}0848+4453$ at $z=1.27$ revealed many ongoing dissipationless mergers of galaxies \\citep{dok01}. These observations go against the view of monolithic collapse where all the ellipticals formed at the same time, at very high redshift. In this letter we report on the evolution of galaxies in the core region of clusters, measured using bulge-disk decomposition. We show that in the case of brightest cluster galaxies, the fraction of galaxies with $B/T>0.4$ and $n>2.5$ evolved significantly over the redshift range 0.31 to 0.83. Throughout the paper we use the standard concordance cosmology with $\\Omega_\\Lambda = 0.73$, $\\Omega_m = 0.27$ and $H_0 = 71$ km s$^{-1}$ Mpc$^{-1}$. ", "conclusions": "" }, "0911/0911.3922_arXiv.txt": { "abstract": "We present a Nobeyama 45\\,m Radio Telescope map and Australia Telescope Compact Array pointed observations of \\dia\\, 1-0 emission towards the clustered, low mass star forming Oph B Core within the Ophiuchus molecular cloud. We compare these data with previously published results of high resolution \\amm\\, (1,1) and (2,2) observations in Oph B. We use 3D {\\sc clumpfind} to identify emission features in the single-dish \\dia\\, map, and find that the \\dia\\, `clumps' match well similar features previously identified in \\amm\\, (1,1) emission, but are frequently offset to clumps identified at similar resolution in 850\\,\\micron\\, continuum emission. Wide line widths in the Oph B2 sub-Core indicate non-thermal motions dominate the Core kinematics, and remain transonic at densities $n \\sim 3 \\times 10^5$\\,\\cc\\, with large scatter and no trend with $N(\\mbox{H$_2$})$. In contrast, non-thermal motions in Oph B1 and B3 are subsonic with little variation, but also show no trend with H$_2$ column density. Over all Oph B, non-thermal \\dia\\, line widths are substantially narrower than those traced by \\amm, making it unlikely \\amm\\, and \\dia\\, trace the same material, but the $v_{\\mbox{\\tiny{LSR}}}$ of both species agree well. We find evidence for accretion in Oph B1 from the surrounding ambient gas. The \\amm\\,/\\,\\dia\\, abundance ratio is larger towards starless Oph B1 than towards protostellar Oph B2, similar to recent observational results in other star-forming regions. The interferometer observations reveal small-scale structure in \\dia\\, 1-0 emission, which are again offset from continuum emission. No interferometric \\dia\\, emission peaks were found to be coincident with continuum clumps. In particular, the $\\sim 1$\\,M$_\\odot$ B2-MM8 clump is associated with a \\dia\\, emission minimum and surrounded by a broken ring-like \\dia\\, emission structure, suggestive of \\dia\\, depletion. We find a strong general trend of decreasing \\dia\\, abundance with increasing $N(\\mbox{H$_2$})$ in Oph B which matches that found for \\amm. ", "introduction": "Stars form out of the gravitational collapse of centrally condensed cores of dense molecular gas. Recent years have seen leaps forward in our understanding of the structure and evolution of isolated, star forming cores. Most star formation, however, occurs in clustered environments \\citep{lada03}. These regions are more complex, with complicated observed geometries, and contain cores which tend to have higher densities and more compact sizes than those found in isolation \\citep{wardthompson07}. It is likely that due to these differences the evolution of filaments and cores in clustered regions proceeds differently than in the isolated cases. Characterizing the physical and chemical structures of these more complicated regions are thus the first steps towards a better understanding of the process of clustered star formation. It is now clear that molecular cores become extremely chemically differentiated, as many molecules commonly used for tracing molecular gas, such as CO, become severely depleted in the innermost core regions through adsorption onto dust grains \\citep[see, e.g.,][for a review]{difran07}. In a recent paper \\citep[hereafter \\one]{friesen09}, we studied the dense gas in several cluster forming Cores\\footnotemark\\footnotetext{\\edits{In this paper, we describe Oph A, B, C, etc., as `Cores' since this is how these features were named in DCO$^+$ observations of the L1688 region by \\citet{loren90}. Since then, higher-resolution data such as those described in this paper have revealed substructure in these features that could be themselves precursors to stars, i.e., cores. To avoid confusion, we refer the larger features in Oph as `Cores' and Core substructure identified by {\\sc clumpfind} as `clumps'.}} in the Ophiuchus molecular cloud (Oph B, C and F) through high resolution observations of \\amm\\, (1,1) and (2,2) emission made at the Green Bank Telescope (GBT), the Australia Telescope Compact Array (ATCA) and the Very Large Array (VLA). The Ophiuchus molecular cloud \\citep[$\\sim 120$\\,pc distant; ][]{loinard08,lombardi08,knude98} is our closest example of ongoing, clustered star formation. We found that the Ophiuchus Cores presented physical characteristics similar to those found in studies of both isolated and clustered environments. In Oph C, for example, gas motions are subsonic (mean $\\sigma_v = 0.16$\\,\\kms), and decrease in magnitude along with the kinetic gas temperature ($T_K$) towards the thermal dust continuum emission peak in a manner reminiscent of findings in the isolated cores L1544 \\citep{crapsi07}; etc. In contrast to Oph C, \\amm\\, line widths in Oph B are dominated by transonic non-thermal motions and the gas temperatures are warmer than typically found in isolated regions ($\\langle T_K \\rangle = 15$\\,K) and nearly constant across the Core. No contrast in any parameters [such as $T_K$, ratio of non-thermal to thermal line width $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$, or \\amm\\, column density $N(\\mbox{\\amm})$] was found between the two main components of Oph B, labelled B1 and B2, despite the presence of several embedded Class I protostars in B2 while B1 appears starless. Additionally, on small scales ($\\sim 15$\\arcsec, or 1800\\,AU at 120\\,pc), significant offsets were found between the peaks of integrated \\amm\\, (1,1) intensity and those of submillimeter continuum emission, as well as between individual `clumps' identified through 3D {\\sc clumpfind} \\citep{williams94} in the \\amm\\, data cube and those found through 2D {\\sc clumpfind} in 850\\,\\micron\\, continuum emission. Finally, evidence was found for a decreasing fractional \\amm\\, abundance with increasing H$_2$ column density (as traced by 850\\,\\micron\\, continuum emission), suggesting that \\amm\\, (1,1) may not be tracing the densest gas in the Oph Cores. The diazenylium ion, \\dia, has been shown to be a preferential tracer of quiescent dense gas in molecular clouds \\citep{womack92,caselli02,tafalla02,tafalla04,difran04}. The 1-0 rotational transition has a critical density $n_{cr} \\sim 2 \\times 10^5$\\,\\cc, $\\sim 10-100 \\times$ greater than that of \\amm\\, (1,1), and is therefore possibly better suited to probing gas properties at the high densities expected in clustered star forming Cores. Based on both observations and chemical models of starless cores, \\dia\\, also appears resilient to depletion at the high densities ($n \\gtrsim 10^5$\\,\\cc) and cold temperatures ($T \\sim 10$\\,K) characteristic of the later stages of prestellar core evolution \\citep[see, e.g.,][]{tafalla04,bergin02}, and is thus expected to be an excellent tracer of the physical conditions of dense cores and clumps. Recently, \\citet{andre07} observed the Oph A, B, C, E and F Cores in \\dia\\, 1-0 emission at 26\\arcsec\\, angular resolution using the IRAM 30\\,m telescope. They found that the gas traced by \\dia\\, emission near clumps identified in millimeter continuum emission \\citep{motte98} is characterized, on average, with relatively narrow line widths showing small non-thermal velocity dispersions, in contrast to the large non-thermal motions traced by \\amm\\, emission in \\one. The authors derived virial masses from the \\dia\\, emission, and found they generally agreed within a factor $\\sim 2$ with the masses derived from dust emission, suggesting that the clumps are gravitationally bound and prestellar. The relative motions of the clumps are also small and subvirial, with a crossing time larger than the expected lifetime of the objects, such that clump-clump interactions are not expected to impact their evolution. These results suggest that a dynamic picture of clump evolution involving competitive accretion at the prestellar stage does not accurately describe the star formation process in central Ophiuchus. In this work, we discuss the results of higher resolution (18\\arcsec, or $\\sim 2200$\\,AU at 120\\,pc) observations of \\dia\\, 1-0 in the Ophiuchus B Core that reveal the distribution, kinematics and abundance pattern of the Core and associated embedded clumps on smaller physical scales. We additionally examine \\dia\\, 1-0 structure at even higher resolution ($8\\arcsec \\times 5\\arcsec$), through observations made with the ATCA, towards five locations in Oph B2 where small-scale structure was found in \\amm\\, (\\one) and in previously unpublished Berkeley-Illinois-Maryland Association (BIMA) \\dia\\, 1-0 observations. We compare these data with our recently published analysis of \\amm\\, emission in Oph B, in particular focusing on the kinematics and relative abundances of the two species to compare with physical and chemical models of dense core formation and collapse. We discuss the observations and calibration and \\S2. In \\S3, we present the data, and discuss the results of the hyperfine line structure fitting procedure and derivations of the column density, $N(\\mbox{\\dia})$, and fractional \\dia\\, abundance, $X(\\mbox{\\dia})$, in \\S4. In \\S5, we discuss general trends in the data, and compare these results with those found in \\one\\, and with studies of dense cores in isolated environments. We summarize our findings in \\S6. ", "conclusions": "\\subsection{General trends} \\label{sec:trends} We show in Figure \\ref{fig:n2h_trends} the distribution of the ratio of the non-thermal dispersion to the sound speed, $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$, $n_{ex}$, $N(\\mbox{\\dia})$ and $X(\\mbox{\\dia})$ with $N(\\mbox{H$_2$})$ in Oph B, omitting pixels where the 850\\,\\micron\\, continuum flux $S_\\nu \\leq 0.1$\\,Jy\\,beam$^{-1}$. Data points represent values for 18\\arcsec\\, pixels, i.e., approximately the Nobeyama beam FWHM. We show individually pixels in Oph B1 and B2, and also show peak values for \\dia\\, clumps, identified in \\S\\ref{sec:clumps}, and 850\\,\\micron\\, continuum clumps \\citep{jorgensen08}. Note that no Oph B3 values are plotted due to the lack of continuum emission at the B3 location. The Figure shows that the \\dia\\, clumps do not reside at the highest H$_2$ column densities, but scatter over the range of $N(\\mbox{H$_2$})$ calculated above our submillimeter flux threshold. The submillimeter clumps tend to be found at higher $N(\\mbox{H$_2$})$ values than the \\dia\\, clumps, on average, but also show a spread in peak $N(\\mbox{H$_2$})$. In \\one, we used the ratio of \\amm\\, (1,1) and (2,2) emission lines in Oph B to determine the kinetic temperature $T_K$ in each pixel, which we can use to calculate $\\sigma_{\\mbox{\\tiny{NT}}}$ and $c_s$ for \\dia\\, emission in Oph B. Given $T_K$, $\\sigma_{\\mbox{\\tiny{NT}}} = \\sqrt{\\sigma_{obs}^2 - k_B T_K\\,/\\,(\\mu_{mol} m_{\\mbox{\\tiny{H}}})}$, where $k_B$ is the Boltzmann constant, $m_{\\mbox{\\tiny{H}}}$ is the mass of the hydrogen atom, $\\mu_{mol}$ is the molecular weight in atomic units ($\\mu_{\\mbox{\\tiny{\\dia}}} = 29.02$) and $\\sigma_{obs} = \\Delta v\\,/\\,(2\\sqrt{2\\ln2})$. We list values and propagated uncertainties for individual \\dia\\, clumps in Table \\ref{tab:peak_dat}, and give the mean, rms, minimum and maximum $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ in Table \\ref{tab:col} for each of Oph B1, B2 and B3. We find a mean $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s = 1.02$ for \\dia\\, 1-0 emission across Oph B with an rms variation of 0.49, indicating that the non-thermal motions are approximately equal, on average, to the sound speed. When looked at separately, we find a difference in the $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ ratio between Oph B1 and B2. In Oph B1, the mean $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s = 0.71$, with an rms variation of 0.16. In Oph B2, the mean $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s = 1.26$ with an rms variation of 0.4. Oph B3 is dominated by nearly thermal line widths, with the mean $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s = 0.39$ and an rms variation of only 0.13. We find little variation in $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ with $N(\\mbox{H$_2$})$ in Oph B, as shown in Figure \\ref{fig:n2h_trends}a. The difference in the mean $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ ratio between B1 and B2 is clear, as is the significantly larger scatter of non-thermal line widths in B2 than in B1. The scatter in $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ extends to the highest H$_2$ column densities. In B1, all non-thermal line widths $\\sigma_{\\mbox{\\tiny{NT}}} \\leqq c_s$, shown by the dashed line. Similar results are found for most \\dia\\, clumps, and approximately half the continuum clumps. At 18\\arcsec\\, resolution, the observed non-thermal line widths do not extend to arbitrarily low values, as the data show a cutoff in the $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ ratio at $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s \\sim 0.5$ for both B1 and B2, below which no data points are found (but note that in \\S\\ref{sec:analysis-atca}, we found narrow line widths at high resolution). For $T_K = 15$\\,K, the mean temperature found in Oph B, the $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ cutoff corresponds to $\\sigma_{\\mbox{\\tiny{NT}}} \\sim 0.12$\\,\\kms. Note that the velocity resolution of the \\dia\\, data, $\\Delta v_{res} = 0.05$\\,\\kms, is significantly less than the observed lower $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ limit. In \\one, we discussed the possible origins of the large non-thermal motions in Oph B, and based on timescale arguments ruled out primordial motions (i.e., motions inherited from the parent cloud). It is interesting that most of the gas traced by \\dia\\, 1-0 in Oph B2 remains dominated by transonic non-thermal motions, while the line widths in Oph B1 reveal the Core is significantly more quiescent at the $\\sim 10^5$\\,\\cc\\, densities traced by \\dia\\, 1-0 emission. The most obvious source of this difference is the presence of embedded protostars in Oph B2 while Oph B1 is starless. \\citet{kamazaki03} found an outflow in CO 3-2 emission in Oph B2 centred near Elias 33 and Elias 32 and oriented approximately east-west, but were unable to pinpoint which protostar was the driving source. In upcoming results from the JCMT Gould Belt Legacy survey \\citep{gbls}, which mapped Oph B in CO 3-2, $^{13}$CO 3-2 and C$^{18}$O 3-2 at 14\\arcsec\\, resolution, this highly clumped outflow is shown to extend over more than 10\\arcmin\\, (i.e., 0.4\\,pc at 120\\,pc, White {\\it et al.} 2009, in preparation). The outflow axis is also aligned such that the outflow does not appear to be impacting Oph B1 significantly, potentially explaining the difference in non-thermal line widths between the two sub-Cores\\footnotemark\\footnotetext{JCMT Spring 2009 newsletter, http://www.jach.hawaii.edu/JCMT/publications/newsletter/n30/jcmt-n30.pdf}. \\dia\\, is not generally thought to be a tracer of protostellar outflows, since it is expected to be destroyed quickly through reactions with CO, which evaporates from dust grains in the higher temperature gas near the driving protostar. On small scales in B2, however, some variations in $v_{\\mbox{\\tiny{LSR}}}$ and $\\Delta v$ appear correlated with the presence of nearby Class I protostars (see \\S\\ref{sec:vlsr} and Figure \\ref{fig:b-mm8}). \\citet{chen08} suggest \\dia\\, can be entrained in protostellar jets before a molecular outflow releases CO from grain surfaces which then can destroy \\dia. In a small \\dia\\, 1-0 survey of the Serpens NW cluster, \\citet{williams00} found that the largest $\\sigma_{\\mbox{\\tiny{NT}}}$ values ($\\sigma_{\\mbox{\\tiny{NT}}} > 0.6$\\,\\kms, greater than seen here in Oph B2) occurred in cores containing protostellar sources which power strong outflows. If caused by the substantial protostellar outflow, the large line widths on a global scale in Oph B2 suggest that protostellar outflows are able to inject additional turbulence into the {\\it high density} gas in cluster-forming Cores, thereby increasing the mass at which clumps become gravitationally unstable and altering the consequent fragmentation and evolution of existing embedded clumps. \\citet{difran01} did not see this effect in the \\dia\\, 1-0 observations of the circumstellar dense gas associated with the protostellar, outflow-driving source NGC 1333 IRAS4B, suggesting that the ability of the outflow to inject turbulent energy may be determined by additional factors, such as the collimation or strength of the outflow. A comparison of the outflow properties in these regions would be useful to probe further the impact of outflows on cluster forming dense gas. Alternatively, the larger line widths in B2 may be due to global infall motions. In \\one, we determined a virial mass $M_{vir} \\sim 8$\\,M$_\\odot$ for Oph B2 based on \\amm\\, line widths, which is a factor $\\sim 5$ less than Core mass estimates based on thermal dust continuum emission (with a factor $\\sim 2$ uncertainty). In contrast, $M/M_{vir} \\sim 1$ in Oph B1. Some evidence for infall has been observed in self-absorbed molecular line tracers in Oph B2 \\citep{andre07,gurney08}, but it is difficult to determine unambiguously given the complex outflow motions also present. Additionally, the consistent line widths found in \\amm\\, gas for both B1 and B2 (discussed further in \\S\\ref{sec:sig} in comparison with the \\dia\\, results) suggest a common source for non-thermal motions at gas densities $n \\sim 10^{3-4}$\\,\\cc, which then impacts differently the gas at higher densities in the two sub-Cores. Figure \\ref{fig:n2h_trends}b shows the volume density $n_{\\mbox{\\tiny{ex}}}$ as a function of $N(\\mbox{H$_2$})$. Nearly all pixels have $n_{\\mbox{\\tiny{ex}}} \\gtrsim 10^5$\\,\\cc, and we find slightly greater $n_{\\mbox{\\tiny{ex}}}$ at higher column densities. The \\dia\\, and continuum clumps scatter over the full range of density values found in Oph B. We find a slight trend of increasing \\dia\\, column density, $N(\\mbox{\\dia})$, with $N(\\mbox{H$_2$})$, shown in Figure \\ref{fig:n2h_trends}c, although the mean $N(\\mbox{\\dia})$ changes by less than a factor of $\\sim 2$ over the order of magnitude range in $N(\\mbox{H$_2$})$ found in Oph B. Several pixels in Oph B1, and one \\dia\\, clump, have significantly greater $N(\\mbox{\\dia})$ for their $N(\\mbox{H$_2$})$ values compared with the rest of the data. The small gradient in $N(\\mbox{\\dia})$ with $N(\\mbox{H$_2$})$ leads to a trend of {\\it decreasing} fractional \\dia\\, abundance with increasing H$_2$ column density, shown in Figure \\ref{fig:n2h_trends}d. We find \\dia\\, abundances decrease by an order of magnitude between the limiting $N(\\mbox{H$_2$})$ threshold and the peak $N(\\mbox{H$_2$})$ values. We fit a linear relationship to $\\log X(\\mbox{\\dia})$ versus $\\log N(\\mbox{H$_2$})$ for Oph B1 and B2 separately. In B1 alone, where we have relatively few data points, we find $\\log X(\\mbox{\\dia})$ is consistent within the uncertainties with being constant with $N(\\mbox{H$_2$})$, but in conjunction with the Oph B2 data, we find \\begin{equation} \\label{eqn:xvsh} \\log X(\\mbox{\\dia}) = (7.1 \\pm 0.9) - (0.74 \\pm 0.04) \\times \\log N(\\mbox{H$_2$}) \\end{equation} \\noindent using the same H$_2$ column density limits as in Figure \\ref{fig:n2h_trends}. \\subsection{Small-Scale Features} We next discuss the physical properties of the small-scale structure present in Oph B, including the \\dia\\, clumps, continuum clumps, and embedded protostars, and describe in detail the compact, thermal object B2-N6. \\subsubsection{Comparison of \\dia\\, clumps, continuum clumps and protostars} \\label{sec:compare_clumps} We list in Table \\ref{tab:prot_dat} the physical parameters derived from \\dia\\, 1-0 emission towards 850\\,\\micron\\, continuum clump locations \\citep{jorgensen08} and embedded Class I protostars \\citep{enoch09}. Columns are the same as in Table \\ref{tab:peak_dat}. In Table \\ref{tab:mean_dat}, we list the mean values of each physical parameter towards \\dia\\, clumps, continuum clumps and protostars. Note that we were only able to solve for all parameters listed towards a single protostar, and discuss below only those mean parameters where we obtained values from three or more objects. We find no difference in the mean $v_{LSR}$ between \\dia\\, clumps, continuum clumps and protostars. There is a clear reduction in $\\Delta v$ towards \\dia\\, clumps relative to continuum clumps (mean $\\Delta v = 0.49$\\,\\kms and 0.68\\,\\kms, respectively), but excluding one continuum clump (162712-24290) reduces the mean $\\Delta v$ to 0.58\\,\\kms. Protostars are associated with wider mean $\\Delta v$, but the larger mean is also driven by large $\\Delta v$ towards a single object. The differences between \\dia\\, and continuum clumps are marginally significant given the variance of $\\Delta v$ across the entire Core is only 0.08\\,\\kms\\, in Oph B1 and 0.21\\,\\kms\\, in Oph B2. We find similar \\dia\\, line opacities and excitation temperatures towards continuum and \\dia\\, clumps. Continuum clumps are associated with larger non-thermal motions, with a mean $\\sigma_{NT} / c_s = 1.24$ which is 1.5 times that associated with \\dia\\, clumps (mean $\\sigma_{NT}/c_s = 0.86$). \\dia\\, column densities are similar towards \\dia\\, and continuum clumps (within $\\sim 25$\\,\\%), but \\dia\\, clumps are associated with smaller $N(\\mbox{H$_2$})$ and hence greater $X(\\mbox{\\dia})$ by a factor $> 2$. In fact, the mean \\dia\\, abundance towards continuum clumps is less than $X(\\mbox{\\dia})$ averaged over the entire Oph B Core. Derived volume densities are the same for \\dia\\, and continuum clumps. In summary, the \\dia\\, clumps are associated with smaller line widths and subsonic non-thermal motions relative to continuum clumps and protostars, but the difference in mean values is similar to the variance in line widths across the Core. The most significant difference between \\dia\\, and continuum clumps are found in their \\dia\\, abundances, with a mean $X(\\mbox{\\dia})$ towards continuum clumps that is lower than $X(\\mbox{\\dia})$ in \\dia\\, clumps. \\subsubsection{Oph B2-N6} \\label{sec:n6} The B2-N6 clump is a pocket of narrow \\dia\\, 1-0 line emission within the larger, more turbulent B2 Core, with a large fraction ($\\gtrsim 50$\\%) of the emission at the clump peak coming from the small size scales probed by the ATCA observations (see Figure \\ref{fig:nh3peak}). Although we do not consider all \\dia\\, clumps to trace underlying physical structure in Oph B, the remarkable properties of B2-N6 suggest that further attention be paid to that particular clump. Line widths measured from the interferometer data are a factor $\\sim 2$ less than those recorded from the single-dish data ($\\Delta v = 0.21 \\pm 0.04$\\,\\kms\\, compared with $0.39 \\pm 0.01$\\,\\kms). The millimeter continuum object B2-MM15, identified through a spatial filtering technique \\citep{motte98}, lies offset by $\\sim 12$\\arcsec\\, from the peak interferometer integrated intensity (see Figure \\ref{fig:nh3peak}). \\citeauthor{motte98} calculate a mass of only $0.17$\\,M$_\\odot$ associated with the B2-MM15 clump, but based on its small size (unresolved at 15\\arcsec\\, resolution, using an estimate of $r \\sim 800$\\,AU) derive an average density of $n \\sim 1.2 \\times 10^8$\\,\\cc. This density is significantly larger than the densities calculated in \\S\\ref{sec:nex} at 18\\arcsec\\, resolution, but higher sensitivity \\dia\\, 1-0 interferometer observations are needed to probe the clump density on small scales. We calculated in \\one\\, a gas kinetic temperature $T_K = 13.9\\pm1.1$\\,K for the co-located B2-A7 \\amm\\, (1,1) clump. We find $\\sigma_{\\mbox{\\tiny{NT}}} = 0.06$\\,\\kms\\, for B2-N6, equal to the expected \\dia\\, thermal dispersion, and $\\sigma_{\\mbox{\\tiny{NT}}}/c_s = 0.26$. A similar compact \\dia\\, 1-0 clump with nearly thermal line widths was found in Oph A \\citep[Oph A-N6, ][]{difran04}. While both clumps are associated with larger-scale \\dia\\, emission, these objects are only distinguishable as compact, nearly purely thermal clumps when observed with interferometric resolution and spatial filtering. \\citeauthor{difran04} proposed Oph A-N6 may be a candidate thermally dominated, critical Bonnor-Ebert sphere embedded within the more turbulent Oph A Core. Such objects have sizes comparable to the cutoff wavelength for MHD waves and are in equilibrium between their self-gravity and internal and external pressures. \\citet{myers98} called these objects `kernels', and showed that they could exist in cluster-forming cores with FWHM line widths $\\Delta v > 0.9$\\,\\kms\\, and column densities $N(\\mbox{H$_2$}) > 10^{22}$\\,\\cc, with sizes comparable to the spacing of protostars in embedded clusters, suggesting that a population of kernels within a dense Core could form a stellar cluster. The physical conditions in both Oph A and Oph B2 fit these requirements. The B2-N6 clump, however, is significantly smaller (0.004\\,pc compared with a predicted $\\sim 0.03$\\,pc) and less massive (0.17\\,M$_\\odot$ compared with a predicted $\\sim 1$\\,M$_\\odot$) than the objects discussed by \\citeauthor{myers98}. If we calculate the mass of a critical Bonnor-Ebert sphere given the ATCA \\dia\\, 1-0 line width and a radius of $\\sim 800$\\,AU, where $M/R = 2.4 \\sigma^2/G$ \\citep{bonnor56}, we find $M_{\\mbox{\\tiny{BE}}} = 0.02$\\,M$_\\odot$, or approximately $10\\times$ smaller than the mass determined by \\citeauthor{motte98} We find a virial mass $M_{vir} = 0.04$\\,M$_\\odot$, assuming B2-N6 is a uniform sphere and $M_{vir} = 5 \\sigma^2 R/G$. If the clump mass is accurate, this suggests that while small, B2-N6 may be gravitationally unstable. \\citet{andre07} made a tentative detection of infall motions towards the clump through observations of optically thick lines such as CS, H$_2$CO or HCO$^+$. Thus B2-N6 may be at a very early stage of clump formation, and may eventually form an additional low mass protostar to the three YSOs already associated with dense gas in Oph B2. \\subsection{Comparison of \\dia\\, and \\amm\\, emission in Oph B} We next compare the physical properties of the gas in Oph B derived here from \\dia\\, observations with those derived from \\amm\\, observations, described in \\one. In the following discussion, all comparisons are made after convolving the \\amm\\, data to a final angular resolution of 18\\arcsec\\, to match the Nobeyama \\dia\\, observations. \\subsubsection{$v_{\\mbox{\\tiny{LSR}}}$ and $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$} \\label{sec:sig} We first look at the velocities and non-thermal line widths of the gas traced by \\dia\\, gas and those determined from \\amm\\, emission in \\one. Figure \\ref{fig:compare_nh3}a shows the distribution of $v_{\\mbox{\\tiny{LSR}}}$ derived from \\dia\\, and \\amm\\, emission in Oph B. There is no significant difference in the measured line-of-sight velocity between the two dense gas tracers, with the mean $v_{\\mbox{\\tiny{LSR}}} = 3.98$\\,\\kms\\, for \\amm\\, and 4.01\\,\\kms\\, for \\dia. In Figure \\ref{fig:compare_nh3}b, we show the distribution of $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ as determined with \\dia\\, or \\amm\\, using the \\amm-derived $T_K$ values. The non-thermal line widths measured in \\dia\\, 1-0 are substantially smaller than those measured in \\amm\\, in \\one, and additionally the variation in the relative magnitude of the non-thermal motions between Oph B1 and B2 was not found in \\amm\\, emission. The mean \\amm\\, $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s = 1.64$ in Oph B, and does not vary significantly between B1 ($\\langle \\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s\\rangle = 1.62$) and B2 ($\\langle \\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s\\rangle = 1.68$). The assumption that $T_K$ is similar for both \\dia\\, and \\amm\\, is reasonable if both molecules are excited in the same material. The good agreements between the locations of \\amm\\, and \\dia\\, clumps, illustrated in Figure \\ref{fig:b_sep}, and the $v_{\\mbox{\\tiny{LSR}}}$, illustrated in Figure \\ref{fig:compare_nh3}a, support this assumption. The substantial offset in the $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ ratio seen in Figure \\ref{fig:compare_nh3}b, however, between \\amm\\, and \\dia\\, emission indicates significantly different motions are present in the gas traced by \\amm\\, than in the gas traced by \\dia. The gas densities traced by \\dia\\, are $\\sim 1-2$ orders of magnitude greater than those traced by \\amm\\, ($\\sim 10^5$\\,\\cc\\, and $\\sim 10^{3-4}$\\,\\cc, respectively). In fact, if we attempt to derive the gas kinetic temperature by assuming equal non-thermal motions for the \\dia\\, 1-0 and \\amm\\, (1,1) emission (effectively assuming \\dia\\, and \\amm\\, trace the same material), the resulting average value is a highly unlikely $T_K = 190$\\,K for Oph B2. Given the narrow \\dia\\, lines observed in Oph B1, no physical solution for $T_K$ can be found under the assumption of equal $\\sigma_{\\mbox{\\tiny{NT}}}$. Starless cores are typically found to be well described by gas temperatures that are either constant or decreasing as a function of increasing density \\citep{difran07}. Given that \\dia\\, traces denser gas, we then expect $T_K$ from \\amm\\, measurements to be biased high in starless cores, e.g., gas traced by \\dia\\, should be colder than gas traced by \\amm. The mean $T_K$ found in Oph B in \\one\\, was 15\\,K. With a lower $T_K$, the sound speed would be smaller and the returned $\\sigma_{\\mbox{\\tiny{NT}}}$ would be larger on average. A gas temperature $T_K = 10$\\,K rather than 15\\,K would increase the average \\dia\\, $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ ratio in Oph B by $\\sim 20 - 25$\\,\\%. This increase, while significant, is not large enough for the mean \\dia\\, $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ to match the \\amm\\, results in Oph B. With both a constant temperature or decreasing temperature with density, non-thermal motions in Oph B1 and B3 would remain subsonic, while Oph B2 would still be dominated by transonic non-thermal motions. Alternatively, it is possible that the two Class I protostars in Oph B2 could raise the temperature of the dense gas above that traced by \\amm\\, (no $T_K$ difference was found between protostars and starless areas in \\one). In this case, the returned \\dia\\, $\\sigma_{\\mbox{\\tiny{NT}}}$ would be smaller, and would further increase the differences in magnitude of non-thermal motions seen between the \\amm\\, and \\dia\\, emission in Oph B1. In B2, a temperature of 20\\,K would decrease the mean $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s$ ratio to 1.1. In a study of dust temperatures in the Ophiuchus Cores, \\citet{stamatellos07} showed, however, that embedded protostars in the Cores will only heat very nearby gas and will not raise the mean temperature of the Cores by more than $\\sim 1-2$\\,K, so an average $T_K = 20$\\,K over all Oph B2 is unlikely. \\subsubsection{Double peaked line profiles} \\label{sec:dblpeak} In Figure \\ref{fig:compare_peaks}, we show the isolated $F_1F \\rightarrow F'_1F' = 0 1 \\rightarrow 1 2$ component of the \\dia\\, 1-0 emission line towards the three \\dia\\, clumps in Oph B1 (B1-N1, B1-N3, and B1-N4) which show double peaked line profiles, as described in \\S\\ref{sec:analysis}. Figure \\ref{fig:compare_peaks} also shows a single Gaussian line profile overlaid on each \\dia\\, spectrum, which represents the $v_{\\mbox{\\tiny{LSR}}}$ and $\\Delta v$ of the 18 component Gaussian fit to the \\amm\\, (1,1) emission at the clump peak, normalized to a peak amplitude of 1\\,K. We plot the Gaussian fit for clarity since there are no isolated components in the \\amm\\, (1,1) hyperfine structure. Additionally, spectra of C$_2$S $2_1 - 1_0$ emission line at the peak clump locations are shown. Note that over all Oph B, C$_2$S emission was only detected towards the southern tip of Oph B1, and thus no significant C$_2$S emission was found at the peak location of B1-N1. The C$_2$S emission was observed with the GBT at $\\sim 32$\\arcsec\\, spatial resolution and 0.08\\,\\kms\\, spectral resolution, with the observations and analysis described in \\one. Arrows show the locations of the fitted $v_{\\mbox{\\tiny{LSR}}}$ for all species, including both \\dia\\, velocity components. The clumps which show the double-peaked line profile structure are the clumps with the highest total \\dia\\, 1-0 line opacities based on a single velocity component fit, with $\\tau = 5, 6$ and 9, respectively for B1-N1, B1-N3 and B1-N4. The $F_1F \\rightarrow F'_1F' = 0 1 \\rightarrow 1 2$ component is expected to have an opacity $\\tau_{iso} = 1/9 \\,\\tau$, so $\\tau_{iso} = 0.6, 0.7$, and $\\sim 1$ for the three clumps, based on a single component fit. If the line is optically thick, however, this fit is likely to underestimate the true opacity given the missing, self-absorbed flux. This is suggestive that the lines are indeed self-absorbed. Since the \\amm\\, and C$_2$S emission is optically thin (\\one), however, if the \\dia\\, emission was self-absorbed due to high optical depth we would expect the emission peak of both \\amm\\, and C$_2$S to be found between the two \\dia\\, components in $v_{\\mbox{\\tiny{LSR}}}$, which is not the case. In all three clumps, we find the $v_{\\mbox{\\tiny{LSR}}}$ of the \\amm\\, emission more closely matches the $v_{\\mbox{\\tiny{LSR}}}$ of the red \\dia\\, component, while the C$_2$S emission in two clumps ($v_{\\mbox{\\tiny{LSR}}} = 3.67$\\,\\kms) more closely matches the $v_{\\mbox{\\tiny{LSR}}}$ of the blue \\dia\\, component ($v_{\\mbox{\\tiny{LSR}}} = 3.85$\\,\\kms\\, and 3.62\\,\\kms\\, for B1-N3 and B1-N4, respectively). This behaviour suggests the \\dia\\, 1-0 emission is not self-absorbed towards these positions. A similar offset of C$_2$S emission relative to \\dia\\, was found by \\citet{swift06} towards the starless Core L1551, however the case of L1551 the C$_2$S emission is redshifted with respect to the systemic velocity of the Core rather than blueshifted, as we find in B1. The GBT \\amm\\, (1,1) observations described in \\one\\, found extensive emission around Oph B, such that it is possible to determine the $v_{\\mbox{\\tiny{LSR}}}$ and $\\Delta v$ of \\amm\\, beyond where continuum emission traced the Oph B Core. We use an intensity threshold of 2\\,K in the \\amm\\, (1,1) main component to separate Core and off-Core gas. The off-Core gas surrounding Oph B1 has a mean $v_{\\mbox{\\tiny{LSR}}} = 3.72$\\,\\kms\\, with an rms variation of 0.13\\,\\kms, which is significantly different from the mean B1 Core $v_{\\mbox{\\tiny{LSR}}} = 3.96$\\,\\kms\\, and rms variation of 0.15\\,\\kms. Both the C$_2$S emission and the blue \\dia\\, component are thus more kinematically similar to the off-Core \\amm\\, gas. The critical densities of C$_2$S $2_1 - 1_0$ and \\amm\\, (1,1) are similar \\citep{suzuki92,rosolowsky08}. Kinematically, we find $\\sigma_{\\mbox{\\tiny{NT}}}\\,/\\,c_s = 0.7$ for the C$_2$S emission if we assume the $T_K$ determined from \\amm\\, emission accurately represents the temperature of the gas traced by C$_2$S. This result closely matches the \\dia\\, results in Oph B1, but the non-thermal motions in the gas traced by \\amm\\, are significantly larger (\\S\\ref{sec:trends} and \\one). It is impossible to match the non-thermal C$_2$S and \\amm\\, motions for a given $T_K$, since the \\amm\\, (1,1) line $\\sigma_{\\mbox{\\tiny{NT}}} > \\sigma_{obs}$ of the C$_2$S. At densities $n \\gtrsim$ a few $\\times 10^3 - 10^4$\\,\\cc, chemical models predict C$_2$S is quickly depleted from the gas phase \\citep[timescale $t_{dep} \\sim 10^5$\\,yr;][]{millar90,bergin97}, and accordingly the molecule is observed to be a sensitive tracer of depletion in isolated cores \\citep{lai00,tafalla06}. The volume density in Oph B1 is greater than that required for C$_2$S depletion ($n \\gtrsim 10^5 - 10^6$\\,\\cc\\, from \\S\\ref{sec:nex} and continuum observations), so we would not expect to see any C$_2$S emission unless the gas has only been at high density for $t < t_{dep}$. C$_2$S was only detected towards southern B1. An explanation, therefore, for both the presence of C$_2$S in southern B1 and for its velocity offset relative to the dense gas is that gas from the ambient molecular cloud is accreting onto B1, reaching a density high enough to excite the C$_2$S emission line. This material would be chemically `younger' than the rest of the B1 Core gas, such that the C$_2$S has not had enough time to deplete from the gas phase. A similar result was found in a multi-species study of the starless L1498 and L1517B cores, where \\citet{tafalla04} conclude that non-spherical contraction of the cores produced asymmetric distributions of CS and CO, where the CS and CO `hot spots' reveal the distribution of recently accreted, less chemically processed dense gas. A single 850\\,\\micron\\, continuum clump \\citep[162715-24303;][]{jorgensen08} was identified in southern Oph B1 where we find B1-N3 and B1-N4. The authors found a clump mass $M = 0.2\\,M_\\odot$, which is a factor $\\sim 2$ less than the virial mass $M_{vir} = 0.4\\,M_\\odot$ we calculate based on the mean line width (of the single velocity component fit) of B1-N3 and B1-N4 ($\\Delta v = 0.52$\\,\\kms, or $\\sigma = 0.22$\\,\\kms) and the continuum clump radius. This suggests that the clump is currently stable against collapse. If the clump is gaining mass through ongoing accretion from the ambient gas, however, then eventually the non-thermal support may not be large enough to prevent gravitational collapse, leading to the formation of a low-mass protostar. \\subsubsection{Opacity and Excitation Temperature} We show in Figure \\ref{fig:compare_nh3}c and \\ref{fig:compare_nh3}d the distribution of total line opacity $\\tau$ and excitation temperature $T_{ex}$ derived from hyperfine line fitting of \\amm\\, (1,1) and \\dia\\, 1-0 in Oph B. We find higher total opacities in \\dia\\, emission ($\\langle \\tau \\rangle = 2.5$) than in \\amm\\, emission ($\\langle \\tau \\rangle = 1.0$) in Oph B. The \\dia\\, emission is characterized by lower excitation temperatures than those found for \\amm\\, emission in \\one, with a mean $T_{ex} = 7.1$\\,K for \\dia\\, 1-0 compared with $T_{ex} = 9.5$\\,K for \\amm\\, (1,1). \\subsubsection{Fractional Abundances and Chemical Evolution} \\label{sec:frac} We next compare the column densities of \\amm\\, and \\dia\\, towards the Oph B Core. We only show results for pixels which have well-determined values for both $N(\\mbox{\\amm})$ and $N(\\mbox{\\dia})$. In \\one, we found a trend of decreasing fractional \\amm\\, abundance with increasing $N(\\mbox{H$_2$})$ for {\\it both} B1 and B2. After convolving the \\amm\\, data to match the single-dish \\dia\\, resolution, we find the following linear relationships between $\\log X(\\mbox{\\amm})$ and $\\log N(\\mbox{H$_2$})$: \\begin{eqnarray} \\mbox{B1}&:& \\log X(\\mbox{\\amm}) = (5 \\pm 4) - (0.6 \\pm 0.2) \\log N(\\mbox{H$_2$}) \\\\ \\mbox{B2}&:& \\log X(\\mbox{\\amm}) = (3.5 \\pm 2.2) - (0.5 \\pm 0.1) \\log N(\\mbox{H$_2$}) \\end{eqnarray} \\noindent The slopes found for Oph B1 and B2 are both negative and similar to that found in Equation \\ref{eqn:xvsh} for the \\dia\\, fractional abundance as a function of $N(\\mbox{H$_2$})$. This result is suggestive of a trend of increasing depletion of both \\amm\\, and \\dia\\, with increasing $N(\\mbox{H$_2$})$ in Oph B. We show in Figure \\ref{fig:compare_column} the \\amm\\, column density, $N(\\mbox{\\amm})$, versus the \\dia\\, column density, $N(\\mbox{\\dia})$ in Oph B1 and B2, and also the ratio of $N(\\mbox{\\amm})$ to $N(\\mbox{\\dia})$ as a function of $N(\\mbox{H$_2$})$, calculated in \\S\\ref{sec:nh2} (we do not show results in B3 due to the small number of pixels with both well-determined \\amm\\, and \\dia\\, column densities, but note mean values below). Note that the column density ratio, $N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia})$, is equivalent to the ratio of fractional abundances, $X(\\mbox{\\amm})\\,/\\,X(\\mbox{\\dia})$. Figure \\ref{fig:compare_column}a shows a significant difference in the $N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia})$ ratio between Oph B1 and B2. We find the mean and the standard deviation of this ratio are each twice as large in B1 ($N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia}) = 135$, $\\sigma = 52$) as in B2 ($N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia}) = 65$, $\\sigma = 31$) and B3 ($N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia}) = 80$, $\\sigma = 56$). In Oph B1 and B2, Figure \\ref{fig:compare_column}a illustrates that this difference is largely driven by lower $N(\\mbox{\\dia})$ values towards B1, while the spread in the $N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia})$ appears due to a wider spread of $N(\\mbox{\\amm})$ values in B1. We find no variation in the $N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia})$ ratio as a function of $N(\\mbox{H$_2$})$. In the clustered star forming region IRAS 20293+3952, \\citet{palau07} found strong \\amm\\, and \\dia\\, differentiation associated with the presence or lack of embedded YSOs. Low $N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia}) \\sim 50$ values were found near an embedded YSO cluster, while higher $N(\\mbox{\\amm})\\,/\\,N(\\mbox{\\dia}) \\sim 300$ were found towards cores with no associated YSOs. In a study of two dense cores, each containing a starless main body and a YSO offset from the core center, \\citet{hotzel04} find a factor $\\sim 2$ smaller $X(\\mbox{\\amm})\\,/\\,X(\\mbox{\\dia})$ values towards the starless cores compared with \\citeauthor{palau07}, with abundance ratios ($\\sim 140-190$, towards the starless gas, and $\\sim 60-90$ towards the YSOs) similar to those found in this study. The relative abundance variation found by \\citeauthor{hotzel04} is driven by a varying \\amm\\, abundance while $X(\\mbox{\\dia})$ remains constant over the cores. Although we find both a decrease in \\dia\\, column density as well as an increase in \\amm\\, column density drives the greater fractional \\amm\\, abundance towards Oph B1 (see Figure \\ref{fig:compare_column}), both $X(\\mbox{\\dia})$ and $X(\\mbox{\\amm})$ are greater in B1, by factors of $\\sim 2$ and $\\sim 4$, respectively, compared with B2 given the lower H$_2$ column densities found in B1 in \\S\\ref{sec:nh2}. In a survey of 60 low mass cloud cores, \\citet{caselli02} find correlations between $N(\\mbox{\\amm})$ and $N(\\mbox{\\dia})$ in starless cores, but the relative column density values do not vary significantly between starless and protostellar objects. Based on this work and the studies described above, it appears that the relative fractional abundance of \\amm\\, to \\dia\\, remains larger towards starless cores than towards protostellar cores by a factor of $\\sim 2 - 6$ in both isolated and clustered star forming regions. \\edits{We find a 1-$\\sigma$ positive trend in the relative \\amm\\, to \\dia\\, abundance in Oph B2 with increasing distance to an embedded protostar, suggestive of a direct impact by the protostars on the relative abundances. In their work on x-ray emission from young protostars in Ophiuchus, \\citet{casanova95} show that protostellar x-ray emission is potentially significant enough to increase the ionization fraction in nearby dense gas, potentially affecting the local gas chemistry. In this region, the complicated line structures found in both species near the protostars make accurate column density measurements difficult, however, with resulting large uncertainties. } In their models of collapsing prestellar cores, \\citet{aikawa05} found that \\amm\\, can be enhanced relative to \\dia\\, in core centers at high densities ($n = 3 \\times 10^5 - 3 \\times 10^6$\\,\\cc) due to dissociative recombination reactions of \\dia\\, and $e^-$ to form NH and N. NH then reacts with H$_3^+$ and H$_2$ to form \\amm. The \\dia\\, recombination reaction is dominant only where CO is depleted (if CO is abundant, \\dia\\, is destroyed mainly through proton transfer to CO), and occurs when the abundance ratio of CO to electrons, $n(\\mbox{CO})\\,/\\,n_e \\lesssim 10^3$. When a collapsing core reaches higher central densities ($n \\sim 10^7$\\,\\cc), however, \\citeauthor{aikawa05} predict that the $N(\\mbox{\\amm})$ will begin to decrease towards the core centre due to depletion, while $N(\\mbox{\\dia})$ continues to increase, leading to a higher fractional \\dia\\, abundance relative to \\amm\\, at later times. This prediction is in agreement with $X(\\mbox{\\amm})/X(\\mbox{\\dia})$ results since B2 is more evolved than B1, having already formed protostars, but seems inconsistent with the common slope we find for the decrease in $X(\\mbox{\\amm})$ and $X(\\mbox{\\dia})$ with $N(\\mbox{H$_2$})$ in both Cores. Detailed modelling of the physical and chemical structure of the (non-spherical and clumpy) Cores is beyond the scope of this study, but would help to constrain the central density and abundance structure as a function of Core radius needed for a direct comparison with the \\citeauthor{aikawa05} models. \\subsection{Are \\dia\\, 1-0 and \\amm\\, (1,1) tracing the Oph B Core interior?} The hyperfine structure of the \\amm\\, (1,1) inversion transition allows the calculation of volume density $n_{ex}$ as in Equation \\ref{eqn:nex}, with the opacity of a typical hyperfine component $\\tau_{\\mbox{\\tiny{\\it hf}}} = 0.233 \\tau$. The resulting mean density $n_{ex, \\mbox{\\tiny{\\amm}}} = 7.8 \\times 10^3$\\,\\cc\\, and $n_{ex, \\mbox{\\tiny{\\amm}}} = 1.7 \\times 10^4$\\,\\cc\\, for Oph B1 and B2, respectively (note that these values are slightly less than reported in \\one\\, due to using the larger 18\\arcsec\\, beam FWHM). These volume densities are a factor $\\gtrsim 20$ less than those calculated in \\S\\ref{sec:nex} from \\dia\\, 1-0 emission ($\\sim 26$ in B1 and $\\sim 18$ in B2). There is some question whether the \\amm\\, volume densities are accurate, as recent studies \\citep[\\one; ][]{foster09} have found \\amm-derived $n_{\\mbox{\\tiny{ex}}}$ values which are an order of magnitude less ($10^4$\\,\\cc\\, versus $10^5$\\,\\cc\\, and greater) than volume densities determined from continuum emission. One possible source of error is the collisional de-excitation rate coefficient, $\\gamma_{ul}$, which \\citeauthor{foster09} note is reported in various publications with factor of $\\sim 10$ variation. Since $n_{cr} = A_{ul}\\,/\\,\\gamma_{ul}$, the critical density is therefore also suspect within a factor $\\sim 10$. It is clear that in Oph B \\amm\\, and \\dia\\, are tracing different motions in the gas. The gas traced by \\dia\\, emission is significantly more quiescent, as has been observed in high density gas in other star forming cores. The fact that higher densities are calculated from \\dia\\, emission relative to those calculated from \\amm\\, emission is thus reasonable, and bolsters the hypothesis that \\amm\\, (1,1) emission is not tracing the highest density gas. \\edits{How well does the \\dia\\, 1-0 emission trace the highest density gas?} In agreement with results in \\one, we find that the clumps found in \\dia\\, 1-0 emission also do not match well continuum clump locations, and significant offsets between \\dia\\, 1-0 and continuum emission are also apparent in the integrated intensity maps. In particular, both the single-dish- and interferometer-integrated \\dia\\, 1-0 intensity towards the continuum clump B2-MM8 peak offset to the clump, and in the interferometer data we see an integrated intensity minimum at the MM8 location. On larger scales, we find a trend of {\\it decreasing} \\dia\\, abundance with increasing H$_2$ column density, described above, with a slope in Oph B2 that agrees within uncertainties with that determined for $N(\\mbox{\\amm})$ versus $N(\\mbox{H$_2$})$ for both Oph B1 and B2 (see \\S\\ref{sec:trends} and \\S\\ref{sec:frac}). \\edits{Furthermore, the mean $X(\\mbox{\\dia})$ for continuum clumps alone in Oph B is lower than the $X(\\mbox{\\dia})$ over the whole core (see \\S\\ref{sec:compare_clumps}). Together, these findings suggest that \\dia\\, 1-0, though doing a better job than \\amm\\, (1,1), may not itself trace well the coldest, densest material in Oph B. Appropriate caution must be made when using even \\dia\\, 1-0 to probe dense gas in some star-forming regions.}" }, "0911/0911.3055_arXiv.txt": { "abstract": "ANTARES is a submarine neutrino telescope deployed in the Mediterranean Sea, at a depth of about 2500 m. It consists of a three-dimensional array of photomultiplier tubes that can detect the Cherenkov light induced by charged particles produced in the interactions of neutrinos with the surrounding medium. Down-going muons produced in atmospheric showers are a physical background to the neutrino detection, and are being studied. In this paper the measurement of the Depth Intensity Relation (DIR) of atmospheric muon flux is presented. The data collected in June and July 2007, when the ANTARES detector was in its 5-line configuration, are used in the analysis. The corresponding livetime is $724\\,h$. A deconvolution method based on a Bayesian approach was developed, which takes into account detector and reconstruction inefficiencies. Comparison with other experimental results and Monte Carlo expectations are presented and discussed. ", "introduction": "The largest event source in neutrino telescopes is \\textit{atmospheric muons}, particles created mainly by the decay of $\\pi$ and $K$ mesons originating in the interaction of cosmic rays with atmospheric nuclei. Although ANTARES \\cite{antares_web, antares_daq, antares_OM, antares_OM1, pcoyle} \"looks downwards\" in order to be less sensitive to signals due to downward going atmospheric muons, these represent the most abundant signal due to their high flux. They can be a background source because they can be occasionally wrongly reconstructed as upward going particles mimicking muons from neutrino interactions. On the other hand they can be used to understand the detector response and possible systematic effects. In this scenario the knowledge of the underwater $\\mu$ intensity is very important for any Cherenkov neutrino telescope and the future projects \\cite{km3net, nemo}. Moreover, it would also provide information on the primary cosmic ray flux and on the interaction models. \\\\ ", "conclusions": "{}} The aim of the presented analysis is the measurement of the muon flux at the depth of ANTARES and the derivation of the vertical component of the atmospheric muon flux as a function of the sea depth. The goal is also to assess the performance of ANTARES in detecting muons. The analysis has been performed on a selection of the experimental data of June and July 2007 when the ANTARES detector was in its 5-line configuration. Several quality cuts have been applied on the reconstructed events in order to improve their purity, in particular concerning the zenith angle reconstruction. An unfolding algorithm, based on an iterative method, has been applied on the selected experimental data in order to retrieve back the flux of atmospheric muons with $E_\\mu>20\\,GeV$ at the fixed sea depth $h_0=2000\\,m$. The experimental DIR was finally obtained. The results are in good agreement, within the uncertainties, with the experimental fluxes obtained by other Cherenkov telescopes. \\begin{figure}[!t] \\begin{center} \\includegraphics[width=8. cm] {bazzotti_fig5} \\caption{\\label{DIRallNEW} \\textbf{PRELIMINARY}. Depth Intensity Relation of atmospheric muons for $E_\\mu>20\\,GeV$, with systematic uncertainties (the statistical uncertainties are negligible). The DIR obtained from other underwater measurements are also shown: Higashi \\cite{d-higa}, Davitaev \\cite{d-davitaev}, Vavilov \\cite{d-vavilov}, Fyodorov \\cite{d-fyodorov}, DUMAND-SPS \\cite{dumand}, BAIKAL NT-36 \\cite{baikal2}, NESTOR \\cite{d-nestor}, AMANDA B-4 \\cite{d-amandab4}, AMANDA-II \\cite{d-amandaII}. The Sinegovskaya parameterization ($E_\\mu>20\\,GeV$) \\cite{sine} and the MUPAGE simulation are superimposed.} \\end{center} \\end{figure}" }, "0911/0911.2786_arXiv.txt": { "abstract": "This talk briefly explains how the breaking of a Lorentz-invariant description of nature at tiny space-time intervals might affect the non-Gaussian character of the primordial fluctuations left by inflation. For example, a model that contains irrelevant operators that only preserve the spatial symmetries along constant-time surfaces can generate a larger non-Gaussian component in the pattern of primordial fluctuations than is ordinarily predicted by inflation. This property can be useful for constraining models that allow some Lorentz violation at short distances, beyond the constraints possible from the power spectrum alone. ", "introduction": "The universe today is filled with a quite bewildering variety of structures, from stars and galaxies at comparatively small scales (when judged by the size of the observable universe) to the vast clusters and networks of galaxies and emptier voids among them at the largest scales. But the earlier universe differs markedly from its current appearance. The earliest relics that we can see directly show a remarkable uniformity in how the material was distributed at those times. For example, the temperature of the radiation that first escaped as the universe cooled from a hot plasma into a transparent gas varies less than one part in one hundred thousand from one place to another. Moreover, the abundances of the primordially produced lighter elements also seem to be the same everywhere in the universe, suggesting that this uniformity extends to still earlier times. Yet the universe needed to have had at least {\\it some\\/} variation, even at the very earliest of times, for without any tiny spatial variations, the growth of the structures that we see today could never have begun in the first place. The theory of inflation provides one elegant mechanism for explaining the origin of these tiny {\\it primordial fluctuations\\/}. One very appealing property of the inflationary picture is that such fluctuations are unavoidable, being a simple consequence of the two basic ingredients needed in any inflationary model. A typical model for inflation requires a space-time that is expanding at an accelerating rate and a quantum field whose dynamics are responsible for the expansion. But a quantum field is always fluctuating. It would be quite impossible for it to be otherwise, as it would be inconsistent with the Heisenberg principle for a field to be at once stationary and without any spatial fluctuations. Let us describe this behavior a little more precisely. If we denote the quantum field by $\\varphi(t,\\vec x)$ and its quantum state by $|0(t)\\rangle$, then while its mean value in this state may vanish on average, \\begin{equation} \\langle 0(t)| \\varphi(t,\\vec x) |0(t)\\rangle = 0 , \\label{onepoint} \\end{equation} the variance of the field never does, \\begin{equation} \\langle 0(t)| \\varphi(t,\\vec x) \\varphi(t,\\vec y) |0(t)\\rangle \\not= 0 . \\label{variance} \\end{equation} Of course, such quantum fluctuations are also occurring at tiny scales today. The difference between the early universe and the universe today is that, during the former, the space-time expands very rapidly, stretching these tiny fluctuations to very large scales---even as large as the observable universe today. Because these fluctuations are such a basic consequence of inflation, the pattern that inflation predicts for them is fairly insensitive to the details of a particular model, though measurements have actually reached a precision that can test inflation beyond its most basic predictions and can even exclude individual models. Nevertheless, the overall inflationary picture is still in a superb agreement with what is inferred from observations. Inflation predicts \\begin{itemize} \\item that the fluctuations have more or less the same amplitude at all length scales---including those that otherwise would not have been in causal contact without an inflationary period, \\item that the fluctuations are all in phase and are adiabatic, \\item that their pattern is largely a {\\it Gaussian\\/} one, the meaning of which we shall explain later, and \\item that there should also be a background of primordial gravity waves. \\end{itemize} Each of these predictions---except for the last---matches what is seen in all experiments so far. With this success, it is important to test the underlying assumptions of the inflationary picture further and to shed some light on its otherwise mysterious elements. What exactly is this field that is producing the inflation and how does it fit into the rest of our picture for particle physics? How did the universe find itself initially in a spatially flat configuration, at least to a sufficient degree over a sufficiently large patch? Does the size of the field change rapidly over small enough distances that we ought to worry about quantum gravity? Which is the correct quantum state $|0(t)\\rangle$ to choose? And is it possible to have {\\it too much\\/} expansion of the quantum fluctuations for the consistency of our picture? If these questions seem to be only so much cavilling, it instructive to compare the setting used in inflation with that of ordinary particle physics to realize how much less we know about the former and how much more limited is our ability to test it. In particle physics, \\begin{itemize} \\item We know the background (flat space). \\item We know all the relevant ingredients up to some energy scale (and any unseen stuff remains largely decoupled from the lower energy world). \\item We have a well developed, {\\it and well tested\\/}, theoretical framework (quantum field theory in flat space). \\item Gravity is entirely negligible. \\item And perhaps most importantly, we are free to test anything we want, in a controlled environment, up to a limiting energy scale. \\end{itemize} Now contrast each of these points with what we assume for the early universe. \\begin{itemize} \\item We probably know the background (it seems to be approaching a spatially independent one at early enough times). \\item We can only guess what are relevant ingredients, and have no way to observe them directly (at least so far, and probably for a long while yet to come). \\item We have a framework, quantum field theory in curved space; and while it is generally self-consistent, it has not been tested. \\item Gravity is an essential ingredient. \\item And experimentally, we can only observe a very limited and indirect set of evidence, and we are not able to test the picture directly (at least not to the standard applied in an accelerator experiment). \\end{itemize} Nothing on this list implies that the inflationary picture is necessarily wrong; but each item does indicate some unresolved piece of the inflationary framework which we would like to understand better. Because of its remoteness from anything that we can test directly and because it will probably be difficult to find {\\it complementary\\/} observations to those that we already have which can test this picture, it is important to address some of these questions if we are to have a little more confidence that inflation really did occur in our own universe. ", "conclusions": "" }, "0911/0911.0921_arXiv.txt": { "abstract": "We evaluate the achievable maximum energy of nuclei diffusively accelerated by shock wave in the jet of Cen\\,A, based on an updated model involving the stochastic magnetic fields that are responsible for recent synchrotron X-ray measurements. For the maximum energy analysis, conceivable energy constraints from spatiotemporal scales are systematically considered for the jet-wide including discrete X-ray knots. We find that in the inner region within $\\sim 1\\,{\\rm arcmin}$ from galactic core, which includes knots AX and BX, proton and iron nucleus can be accelerated to $10^{19}-10^{20}$ and $10^{21}~{\\rm eV}$ ($10-100~{\\rm EeV}$ and ${\\rm ZeV}$) ranges, respectively. The upper cutoff energy of the very energetic neutrinos produced via photopion interaction is also provided. These are essential for identifying the acceleration site of the ultra-high energy cosmic ray detected in the Pierre Auger Observatory, which signifies the arrival from nearby galaxies including Cen\\,A. ", "introduction": "To date, the 27 ultra-high energy cosmic ray (UHECR) events of the energy exceeding $5.7\\times 10^{19}~{\\rm eV}$ have been detected in state-of-the-art Auger observatory; in particular, the anisotropy and significant correlation with the active galactic nuclei (AGNs) residing within $75~{\\rm Mpc}$, namely GZK horizon \\citep{greisen,zatsepin}, were discovered \\citep{abraham07}. Surprisingly, the results indicate that two of these UHECR events have arrived within $3\\degr$ of Centaurus\\,A (NGC\\,5128), a closest galaxy \\citep[$3.7~{\\rm Mpc}$ from us;][]{ferrarese}. Although the detailed position of the UHECR production site is still unresolved, the galactic core accompanied by a supermassive black hole, bipolar jets, giant radio lobes \\citep{hardcastle09}, and so on, will be enumerated as the favored candidates. It is desired that genuine theoretical survey be expanded for clarifying the particle acceleration mechanism feasible at these sites, in order to provide the physical interpretation of the observed intriguing results, which include the recent detection of high-energy gamma rays \\citep{aharonian}. When closely looking at the anatomy of the large-scale jets in the Cen\\,A galaxy \\citep[see,][for a review]{israel}, a clue to solve this challenging problem seems to be in the knotty regions resolved in a deep X-ray image \\citep[e.g.,][]{kraft02}. In an inner region near the galactic core, the rugged features are associated with the shocks, which are considered to be formed where a flowing plasma collides with obstacles \\citep{hardcastle07}. According to the latest theory of plasma kinetic transport, it is known that such a colliding plasma is likely unstable for the electromagnetic current filamentation instability, which generates small-scale magnetic fluctuations with the order of plasma skin depth (\\citealt{medvedev}; \\citealt{honda04} and references therein). In the nonlinear phase, the filaments preferentially coalesce one another, self-organizing larger scale filaments with stronger magnetic fields. This reflects a stochastic nature of the inverse energy cascade of plasma turbulence. The magnetic fluctuations are non-perturbative, and in the strong turbulence regime, in that the energy density is comparable to the thermal energy density of plasma bulk \\citep{honda00a,honda00b}. Indeed, recent high-resolution observations reveal that the jet of Cen\\,A is in part dominated by the filamentary, sometimes edge-brightened features \\citep{hardcastle07}, as inspected in the well-confirmed jet of a nearby galaxy \\citep[Vir\\,A;][]{owen}. Worth noting thing is that inherent inner structure similar to these has been discovered in many other jets (e.g., Cyg\\,A: \\citealt{perley}; 3C\\,353: \\citealt{swain}; 3C\\,273: \\citealt{lobanov}; 3C\\,438: \\citealt{treichel}; Mrk\\,501: \\citealt{piner}). Also, it was recently found that a filamentary jet model naturally provides the comprehensive explanation for the complicated spectral variability observed in a TeV blazar object \\citep[Mrk\\,421;][]{honda08}. These ingredients strongly encourage us to take the inhomogeneous magnetic field effects into account for precisely modeling the Cen\\,A jet as a cosmic-ray accelerator. For a standard, diffusive shock acceleration (DSA) scenario \\citep[e.g.,][]{blandford87}, the strong turbulence plays an essential role in particle scatterers is expected. In the stochastic medium, the higher energy heavy particles tend to freely meander \\citep{hh05}, albeit electrons (and positrons) are, if anything, likely gyro-trapped in the magnetized filaments, suffering stronger radiative cooling (Section~3.1). From the fact that in the Cen\\,A jet, electron synchrotron emissions appear to be, in part, already diffusive, it is thus inferred that for nuclei (say, proton) the conventional approximation of small-angle resonant scattering will be inadequate for describing the spatial diffusion, and instead, the three-dimensional rms deflection becomes rather feasible. Importantly, the latter facilitates the back and forth of particles across the shock, increasing the efficiency of DSA. Based on this notion, \\citet{hh04} have argued that a nucleus could be accelerated to the UHE range at a bright jet knot of Vir\\,A, though the arrival from the galaxy ($\\sim 4$ times more distant than the distance to Cen\\,A) is not yet signified \\citep[see, e.g.,][for the useful discussions]{stanev}. Besides, the bound electrons are co-accelerated, to emit the synchrotron photons, which could serve as a target of the accelerated protons. Then, the resonant interaction of the UHE protons with the target photons could be a major neutrino production process via the decay channels triggered by photopionization: $p\\gamma\\rightarrow \\pi^{\\pm}X\\rightarrow\\mu^{\\pm}\\nu_{\\mu} \\rightarrow e^{\\pm}\\nu_{e}\\nu_{\\mu}$ \\citep{romero}. The emitted neutrinos propagate on the straight, unlike the charged particles that suffer, more or less, deflection by intergalactic magnetic fields \\citep[e.g.,][]{vallee}, so that they play the role of a powerful and complementary messenger of the UHECR production, in light of the observability at a forthcoming kilometer neutrino telescope \\citep[e.g.,][]{halzen}. Recently, \\citet{cuoco} have proposed the model spectra of high-energy neutrinos from Cen\\,A, but putting the detailed mechanisms of the particle accelerator into effect has remained unsolved. In this paper, based on the filament model, we estimate the maximum possible energies of a proton and iron diffusively accelerated in the Cen\\,A jet, and also, the energy of the produced neutrinos, providing the energy-equipartition among the flavors. The conceivable mechanisms of the energy restriction are taken into consideration for the jet wide including bright knots. It is addressed that the energy is limited dominantly by the shock operation time or particle escape loss, rather than radiative loss and nucleus--nucleus collision timescales. As a result, we find that in the inner region, which contains the X-ray knots BX$n$, proton and iron can be accelerated beyond the Auger limit. In particular, the expected highest energy reaches, for iron, $3~{\\rm ZeV}$ and more (around knot BX5); that is to say, the Cen\\,A jet is a \"Zevatron\" worthy of the candidate (in addition to the Vir\\,A, put forth by \\citet{hh04}). The present analysis virtually provides a substantially extended version of the previous simple analysis by \\citet{romero}, and hence, a particular attention has been paid to highlight the new points including (1) the proposal of the improved theoretical model compatible with updated X-ray measurements (Section~2.1) and extended arguments on the turbulent magnetic field (Section~2.2), (2) the non-resonant diffusion scenario involved in the DSA (Section~3.1), and (3) thorough survey of the temporal (Section~3.2) and spatial (Section~3.3) limits, in order to figure out the maximum energies of proton and iron achievable in the inner region (Section~3.4), and estimate neutrino energy (Section~3.5). The application of the maximum energy analysis to the outer region of the large-scale jet is also provided (Section~4). At last, the discussion on the comparison with the previous relevant results is expanded (Section~5). ", "conclusions": "The validity of the filamentary model has been checked by reproducing the complicated blazar variabilities that involve the non-monotonous hysteresis patterns of broadband electromagnetic spectrum in flare phases \\citep{honda08}. By recalling the idea that regards FR-I radio sources as misaligned BL\\,Lac objects \\citep{urry,tsvetanov}, the application of the filamentary model to the Cen\\,A jet will be reasonably justified. As for the particle acceleration in the filamentary medium, we point out that the present DSA mechanism for nuclei, which owes to the (off-resonant) three-dimensional rms diffusion across the shock (Section~3.1), is simpler than the electron acceleration mechanism that is incorporated with the complexity of the energy hierarchy mediated by the transition injection \\citep{hh07}. The acceleration of highest energy particle in the Cen\\,A was first considered by \\citet{romero}. In the early work, they dealt with the conventional gyro-resonant diffusion model even for proton, that is to say, implicitly supposed a large-scale ordered magnetic field (with small-amplitude perturbations). They addressed that a proton could be accelerated to the energy of $2.7\\times 10^{21}~{\\rm eV}$ in situ, which is an order of magnitude larger than the present $E_{p,m}$ values that are in the range of $\\sim 10^{19}-10^{20}~{\\rm eV}$ (Section~3.4; Table~\\ref{tbl:2}). Such a larger value was derived from simply equating a classical acceleration timescale \\citep[e.g.,][]{biermann} with radiative loss timescales, but this fashion now appears to be somewhat optimistic (Section~3.2.2; Figure~2). It can be claimed that taking into account of the shock operation time or particle escape is essential for the maximum energy analysis for the concerned source, to yield the reduced proton energy, which can hardly reach the $10^{21}~{\\rm eV}$ energy range. It is also mentioned that the situation in which the dynamical timescale limits the particle acceleration is analogous to the situation that typically appears in the supernova remnant shock acceleration, in which the adiabatic expansion of spherical shell likely becomes a major loss mechanism \\citep[e.g.,][]{kobayakawa}. In conclusion, we have evaluated the achievable maximum energies of the proton and iron nucleus diffusively accelerated by the shock in the Cen\\,A jet including X-ray knots. In particular, we have taken into account of the more realistic DSA scenario that relies on the filamentary jet model responsible for the recent X-ray measurements, and elaborated the conceivable energy restriction stemming from spatiotemporal scales. The key finding is that, for the plausible ranges of the shock speed of $\\sim 0.1c$, viewing angle of $\\gtrsim 10\\degr$, and magnetic intensity of $\\lesssim 500~{\\rm\\mu G}$, the uppermost particle energy tends to be inevitably limited by the shock accelerator operation timescale ($<10^{13}~{\\rm s}$), rather than the radiative cooling losses ($\\gtrsim 10^{13}~{\\rm s}$). As a result, it has been demonstrated that there exists the acceleration region (off the galactic core), in which proton and iron can be energized to $\\sim 10^{20}~{\\rm eV}$ and $\\sim 10^{21}~{\\rm eV}$ ranges, respectively. In particular, for the inner region including the X-ray knot BX5, we work out at the maximum energies of $1\\times 10^{20}~{\\rm eV}$ and $3\\times 10^{21}~{\\rm eV}$ for proton and iron, respectively; and estimate the corresponding cutoff energy of the neutrino produced via the photopionization, as $6\\times 10^{18}~{\\rm eV}$. For the stationary shock case, we can expect the enhancement of the maximum energies, and read that the large-scale region up to the projection of about $200\\arcsec$, as well, has a potential to energize proton and iron beyond $10^{20}~{\\rm eV}$ and $10^{21}~{\\rm eV}$, respectively, to yield the $p\\gamma$ neutrino with the energy exceeding $5\\times 10^{18}~{\\rm eV}$. We here manifest that the jet wide of the Cen\\,A galaxy is a promising candidate for the UHE proton accelerator and Zevatron for high-$Z$ nucleus, as desirable to account for the outcome from the Auger observatory \\citep{abraham07}. The derived large values of particle energies are of those achievable at the acceleration sites. The observable energies ought to be, of course, reduced, due to a major energy loss mechanism including the photopion interaction with cosmic microwave background. However, the Cen\\,A is so nearby that the ballistic transport is anticipated to be not crucially degraded, on account of the weak decay property (such as $-(1/E)(dE/dt)\\sim 10^{-8}~{\\rm yr^{-1}}$ for $E>10^{20}~{\\rm eV}$; \\citet{romero}). We hope, in near future, the application of the present scenario to cosmologically distant (super GZK) AGNs, to solidify the point source scenario for dominant UHECR production, which is responsible for the suppression of cosmic-ray flux above $4\\times 10^{19}~{\\rm eV}$ that has recently been confirmed by the Auger's experiments \\citep{abraham08}, and also, is of interest in conjunction with the detection of super GZK neutrinos \\citep[e.g.,][]{ringwald}. I am grateful to Y.~S.~Honda for a useful discussion.\\\\" }, "0911/0911.2115_arXiv.txt": { "abstract": "We present new and archival multi-frequency radio and X-ray data for Centaurus A obtained over almost 20 years at the VLA and with {\\it Chandra}, with which we measure the X-ray and radio spectral indices of jet knots, flux density variations in the jet knots, polarization variations, and proper motions. We compare the observed properties with current knot formation models and particle acceleration mechanisms. We rule out impulsive particle acceleration as a formation mechanism for all of the knots as we detect the same population of knots in all of the observations and we find no evidence of extreme variability in the X-ray knots. We find the most likely mechanism for all the stationary knots is a collision resulting in a local shock followed by a steady state of prolonged, stable particle acceleration and X-ray synchrotron emission. In this scenario, the X-ray-only knots have radio counterparts that are too faint to be detected, while the radio-only knots are due to weak shocks where no particles are accelerated to X-ray emitting energies. Although the base knots are prime candidates for reconfinement shocks, the presence of a moving knot in this vicinity and the fact that there are two base knots are hard to explain in this model. We detect apparent motion in three knots; however, their velocities and locations provide no conclusive evidence for or against a faster moving `spine' within the jet. The radio-only knots, both stationary and moving, may be due to compression of the fluid. ", "introduction": "\\label{introjlg} It is generally agreed that the observed emission from Fanaroff-Riley class I \\citep[FR\\,I;][]{fr74jlg} radio jets is due to the synchrotron process at all wavelengths, with similar jet structure observed from the radio through the optical into the X-ray \\citep[e.g.][]{hardcastle02jlg,harris02jlg}. The jets of FR\\,I radio galaxies are thought to decelerate as they move away from the core, entraining material and expanding into a plume of diffuse matter \\citep[e.g.][]{bicknell84jlg}. One of the most significant implications of the synchrotron emission model is reflected in the characteristic loss timescales: in a stable environment, the X-ray emitting electrons have lifetimes of the order of tens of years, tracing regions of current, in situ particle acceleration, while the radio emitting electrons last for hundreds of thousands of years, showing the history of particle acceleration in the jet. In order to investigate these regions of particle acceleration we need data with sensitivity and resolution sufficient to detect jet substructure on spatial scales comparable to the synchrotron loss scales. This prompts us to look to the two closest bright FR\\,I radio jets: M87 and Centaurus A (NGC\\,5128, hereafter Cen A). Both of these jets have been detected in multiple frequencies from the radio through to the X-ray \\citep[e.g.][]{feigelson81jlg,kraft02jlg,hardcastle03jlg,hardcastle06jlg,harris02jlg}. The proximity of these radio galaxies, 16.7\\,Mpc and 3.7\\,Mpc respectively \\citep{blakeslee09jlg,mei07jlg,ferrarese07jlg}, make them unique jet laboratories with spatial scales of 77\\,pc and 17\\,pc per arcsec respectively. The details revealed in the structure of these jets have been the focus of many recent studies \\citep[e.g.][]{biretta99jlg,hardcastle03jlg,kataoka06jlg,cheung07jlg}. Within the smooth surface brightness observed in both of these jets are clumps of bright material -- the knots -- embedded in diffuse material, all emitting via synchrotron emission. The precise mechanisms causing the particle acceleration responsible for the diffuse structure and the knots are still unknown. The most surprising result of recent observations of these two systems was the radio-to-X-ray synchrotron flare of HST-1 in M87. In 2002, the X-ray flux of HST-1 increased by a factor of 2 in only 116 days \\citep{harris03jlg}, implying a change within an emitting volume with a characteristic size less than 0.1\\,pc for a stationary source (much less than the size of HST-1, $\\sim 3$\\,pc). The X-ray brightness then faded in the following months only to flare again, peaking in 2005. At its brightest, this flare was higher than its 2001 level by a factor of $\\sim50$. The UV and radio light-curves were found to vary in step with the X-ray up to this peak \\citep{perlman03jlg,harris06jlg}, but the subsequent decrease appeared to drop off faster in the X-ray than in either the optical or the UV, which drop off in step \\citep{harris09jlg}. In additional to this spectral variability, it was established with the {\\it Hubble Space Telescope (HST)} and the NRAO Very Long Baseline Array (VLBA) that some of the knots in M87 move superluminally, including subregions of the HST-1 knot \\citep{biretta99jlg,cheung07jlg}. Together, this suggests that we are observing synchrotron losses in addition to either beaming or compression/rarefaction of the fluid. Cen A is a factor of 4.5 closer than M87 so we can resolve more details in the complicated fine structure of the jet. The originally identified features, named A-G by \\citet{feigelson81jlg}, have since been resolved into at least 40 individual knots \\citep{kraft02jlg,hardcastle03jlg} with additional emission from diffuse material. Some of the diffuse emission has been described as downstream `tails' of emission from the knots \\citep{hardcastle03jlg} or as evidence for limb-brightening of the jet \\citep{kraft00jlg}. In 2003, Hardcastle et al. presented 8.4\\,GHz radio observations from the NRAO Very Large Array (VLA) of Cen A. That work used archival data from 1991 and new observations from 2002 to study the jet knots and investigated the offsets and relationships between the radio knots and their X-ray counterparts and vice versa. They found that only some of the radio knots appeared to have X-ray counterparts, leaving many as `radio-only' knots and `X-ray-only' knots. They also considered the temporal changes in the radio knots, specifically their proper motions, finding that some of the radio knots were moving. These moving knots had comparatively little X-ray emission suggesting that high-energy particle acceleration is less efficient in these regions than in the jet as a whole. Some of the current models explaining the presence of knots within the generally smooth diffuse material of the jet include compressions in the fluid flow, collisions with obstacles in the galaxy causing local shocks, reconfinement of the jet or some other jet-wide process, and magnetic reconnection. \\citet{hardcastle03jlg} ruled out simple compression of the fluid as a mechanism for producing X-ray bright, radio faint compact knots in favor of in situ particle acceleration associated with local shocks; however, compression could still play a part in the other knots. They concluded that the most likely model to describe the majority of these knots is an interaction between the jet fluid and an obstacle such as a molecular cloud or a high mass-loss star. By exploring the temporal behavior of the X-ray and radio emission, we can understand the evolution of the knots and constrain the various models of particle acceleration used to describe the jet features. In this work we use Chandra and VLA data spread over almost 20 years to measure the X-ray and radio spectral indices knots, flux density variations, polarization variations, and the proper motions of the jet knots in Cen A. Our aims are to detect variability in the radio and X-ray properties of the knots, either extreme variability similar to that of HST-1 in M87 or more subtle changes, and to compare these properties. which will allow us to constrain the knot formation processes at work in the jet of Cen A. The details of our radio and X-ray data reduction are discussed in Section~\\ref{sec:data}. In Section~\\ref{sec:res} we discuss the details of our analysis methods and the global results for the knot population, highlighting particularly interesting features. In Section~\\ref{sec:dis}, we compare these knot properties with the predictions of various models for the formation of knots, their particle acceleration and the jet structure. Finally we outline the most likely processes for forming knots in Cen A in Section~\\ref{sec:con}. ", "conclusions": "\\label{sec:con} Our results can be summarized as follows: \\begin{itemize} \\item We rule out impulsive particle acceleration in the knots of Cen A as we detect no extreme variability in the X-ray knots, in contrast to what is seen in knot HST-1 in M87. We see essentially the same distribution of X-ray knots in our most recent observation as was seen in the earliest {\\it Chandra} observations in 1999. This would not be the case if the knots were impulsive as they would fade due to synchrotron losses indicating long-lived particle acceleration in the knots of Cen A. \\item For those radio knots with X-ray counterparts, the most likely formation mechanism is a collision between the jet and an obstacle, resulting in a local shock. We see no significant variability in many of these knots, suggesting a long-lived, stable stage of particle acceleration during the interaction between the jet and the obstacle. \\item The formation of knots at the point where the inner hundred-parsec-scale jet broadens abruptly suggests that these base knots (A1A and A1C) may be reconfinement shocks; however, this is complicated by the presence of a radio-only knot (A1B) moving downstream between the possible confinement-shock knots. \\item We detect a factor of 3 increase in radio flux density of the counterjet knot SJ1. This knot lies only 17\\,pc from the nucleus so is unresolved in the X-ray; however, it was still increasing in flux in the most recent observation (Dec 2008) so we plan to continue to monitor its radio behavior with the VLA. \\item We detect proper motions in three of our radio knots; two of which have no compact X-ray counterparts and a third which has only diffuse X-ray emission associated with it. Studies of the distribution of the moving knots are inconclusive due to the low number of well-established proper motions; however, that the direction of motion of the knots may not be directly parallel to the jet axis which appears to varies along the jet. Their motions are all downstream and they show no dependence on the position of the knot within the jet. \\end{itemize} The most likely cause of knots in the jet is collisions; if the X-ray-only knots have faint radio counterparts and the radio-only knots are seen only during the latter stages of the collision when the interaction is weaker, then only the moving knots are a separate population. These may include all the radio-only knots but our proper motion measurements are inconclusive for many of these. We argue that these moving knots are due to compressions in the fluid flow that do not result in particle acceleration to X-ray emitting energies. It is possible, however, that the X-ray-only knots are also a separate population with flatter X-ray to radio spectra than those with counterparts, in which case we currently have no model for their formation. Compared to other FR\\,I jets, Cen A is atypical, with an obscuring dust lane extending out to 1\\,kpc from the core which greatly affect the jet and its knots. Other galaxies where dust has been detected, such as 3C31 and 3C449, have much smaller disks, which cannot affect even the innermost regions of the observed X-ray jet. If we can attribute the knot dominated particle acceleration of the inner kpc to the presence of this disk then we can postulate that the X-ray jet emission seen in other FR\\,I galaxies should be comparable to the dominant diffuse particle acceleration that dominates farther out in the Cen A jet. We would then predict that knot-dominated structure will not be seen in other FR\\,I galaxies. \\vspace{-10pt}" }, "0911/0911.0326_arXiv.txt": { "abstract": "The study of scattering and extinction properties of possible nanodiamond grains in the ISM are reported. Calculations using Discrete Dipole Approximation (DDA) for varying ellipsoidal shapes and sizes from $2.5$ to $10 ~nm$ are considered. Nanodiamonds show negligible extinction from IR to near-UV and very sharp far-UV rise. Comparison with observations rule out possibility of independent nanodiamond dust but point towards possibility of nanodiamonds as a component in the ISM. Radiation induced transformations may lead to carbonaceous grains with different core and mantles. So calculations are also performed for a core-mantle target model with nanodiamond core in graphite mantles. The graphite extinction features get modified with the peak at 2175 \\AA{} being lowered, broadened, blue shifted and accompanied by enhanced extinction in the far-UV. Such variations in the 2175 \\AA{} band and simultaneous far-UV rise are observed along some sources. A three component dust model incorporating silicate, graphite and graphite with nanodiamond core is also considered. The model extinction compares very well with the average galactic extinction in the complete range from $0.2$ to $10 ~\\mu m^{-1}$. The best fit requires small size and small number of nanodiamonds. ", "introduction": "To account for the ultraviolet extinction in the diffuse interstellar medium \\citet{saslaw69} first proposed the possibility of diamonds in Interstellar Medium (ISM). The 3.43 and 3.53 $\\mu m $ emission bands in circumstellar media of Ae/Be Herbig stars HD~97048 and Elias~1 show convincing presence of nanodiamonds \\citep{guillois99}. % \\citet{kerckhoven02} attribute these bands to C-H stretching in hydrogenated nanodiamond, as distinct from the 3.3 $\\mu$m PAH feature. The 3.47 $\\mu m $ feature in absorption toward a number of proto-stars is attributed to the tertiary C-H stretching mode in diamond like structures \\citep{allamandola92}. Spectra of diamondoid molecules \\citep{pirali07} show that nanodiamonds a few nanometre in size could be responsible for the 3.53 and 3.43 $\\mu$m emission lines and smaller diamondoids give the 3.47 $\\mu$m absorption feature. Cosmic nanodiamonds are also detected in primitive carbonaceous meteorites of presolar origin \\citep{lewis87, daulton96, dai02, garai06}. In fact, nanodiamonds are considered to be the most abundant presolar component in meteorites \\citep{anders93, zinner98}. \\citet{jones04} studied $C-H$ stretching mode of nanodiamonds extracted from Orgueil meteorite and observed the above mentioned IR bands. If the ISM nanodiamonds are not hydrogenated their detection is hard \\citep{kerckhoven02} and they could be more abundant than estimated by the observations of infrared $C-H$ bands \\citep{dedeigo07}. It is, therefore, important to incorporate nanodiamonds in UV extinction models for which their independent scattering and extinction properties need to be understood. The extinction properties of nanodiamond dust have been reported \\citep{mutschke04} for spherical shape and the far-UV break of spectral energy in quasars is attributed to nanodiamond dust \\citep{binette05, binette06}. In this communication extinction properties of nanodiamond grains of different ellipsoidal shapes and sizes are reported. To study the effect of nanodiamond within graphite extinction efficiencies are also reported for nanodiamond-graphite as core-mantle grains. An extinction curve modeling is proposed including nanodiamond as a component. ", "conclusions": "Various experimental, theoretical and observational studies suggest the possibility of carbon in diamond form in the ISM. Besides enhanced far-UV extinction \\citep{binette05, binette06}, luminescence from diamond nano-crystals could be responsible for the Extended Red Emission (ERE) \\citep{chang06}. The study of extinction properties of possible nanodiamond grains of different shapes and sizes is done to understand their role in far-UV extinction and in modification of the overall extinction curve. The extinction due to nanodiamonds of all shapes and sizes is essentially similar with negligible extinction from IR to near UV range and sharp rise in extinction in the far-UV. The extinction due to non-spherical shape is higher than that due to spherical. For nanosized particles the extinction efficiency increases linearly with effective radius. Considering nanodiamond-graphite core-mantle target, modification in the 2175 \\AA{} peak and graphite extinction is studied. In general the 2175 \\AA{} peak gets lowered, broadened and there is enhanced far-UV extinction. Increase in \\% of $ sp^3 $ character in graphite gives rise to higher far-UV extinction. The three component modeling of average galactic extinction gives best results for small sized and low \\% of nanodiamond. This enables us to explain far-UV rise and simultaneous modification of the 2175 \\AA{} bump without putting constraints on the interstellar abundances. Extinction from specific objects that show enhanced far-UV rise can be attempted by incorporating a nanodiamond component. This will enhance understanding of radiation induced transformations of carbonaceous matter in the ISM." }, "0911/0911.3148_arXiv.txt": { "abstract": "Radial velocities measured from near-infrared spectra are a potentially powerful tool to search for planets around cool stars and sub-stellar objects. However, no technique currently exists that yields near-infrared radial velocity precision comparable to that routinely obtained in the visible. We are carrying out a near-infrared radial velocity planet search program targeting a sample of the lowest-mass M dwarfs using the CRIRES instrument on the VLT. In this first paper in a planned series about the project, we describe a method for measuring high-precision relative radial velocities of these stars from $K$-band spectra. The method makes use of a glass cell filled with ammonia gas to calibrate the spectrograph response similar to the ``iodine cell'' technique that has been used very successfully in the visible. Stellar spectra are obtained through the ammonia cell and modeled as the product of a Doppler-shifted template spectrum of the object and a spectrum of the cell, convolved with a variable instrumental profile model. A complicating factor is that a significant number of telluric absorption lines are present in the spectral regions containing useful stellar and ammonia lines. The telluric lines are modeled simultaneously as well using spectrum synthesis with a time-resolved model of the atmosphere over the observatory. The free parameters in the complete model are the wavelength scale of the spectrum, the instrumental profile, adjustments to the water and methane abundances in the atmospheric model, telluric spectrum Doppler shift, and stellar Doppler shift. Tests of the method based on the analysis of hundreds of spectra obtained for late M dwarfs over six months demonstrate that precisions of $\\sim$\\,5\\,m\\,s$^{-1}$ are obtainable over long timescales, and precisions of better than 3\\,m\\,s$^{-1}$ can be obtained over timescales up to a week. The obtained precision is comparable to the predicted photon-limited errors, but primarily limited over long timescales by the imperfect modeling of the telluric lines. ", "introduction": "\\subsection{Motivation for NIR radial velocities} The vast majority of known exoplanets have been detected with high-precision (i.e. $\\sigma$ $<$ 10\\,m\\,s$^{-1}$) time-series stellar radial velocity measurements. Invariably, these measurements have been made using spectra obtained in the visible wavelength region. This is natural for two reasons. First, arguably the most interesting targets for planet searches are solar-type stars (broadly defined here as dwarfs with F, G, and K spectral types), and these stars are brightest at wavelengths shorter than 1\\,$\\mu$m. The spectra of solar-type stars at these wavelengths are also rich in deep and sharp spectral lines suitable for Doppler shift measurements. The second reason is that visible wavelength spectrograph technology is much more advanced relative to that required for spectrographs operating in other wavelength regions. When fed by a moderately sized telescope, visible wavelength spectrographs using a cross-dispersed echelle design and CCD detectors can yield high-resolution and high signal-to-noise (S/N) spectra with a large wavelength coverage for a large number of solar-type stars. The more advanced technology in the visible also includes well-established methods for high-precision calibration. The combination of currently obtainable data quality and calibration together make it possible to reach radial velocity precisions of $\\sim$\\,1\\,m\\,s$^{-1}$ for many solar-type stars, which is sufficient to detect planets down to a few Earth masses in short-period orbits around them \\citep[e.g.][]{howard09,bouchy09}. Despite the obvious utility of radial velocity measurements in the visible, measurements at other wavelengths are also desirable. The near-infrared (NIR, i.e. 1 -- 5\\,$\\mu$m) in particular is an interesting wavelength region for radial velocity exoplanet studies. In contrast to solar-type stars, low-mass objects that are cooler than $\\sim$\\,4000\\,K are brightest at wavelengths of 1\\,$\\mu$m and longer. Furthermore, stars cooler than $\\sim$\\,3200\\,K are so faint at visible wavelengths that high-precision radial velocity measurements are impossible for all but a few very nearby examples even with the best instruments on 8-10\\,m class telescopes. For example, the sample of 40 stars in a long-term radial velocity planet search program specifically targeting low-mass stars that utilized UVES \\citep{dekker00} at the VLT only included two objects with masses below 0.2\\,M$_{\\sun}$, and only three with masses below 0.35\\,M$_{\\sun}$ \\citep{zechmeister09}. Very low-mass stars are an interesting sample of potential planet hosts notwithstanding their neglected status. They are useful for probing the correlation between gas giant planet frequency and stellar mass \\citep{endl06, johnson07}, are the most numerous stars in the Galaxy \\citep{reid02, henry06, covey08}, and exhibit a larger dynamical response to orbiting planets and have closer-in habitable zones than higher-mass stars \\citep{kasting93}. The latter point means that much less precision is needed to detect potentially habitable planets around low-mass stars with the radial velocity method than is needed for solar-type stars. The ubiquity of mid to late M dwarfs and the closeness of their habitable zones suggests that the closest stars hosting transiting habitable planets are likely of this stellar type \\citep{deming09}. Furthermore, \\citet{deming09} have estimated that basic atmospheric characterization of a few such planets could be obtained with the \\textit{James Webb Space Telescope} using the techniques of transit and occultation spectroscopy. Therefore, very low-mass stars are an important sample for the possible future study of the atmosphere of a habitable planet. However, we first must be able to use the radial velocity method to identify or confirm (in the case the planet is first identified by a transit detection), and measure the masses of low-mass planets around very low-mass stars before these investigations can be carried out. As faint as the lowest-mass stars are at visible wavelengths, their extreme redness means they are reasonably bright at NIR wavelengths. For example, an M6 star (M$_{\\star}$ $\\approx$ 0.1\\,M$_{\\sun}$) at 10\\,pc will have a $V$-band magnitude of $\\sim$\\,15.5, which would make it a very challenging target for high-precision radial velocity measurements in this wavelength region with current telescopes and instruments. However, the same star will have a $K$-band magnitude of $\\sim$\\,8.5, which is bright enough that spectra with S/N suitable for high-precision radial velocity measurements could be easily obtained using existing high-resolution NIR spectrographs on 8-10\\,m class telescopes - especially those equipped with adaptive optics (AO) systems. Therefore, if the issues and advantages of low-mass stars described above are to be studied and exploited using the radial velocity method, then the NIR spectral region offers the only possible option without building larger telescopes. Radial velocities measured from NIR spectra also likely offer the advantage that activity induced ``jitter'' is reduced relative to measurements in the visible. Magnetic activity plays a role in limiting the effectiveness of radial velocity planet searches because it results in inhomogeneities on the stellar surfaces (spots and plage). Visible spectra are thus modulated by stellar rotation and variations in the activity. These intrinsic variations mimic real radial velocity changes and are very confusing for planet searches \\citep[e.g.][]{queloz01,paulson04,wright05}. It is thought that radial velocities measured from NIR spectra will exhibit reduced jitter relative to the visible because of the smaller contrast between the cool inhomogeneities caused by activity and the average stellar surface at the longer wavelengths. Significant evidence exists to support the existence of this effect, and this evidence is discussed in the following sub-section. The issue of activity induced radial velocity jitter is particularly relevant for low-mass stars because it is well established that a much higher fraction of late-type stars are active than earlier-type stars. In the most comprehensive survey so far, \\citet{west04} found that the fraction of active stars increased from $<$ 10\\% at spectral type M3, to 50\\% at M5, and ultimately 75\\% at the end of the main sequence. As high as these activity fractions are, they are likely only lower limits for nearby stars because the \\citet{west04} survey included stars orbiting perpendicular to the Galactic plane, and such stars are potentially very old. Therefore, the fraction of active stars in a radial velocity planet search of M dwarfs will likely be even higher than the \\citet{west04} statistics suggest because they will be drawn from an intrinsically younger sample. This has serious implications for planet searches because it means that the majority of the lowest-mass stars will exhibit substantial activity induced jitter in radial velocities measured from visible wavelength spectra even if their faintness can be overcome. \\subsection{Previous work in the area} The advantages offered by NIR radial velocities have been identified before and there has been some previous work to realize them. \\citet{martin06} used NIRSPEC \\citep{mclean98} on Keck to make NIR radial velocity measurements of the brown dwarf LP\\,944-20. Their measurements over six nights had an rms dispersion of 360\\,m\\,s$^{-1}$. This result was inconsistent with previously measured visible wavelength radial velocities, which exhibited a coherent periodic signal with an amplitude of 3.5\\,km\\,s$^{-1}$. This result is generally interpreted as confirmation that radial velocity jitter is indeed reduced at NIR wavelengths for active cool dwarfs. However, it should be noted that the NIR and visible data sets were not contemporaneous, and the possibility that the star was experiencing a period of reduced activity during the NIR measurements can not be ruled out. \\citet{zapatero09} have also reported radial velocity measurements obtained from NIRSPEC data, in this case for the very low-mass star VB\\,10. Their data were obtained over seven years and have an estimated precision of $\\sim$\\,300\\,m\\,s$^{-1}$. These measurements seem to support the astrometric detection of a giant planet around VB\\,10 claimed by \\citet{pravdo09}. However, our measurements of VB\\,10, which are presented in paper II and exhibit a dispersion of only 10\\,m\\,s$^{-1}$, are inconsistent with the \\citet{zapatero09} measurements and also appear to completely rule out the existence of the proposed giant planet around this star \\citep{bean10}. In addition, \\citet{anglada10} did not detect the expected reflex motion of VB\\,10 in their visible wavelength radial velocities, which have a typical precision $\\sim$\\,200\\,m\\,s$^{-1}$. \\citet{blake07} reported NIR radial velocity measurements with PHOENIX \\citep{hinkle03} on Gemini South. They targeted L dwarfs and reached a precision of 300\\,m\\,s$^{-1}$ over five nights. \\citet{prato08} observed a sample of young stars and some late-type radial velocity standard stars with CSHELL \\citep{greene93} on the IRTF for seven nights. They reached precisions of $\\sim$\\,100\\,m\\,s$^{-1}$ for the standard stars, and 100 -- 300\\,m\\,s$^{-1}$ for the young stars. The measurements for the young stars again indicated no coherence with significant periodicities seen in visible radial velocities, but also were not contemporaneous. Similarly, \\citet{huelamo08} monitored the young star TW Hya over six nights using CRIRES \\citep{kaeufl04} on the VLT. Their measurements yielded an rms dispersion of 35\\,m\\,s$^{-1}$, which was inconsistent with a coherent signal having an amplitude of $\\sim$\\,300\\,m\\,s$^{-1}$ seen in visible radial velocities obtained previously \\citep{setiawan08} and contemporaneously. \\citet{seifahrt08} also used CRIRES to obtain continuous observations of an M giant over five hours and reached a precision of $\\sim$\\,20\\,m\\,s$^{-1}$. In addition to these stellar observations, there has been some work on measuring NIR radial velocities for the Sun. \\citet{deming87} and \\citet{deming94} presented data from a long-term monitoring program using a Fourier Transform Spectrometer (FTS). Their $K$-band measurements had a precision of better than 5\\,m\\,s$^{-1}$. More recently, \\citet{ramsey08} used a laboratory grating spectrograph to measure the signature of the Earth's rotational velocity in $Y$-band spectra of integrated Sunlight. They maintained a precision of $\\sim$\\,10\\,m\\,s$^{-1}$ over a few hours in two separate experiments. \\subsection{Challenges of NIR radial velocities} Despite the attention paid to obtaining NIR radial velocities of cool stars, no previous work has achieved a long-term precision on a star other than the Sun within an order of magnitude of the precision that is routinely obtained in the visible. One reason for this is simply that enough signal was not obtained to probe the true limit in some cases. However, based on photon statistics, the observations of \\citet{prato08} and \\citet{huelamo08} should have at least achieved precisions around a factor of two better than they did. Another issue is that no studies with an external check on the obtained precision have been reported continuing for longer than seven days. Therefore, the obtainable long-term precision in NIR radial velocities has not been systematically developed or evaluated. The main issue that has limited the pursuit of high-precision NIR radial velocities is the lack of a suitable calibration method. For example, useful lines from available emission lamps \\citep[e.g. ThAR,][]{lovis07} are infrequent in the NIR relative to the visible. This is a problem for the existing high-resolution NIR spectrographs because they generally have relatively short wavelength coverage for single exposures. Laser frequency combs have the potential to provide calibration lines in nearly any spectral region conceivable for stellar radial velocity work \\citep{murphy07,li08,steinmetz08}. However, the technique is still not developed enough for long-term use at an astronomical observatory. Furthermore, none of the existing NIR spectrographs can be fed simultaneously with light from a calibration source like a lamp or laser comb system, and are not stabilized to the level necessary to do without simultaneous calibration. In addition, there is also the issue that current NIR spectrographs experience significant variations in illumination on short timescales. This means that even if these instruments could be simultaneously fed with light from a lamp, the calibration would not track all the necessary effects for high-precision radial velocity measurements. Therefore, emission lamps and laser combs are not useful for high-precision calibration in currently available NIR spectrographs, but could be for future instruments designed for larger wavelength coverage and excellent stability \\citep{reiners10}. All of the previous studies discussed above that targeted stars other than the Sun utilized the telluric spectrum imprinted on the data for \\textit{in situ} calibration, although the lines and methods utilized varied somewhat. In their study, \\citet{seifahrt08} demonstrated that telluric N$_{2}$O lines near 4.1\\,$\\mu$m were stable to a level of 10\\,m\\,s$^{-1}$ over five hours. On the other hand, \\citet{deming87} showed that the telluric methane lines in the $K$-band varied with a semi-amplitude of 20\\,m\\,s$^{-1}$ depending on the hour angle due to the prevailing winds over the observatory. In addition to the possible variability of the telluric lines, another issue with using telluric lines as a radial velocity fiducial is that measurements obtained over a full observing season could be affected by systematics because the stellar lines will move relative to the telluric lines due to the changing barycentric velocity of the observatory along the line of sight. This will lead to more or less blending of features, which could influence the measurements. The well known ``iodine cell'' method \\citep{butler96} has been used to measure radial velocities to precisions of a few m\\,s$^{-1}$ with numerous visible spectrographs that are not highly stabilized. The general gas cell technique involves placing a cell in front of the spectrograph during observations so that a standard spectrum is imprinted on each obtained stellar spectrum. The cell's spectrum recorded during each exposure can then be used to precisely determine the spectrograph response at the exact moment of the observations and, thus, calibrate the simultaneously obtained stellar spectra for radial velocity measurements. The success of the iodine cell in the visible implies that a gas cell could be a useful way to calibrate currently available NIR spectrographs. The main constraint for the gas cell method is that the gas in question should exhibit numerous sharp and deep lines in a wavelength interval where the stars targeted for radial velocity measurements have lines as well \\citep{campbell79}. Ideally, this region would also be free of atmospheric lines. However, this is a particularly challenging requirement in the NIR due to the prevalence of strong atmospheric lines even in the traditional windows between water absorption bands. \\citet{deming87} and \\citet{deming94} used a cell filled with N$_{2}$O to calibrate their FTS measurements of integrated sunlight from 2.0 -- 2.5\\,$\\mu$m. In addition, \\citet{mahadevan09} considered a number of options for NIR gas cells and concluded that the gases H$^{13}$C$^{14}$N, $^{12}$C$_{2}$H$_{2}$, $^{12}$CO, and $^{13}$CO together could provide useful calibration in the $H$-band. However, there have been no reported uses of a gas cell for high-precision NIR radial velocity measurements for stars other than the Sun, and no such measurements obtained using a grating spectrograph. We have recently developed a new gas cell for simultaneous calibration of spectra obtained in the $K$-band, and are currently using it to conduct a radial velocity search for planets around very low-mass stars with the CRIRES instrument at the VLT. In this paper, which is the first in a planned series about our planet search, we describe a method for measuring high-precision relative radial velocities of cool stars using observations made with the cell, and we present extensive data sets demonstrating the performance of the method over long timescales. The paper is laid out as follows. In \\S2 we describe the new cell and its implementation in CRIRES. We describe the typical observing procedure and data reduction used in \\S3. We detail the radial velocity measurement algorithm in \\S4. In \\S5, we present tests of the obtained radial velocity precision. We conclude in \\S6 with a discussion and outlook for future planet searches utilizing NIR radial velocities. ", "conclusions": "The general technique of measuring radial velocities from NIR spectra has recently been receiving a lot of attention as a possible important direction for the observational study of exoplanets for the reasons described in \\S1.1 \\citep[e.g.][]{lunine08}. However, up to now the feasibility of obtaining NIR radial velocity precision similar to that routinely obtained in the visible was largely unknown due to uncertainty about a number of technical issues. The main open question has been how to calibrate NIR spectra, with a second issue being whether the quality of NIR detectors is sufficient. We have described a new gas cell suitable for calibrating spectra of cool dwarfs in the $K$-band, and demonstrated long-term radial velocity precisions of $\\sim$\\,5\\,m\\,s$^{-1}$ from data obtained with CRIRES at the VLT when using the cell. It is worth noting that the CRIRES detectors have severe cosmetic blemishes relative to CCDs. However, the effects seem to be stable and we have been able to calibrate them out to an acceptable level. Furthermore, the detectors were not even the highest-quality NIR detectors that were available at the time CRIRES was built \\citep[ca. 2002,][]{dorn04}, and are certainly not competitive with the best detectors available as of this writing. Therefore, the issue of the quality of NIR detectors requires attention, but will not likely be a limiting factor for new instruments aiming at $\\sim$\\,1\\,m\\,s$^{-1}$ NIR radial velocity precision. Our ammonia gas cell has proven to be useful for NIR calibration down to a few m\\,s$^{-1}$ precision despite not being temperature stabilized. Therefore, it could be useful for high-precision radial velocity measurements with other instruments not originally designed for excellent stability. The precision afforded by the cell and the measurement method we have described is already an order of magnitude better than the best previously obtained long-term NIR precision on a star other than the Sun. As a result, it now enables the possibility of searching for planets around a much larger number of very low-mass stars than was feasible before. For example, our planet search project includes 31 objects with estimated masses below 0.2\\,M$_{\\odot}$, 22 of which have estimated masses below 0.15\\,M$_{\\odot}$. A previous high-precision radial velocity planet search utilizing a visible wavelength spectrograph on an identical telescope (UVES on the UT2 telescope of the VLT) was only able to target two objects with estimated masses below 0.2\\,M$_{\\odot}$, and only one of these two has an estimated mass below 0.15\\,M$_{\\odot}$ \\citep{zechmeister09}. In addition to enabling the search for planets around more low-mass stars, the gas cell and radial velocity measurement algorithm we have developed also opens up a new frontier on the search for potentially habitable planets. The orbital period range for a planet in the habitable zone around a star with a mass M = 0.10\\,M$_{\\odot}$ would be 3 -- 21 days \\citep{selsis07}. The 5\\,m\\,s$^{-1}$ precision obtainable with our method corresponds to the velocity semi-amplitudes induced by a 2.5\\,M$_{\\oplus}$ planet or a 4.6\\,M$_{\\oplus}$ planet on the inner and outer edges of the habitable zone respectively. Therefore, it should be possible to detect Super-Earth type planets in the habitable zones of very low-mass stars with a reasonable expenditure of observing time on current facilities. Altogether our results have shown that obtaining high-precision NIR radial velocities is possible, and we see no technical reason why a level of precision of 1\\,m\\,s$^{-1}$ could not be achievable with a highly stabilized NIR grating spectrograph. However, the design of such a next generation instrument will have to carefully weigh the availability of calibration against the availability of information (i.e. spectral lines) from the stars being targeted like we have done with our observing program. As discussed above, we are carrying out a planet search using CRIRES with the ammonia cell and this paper is the first in a planned series from the project. The next papers in the series will present high-precision NIR radial velocities of our M dwarf planet search targets. The data will allow us to detect planets around these stars with similar masses and orbital separations as those that can be detected around solar-type stars using state-of-the-art visible wavelength radial velocities." }, "0911/0911.4768_arXiv.txt": { "abstract": "We study the halo bispectrum from non-Gaussian initial conditions. Based on a set of large $N$-body simulations starting from initial density fields with local type non-Gaussianity, we find that the halo bispectrum exhibits a strong dependence on the shape and scale of Fourier space triangles near squeezed configurations at large scales. The amplitude of the halo bispectrum roughly scales as $\\fnl^2$. The resultant scaling on the triangular shape is consistent with that predicted by Jeong \\& Komatsu based on perturbation theory. We systematically investigate this dependence with varying redshifts and halo mass thresholds. It is shown that the $\\fnl$ dependence of the halo bispectrum is stronger for more massive haloes at higher redshifts. This feature can be a useful discriminator of inflation scenarios in future deep and wide galaxy redshift surveys. ", "introduction": "The standard cosmological model has successfully explained the observed statistical properties of the cosmic microwave background (CMB) radiation and the large scale structure (LSS) traced by galaxies (e.g., \\cite{WMAP5,Tegmark2006}). The model usually assumes that the primordial density/temperature/curvature fluctuations follow Gaussian statistics. Recently, however, possible deviations from standard Gaussian statistics has attracted great attention with rapid progress in observational techniques. It offers an opportunity to access cosmological information beyond traditional power spectrum analysis. Many recent works have discussed the constraints and future detectability of possible deviations from Gaussianity through the observations of CMB and LSS (e.g., \\cite{Komatsu2001,Carbone2008}). According to the inflationary scenarios, primordial curvature perturbations are generated during the accelerated phase of cosmic expansion. The simplest inflation models, in which the inflation takes place with the slow-roll single scalar field that has a canonical kinetic structure, generally predicts a nearly scale-invariant spectrum of curvature perturbations, and a small departure from Gaussianity. On the other hand, a variety of inflation models that produce a large non-Gaussianity has been recently proposed (see \\cite{Bartolo2004} for a review). Among these, the models with large non-Gaussianity generated by non-linear dynamics on super-horizon scales can have a generic prediction for non-Gaussian properties. Denoting a Gaussian field by $\\Phi_{\\rm G}(\\bfx)$, the Bardeen's curvature perturbation during the matter era is characterized as \\cite{Komatsu2001}: \\begin{eqnarray} \\Phi(\\bfx) = \\Phi_{\\rm G}(\\bfx)+\\fnl\\left\\{ \\Phi_{\\rm G}^2(\\bfx)-\\langle\\Phi_{\\rm G}^2(\\bfx)\\rangle\\right\\} +\\cdots. \\label{eq:fnllocal} \\end{eqnarray} This type of non-Gaussianity, described as a local function of the Gaussian field, is often called {\\it local type} non-Gaussianity, and the leading-order coefficient $\\fnl$, which controls the strength of non-Gaussianity, has information on the generation mechanisms for non-Gaussian fluctuations. Although the current constraint on the parameter $\\fnl$ from CMB data is $-9<\\fnl<111$ ($95\\%$C.L.) \\cite{WMAP5} and it is still consistent with Gaussian ($\\fnl=0$), the constraint will be tightened by the on-going CMB experiment, Planck \\cite{Planck}. As standard inflation models generally predict $|\\fnl|\\ll1$, detection of large non-Gaussianity immediately implies the non-standard mechanism for generation of primordial curvature perturbations. In this paper, we focus on how this type of non-Gaussianity alters the statistical properties of LSS. Let us first define the power spectrum of mass density fluctuations assuming isotropy and homogeneity: \\begin{eqnarray} \\langle \\deltam(\\bfk_1)\\deltam(\\bfk_2)\\rangle \\equiv (2\\pi)^3\\delta_{\\rm D}(\\bfk_1+\\bfk_2)\\Pm(k_1), \\label{eq:pow_def} \\end{eqnarray} where $\\deltam(\\bfk)$ denotes the Fourier transform of the density contrast, while $\\delta_{\\rm D}(\\bfk)$ represents the Dirac delta function. If the density field follows Gaussian statistics, its power spectrum determines all the statistical properties. Next we define the bispectrum of a mass density field: \\begin{eqnarray} \\langle \\deltam(\\bfk_1)\\deltam(\\bfk_2)\\deltam(\\bfk_3)\\rangle \\equiv (2\\pi)^3\\delta_{\\rm D} (\\bfk_1+\\bfk_2+\\bfk_3)\\Bm(\\bfk_1,\\bfk_2,\\bfk_3). \\label{eq:bis_def} \\end{eqnarray} Since this is the lowest-order non-vanishing quantity in the presence of non-Gaussianity, the bispectrum is naively expected as the best measure for non-Gaussianity (e.g., \\cite{Scoccimarro2000,Verde2000,Scoccimarro2004,Sefusatti2007}). Recently, however, the {\\it galaxy} or {\\it halo} power spectrum has been reconsidered in the presence of local and/or equilateral type primordial non-Gaussianity both analytically and numerically (e.g., \\cite{Dalal2008,Afshordi2008,Taruya2008,Matarrese2008,Pillepich2008,Desjacques2009a,Desjacques2009b,Grossi2009,Slosar2008,McDonald2008,Sefusatti2009,Giannantonio2009}). The matter power spectrum or bispectrum is not a direct observable, and the real measurement of LSS gives {\\it galaxy} power spectrum or bispectrum, as defined similarly to equations (\\ref{eq:pow_def}) and (\\ref{eq:bis_def}). Since galaxies are biased tracers of the dark matter distribution, the information of $\\fnl$ is imprinted in a different manner: a new contribution coming from the primordial non-Gaussianity may dominate over the Gaussian term in the galaxy power spectrum, $\\Pg(k)$, at very large scales ($k\\simlt0.01h$Mpc$^{-1}$). This new contribution, which is sometimes referred to as the ``scale-dependent bias'' of the galaxy power spectrum, may be a powerful indicator to constrain $\\fnl$. Indeed, it has been recently applied to the clustering statistics of SDSS LRG and quasar samples, and the tight constraints on $\\fnl$ are comparable to those obtained from CMB measurements have been obtained \\cite{Slosar2008}. The purpose of this paper is to examine the bispectrum of biased tracers in detail. While the matter bispectrum in the presence of primordial non-Gaussianity has been studied in the literature using both perturbation theory and numerical simulations, the galaxy bispectrum may significantly differ from the matter bispectrum in the presence of primordial non-Gaussianity, just like the difference in the power spectra. Since the local type non-Gaussianity can be straightforwardly implemented within $N$-body simulations, numerical study on the bispectrum for the dark matter haloes is the first important step toward a practical understanding of the galaxy bispectrum. Incidentally, Jeong and Komatsu (2009) recently proposed a new parametrized model for the halo/galaxy bispectrum (\\cite{Jeong2009}, see also \\cite{Sefusatti2009}) based on the peak bias model \\cite{Matarrese1986} and the local bias model \\cite{Fry1993}. They found that the formula for the galaxy bispectrum used in \\cite{Sefusatti2007} was missing important contributions from the scale dependent bias effects and they discovered new terms that are important at ``squeezed'' configurations where $k_1, k_2\\gg k_3$. It was argued that these new contributions enable us to put stronger constraints on $f_{\\rm NL}$ than those obtained in \\cite{Sefusatti2007}. It is of great importance to confirm the scale-dependent bias effects in the bispectrum by $N$-body simulations. The rest of this paper is organized as follows: we first review the analytical models of the power spectrum and the bispectrum of biased tracers in section \\ref{sec:theory}. We then describe the setup and initial conditions for $N$-body simulations in section \\ref{sec:simulation}. As a first check of our simulations, in section \\ref{sec:pow_results}, we compute the matter and halo power spectra, and the results are compared with previous works. Section \\ref{sec:bis_results} gives the main results of this paper, in which the simulation results for the matter and halo bispectra are presented and compared with predictions from analytic models, particularly focusing on their scale dependence. The dependence of the halo bispectrum on the halo mass threshold and redshift is also investigated in detail. Section \\ref{sec:discussion} discusses the future prospects for detecting the primordial non-Gaussianity using the scale-dependent properties of the halo/galaxy bispectrum. Finally, section \\ref{sec:conclusion} is devoted to conclusions and discussions. ", "conclusions": "\\label{sec:conclusion} In this paper, we have studied the clustering properties of dark matter haloes from cosmological $N$-body simulations in the presence of local-type primordial non-Gaussianity. We found that the halo bispectrum measured from $N$-body simulations exhibits a strong $f_{\\rm NL}$ dependence which becomes most prominent for squeezed configurations at large scales. In particular, for realistic values of $|\\fnl|\\simlt100$, the dependence of the halo bispectrum on $\\fnl$ is well characterized by the polynomial expansions of $\\fnl$ up to second order. Since the quadratic dependence on $\\fnl$ does not appear in the matter bispectrum at the lowest order in perturbation theory, this would be a clear indicator for the existence of primordial non-Gaussianity of the local type. We have investigated the shape and scale dependence of the halo bispectrum arising from the $\\fnl^2$ term, and the simulation results are compared with theoretical predictions based on the local bias model. For the isosceles triangles characterized by $\\alpha\\equiv k_1 / k_3$ and $k\\equiv k_1=k_2$, the dependence of the halo bispectrum on $\\alpha$ measured from $N$-body simulations is found to be consistent with theoretical predictions by \\cite{Jeong2009}. We also examined the dependence of the halo bispectrum on minimum halo mass and redshift, and showed that the amplitude of the halo bispectrum is more significant for more massive haloes at higher redshifts. Thus, the strong dependence of the halo/galaxy bispectrum on $\\alpha$ makes the detection of primordial non-Gaussianity much more promising in future surveys. As a preliminary investigation, we have evaluated the signal-to-noise ratio for the scale dependence of the bispectrum in three representative surveys, and found that even with the very limited number of configurations of bispectrum it is possible to detect primordial non-Gaussianity if $\\fnl$ is several dozen. Thus, the detectability of primordial non-Gaussianity is expected to be greatly improved if we use all configurations of the bispectrum. We leave the following tasks as a future work: (i) Study the effects of redshift-space distortions. Since we focus on very large scales, we expect that these effects are accurately described by linear theory, i.e., they just enhance the amplitude of the bispectrum in a scale independent way. (ii) Construct more elaborate theoretical models that are applicable to a wider range of triangular configurations and compare them with $N$-body simulations (iii) Run higher resolution simulations where we can populate haloes with galaxies and measure the galaxy bispectrum directly from simulations. These tasks are clearly very important to exploit future surveys. \\ack We thank T.~Sousbie, R.~Nichol, E.~Komatsu, D.~Jeong, and Y.~Suto for useful discussions and comments. T.~N. is supported by a Grant-in-Aid for Japan Society for the Promotion of Science (JSPS) Fellows (DC1: 19-7066). A.~T. is supported by a Grant-in-Aid for Scientific Research from JSPS (No. 21740168). K.~K. is supported by the European Research Council, Research Councils UK and the UK's Science \\& Technology Facilities Council (STFC). C.~S. was funded by a STFC PhD studentship. Numerical computations were in part carried out on Cray XT4 at Center for Computational Astrophysics, CfCA, of National Astronomical Observatory of Japan. We are also grateful for the computational time provided by the U.K. National Grid Service (NGS). This work was supported in part by Grant-in-Aid for Scientific Research on Priority Areas No.~467 ``Probing the Dark Energy through an Extremely Wide and Deep Survey with Subaru Telescope'', and JSPS Core-to-Core Program ``International Research Network for Dark Energy''. \\appendix" }, "0911/0911.3654_arXiv.txt": { "abstract": "\\renewcommand{\\thefootnote}{\\fnsymbol{footnote}} We present kinematic and metallicity profiles for the M\\,31 dwarf elliptical (dE) satellite galaxies NGC~147 and NGC~185. The profiles represent the most extensive spectroscopic radial coverage for any dE galaxy, extending to a projected distance of eight half-light radii ($8r_{\\rm eff} \\sim 14'$). We achieve this coverage via Keck/DEIMOS multislit spectroscopic observations of 520 and 442~member red giant branch stars in NGC~147 and NGC~185, respectively. In contrast to previous studies, we find that both dEs have significant internal rotation. We measure a maximum rotational velocity of $17\\pm 2$\\kms\\ for NGC~147 and $15\\pm5$\\kms\\ for NGC~185. While both rotation profiles suggest a flattening in the outer regions, there is no indication that we have reached the radius of maximum rotation velocity. The velocity dispersions decrease gently with radius with an average dispersion of $16\\pm 1$\\kms\\ for NGC~147 and $24\\pm1$\\kms\\ for NGC~185. The average metallicity for NGC~147 is [Fe/H] = $-1.1\\pm0.1$ and for NGC~185 is [Fe/H] = $-1.3\\pm0.1$; both dEs have internal metallicity dispersions of 0.5\\,dex, but show no evidence for a radial metallicity gradient. We construct two-integral axisymmetric dynamical models and find that the observed kinematical profiles cannot be explained without modest amounts of non-baryonic dark matter. We measure central mass-to-light ratios of $M/L_V = 4.2\\pm0.6$ and $M/L_V = 4.6\\pm0.6$ for NGC~147 and NGC~185, respectively. Both dE galaxies are consistent with being primarily flattened by their rotational motions, although some anisotropic velocity dispersion is needed to fully explain their observed shapes. The velocity profiles of all three Local Group dEs (NGC~147, NGC~185 and NGC~205) suggest that rotation is more prevalent in the dE galaxy class than previously assumed, but is often manifest only at several times the effective radius. Since all dEs outside the Local Group have been probed to only inside the effective radius, this opens the door for formation mechanisms in which dEs are transformed or stripped versions of gas-rich rotating progenitor galaxies. ", "introduction": "\\renewcommand{\\thefootnote}{\\fnsymbol{footnote}} Dwarf elliptical (dE) galaxies are characterized by low surface brightness ($\\mu_{\\rm eff,V} > 22$ mag arcsec$^{-2}$), little to no gas, and old to intermediate age stellar populations. These galaxies are spatially clustered and account for more than 75\\% of objects brighter than $M_V < -14$ in nearby galaxy clusters \\citep{ferguson91a,ferguson94a}. Since there are few, if any, isolated dE galaxies, most proposed dE galaxy formation scenarios are environmentally-driven, such as via gas stripping or gravitational harassment of gas-rich spiral or dwarf irregular galaxies \\citep[e.g.\\,][]{dekel86a, moore98a,Aguerri09a}. Recent evidence for embedded stellar disks and bars in some dEs support these scenarios \\citep{jerjen00a,geha05a,lisker06a,chilingarian07a}. Since gas-rich systems have significant rotation velocities, a key test of this hypothesis class is comparing the kinematics of dE and gas-rich galaxies. Internal kinematic measurements for dEs in the Fornax and Virgo Cluster have revealed an unexpected dichotomy: roughly half of the observed dEs have significant rotation about the major-axis, while the remaining galaxies show no detectable major-axis rotation \\citep{pedraz02a,geha03a, vanzee04a, deRijcke05a}. dE galaxies with no rotation present a challenge to any scenario in which dE progenitors are gas-rich and thus rotating \\citep{moore98a}. Ram pressure stripping alone does not affect the kinematics of the collisionless stellar component. Gravitational processes such as galaxy harassment can reduce the amount of rotational angular momentum, converting this into random velocity dispersion, however, simulations are unable to completely remove rotational support \\citep{mayer01a,mastropietro05a}. Because dEs have low surface brightnesses, integrated light (long-slit) kinematic observations of dEs in both galaxy clusters and the Local Group are currently limited to within the half-light effective radius ($r_{\\rm eff}$). Thus, it is possible that significant rotational angular momentum lies at larger distances than has so far been explored. Addressing this question requires kinematics tracers at larger radius. While the dynamics of dE globular cluster systems suggest that there may be significant rotation at large radius \\citep{beasley06a,beasley09a}, the small number of clusters limits the usefulness of this tracer. The dEs in the Local Group share the same general properties as the more numerous dEs in galaxy clusters, however, their proximity allows us to resolve individual stars \\citep{geha06a}, and thus trace kinematics profiles down to arbitrarily low surface brightness and large radius. The Local Group contains three dE galaxies, NGC~205, NGC~147 and NGC~185, which in addition to the compact elliptical M32, are the brightest satellites around the spiral galaxy M\\,31. We follow the terminology conventions of, e.g.~\\citet{bender92a}, although alternatives are in use. While the differing terminologies can be confusing, the key point is that NGC~205, NGC~147, and NGC~185 on the one hand, and M\\,32 on the other hand, should not be grouped in a single class. The former three galaxies have fundamental plane properties that indicate similarity to galaxies in the dwarf spheroidal class (dSph). By contrast, galaxies such as M\\,32 have properties that indicate similarity to (giant) elliptical galaxies. This likely indicates a difference in formation history, as discussed more fully in, e.g., \\citet{kormendy09a}. The prototype of the dE galaxy class is NGC~205 which is in very close projection to M\\,31 ($40' = 8$\\,kpc). Both photometric and kinematic evidence suggests that NGC~205 is tidally interacting with M\\,31 \\citep{choi02a,geha06a}, and may in fact be on its first orbital approach \\citep{howley08a}. In contrast, NGC~147 and NGC~185 lie at a projected distance of $7^{\\circ}$\\,($\\sim$\\,$150$\\,kpc) from M\\,31. While the metal-poor stellar halo of M\\,31 extends to this distance \\citep{guhathakurta05a,kalirai06a}, its gravitational influence of M\\,31 is not expected to influence the kinematics of these dE satellites (see \\S\\,\\ref{subsec_bound}). NGC~147 and NGC~185 are separated from each other by merely $58'$, leading to speculation that they may be a bound galaxy pair \\citep{baade44a, vandenbergh98a}. An accurate distance estimate suggests that these two galaxies are physically separated by over 60\\,kpc \\citep{mcconnachie05a} and may not be bound (however, see \\S\\,\\ref{subsec_bound}). The large observed distances between these two dEs and their parent M\\,31, combined with their relative proximity to the Sun, make NGC~147 and NGC~185 the best available targets for studying the properties of dE galaxies to large radius. NGC~147 and NGC~185 share several similar fundamental properties, such as absolute luminosity ($M_V\\sim -15.5$), half-light radius ($r_{\\rm eff} \\sim 2' = 0.3$\\,kpc), and central velocity dispersion (for exact values see Table~1). However, they differ markedly in many aspects. NGC~185 contains some gas, dust and evidence for recent star formation confined to its center \\citep{martinez-delgado99a}, while NGC~147 is devoid of gas or dust and shows no sign of recent star formation activity \\citep{young97a, sage98a}. An average metallicity of [Fe/H] = $-1.4$ is inferred for NGC~185 \\citep{martinez-delgado99a}, while a more metal rich population of [Fe/H] = $-1.0$ is inferred for NGC~147 \\citep{han97a, davidge05a, goncalves07a}. HST/WFPC2 imaging of both galaxies also implies the presence of intermediate-age stars, with NGC~147 having a more significant contribution than NGC~185 \\citep{butler05a}. Thus, while both dEs are dominated by old to intermediate-age stars, the metallicity and age mixture of these components are different in each galaxy. The first spectroscopic long-slit observations of Local Group dEs by \\citet{bender91a} suggested that while NGC~185 appeared supported entirely by random motions, NGC~147 had a significant rotational component of $6.5\\pm1.1$\\kms. Somewhat deeper observations by \\citet{simien02a} extended the kinematic profiles out $\\sim2'$ (1$r_{\\rm eff}$), but did not measure a significant rotation for either galaxy. In contrast, both studies measured a rotational component in NGC~205 (within the tidal radius), later confirmed by \\citet{geha06a} to be $11\\pm5$\\kms. Using the Simien \\& Prugniel data, \\citet{derijcke06a} constructed dynamical models for all three Local Group dEs, concluding that, within the half-light radii, all three galaxies have mass-to-light ratios above that expected from the stellar populations alone. In this paper, we present accurate radial velocities out to 8\\,$r_{\\rm eff}$ in the Local Group dE galaxies NGC~147 and NGC~185, based on Keck/DEIMOS multi-slit spectroscopy for 520 and 442~red giant branch (RGB) stars, respectively. The paper is organized as follows: in \\S\\,\\ref{sec_data} we discuss target selection for our DEIMOS slitmasks, the observing procedure and data reduction. In \\S\\,\\ref{sec_results}, we discuss the velocity, velocity dispersion and metallicity profiles. In \\S\\,\\ref{sec_model}, we construct two-integral dynamical models for these galaxies and discuss the best fit mass distribution and dark matter content. Finally, in \\S\\,\\ref{sec_disc}, we discuss the implications of these results for the formation of dE galaxies. Throughout this paper we adopt distance moduli determined by \\citet{mcconnachie05a} via the tip of the RGB method. The distance modulus for NGC~147 is $(m -M)_0 = 24.43 \\pm 0.04$ ($675\\pm 27$\\,kpc) and NGC~185 is $(m -M)_0 = 24.23 \\pm 0.03$ ($616\\pm 26$\\,kpc). This places NGC~147 and NGC~185 at projected distances of 140 and 185\\,kpc from their parent galaxy M\\,31, respectively. \\begin{figure*}[t!] \\plotone{dEfig_dss.eps} \\caption{Palomar Sky Survey images of the two Local Group dE galaxies, NGC~147 ({\\it left}) and NGC~185 ({\\it right}). Images are $30'\\times 30'$; North is up, East is to the left. The major-axis of each galaxy is indicated by the solid line. The spatial position of Keck/DEIMOS spectroscopically confirmed member stars are shown as blue circles, non-members are plotted as open squares. \\label{fig_dss}} \\end{figure*} ", "conclusions": "\\label{sec_disc} We have presented mean velocity, velocity dispersion and metallicity profiles for the Local Group dE galaxies NGC~147 and NGC~185 based on Keck/DEIMOS spectroscopic observations of 520 and 442~member RGB stars in each dE, respectively. The profiles represent the most extensive spectroscopic radial coverage for any dE galaxy, extending to a projected distance of eight half-light radii ($14' \\sim 8r_{\\rm eff}$). Contrary to previous results, we find that both dEs have significant rotation velocities which contributes to the observed flattened shapes of each galaxy. Our two-integral dynamical modeling suggests that the observed rotation velocities cannot fully explained the observed shapes and that some anisotropic velocity dispersion is required. Our modeling estimates $M/L$ ratios of $M/L_V = 4.2\\pm0.6$ and $M/L_V = 4.6\\pm0.6$ for NGC~147 and NGC~185, respectively, which is in excess of that expected from stellar populations alone. Thus, some dark matter is required to fully explain the observe kinematics. The mean velocity profiles of NGC~147 and NGC~185 suggest that rotation may be far more prevalent in dE galaxies than previously assumed. We conclude that these Local Group dEs are rotationally-supported, however, if placed at the distance of the Virgo or Fornax Clusters, integrated-light observations such as those of \\citet{geha03a} would have concluded that these dEs have little to no rotation. It is not guaranteed that the Local Group dE galaxies had similar formation mechanisms as those residing in the much denser environments of, e.g., the Virgo or Fornax clusters. However, this is certainly plausible, given that these galaxies share similar fundamental plane properties, as defined by the photometric and kinematic properties inside $r_{\\rm eff}$ (see e.g., Fig 7). Our results then indicate that cluster dEs may also have significant rotation velocities that are manifest only at larger radii than current observations allow. If true, this significantly modifies the observational constraints under which dEs can form. The observations presented here open the door for formation mechanisms in which dEs are transformed or stripped versions of gas-rich rotating progenitor galaxies." }, "0911/0911.2652_arXiv.txt": { "abstract": "An artificial neural network (ANN) is investigated as a tool for estimating rate coefficients for the collisional excitation of molecules. The performance of such a tool can be evaluated by testing it on a dataset of collisionally-induced transitions for which rate coefficients are already known: the network is trained on a subset of that dataset and tested on the remainder. Results obtained by this method are typically accurate to within a factor $\\sim 2.1$ (median value) for transitions with low excitation rates and $\\sim 1.7$ for those with medium or high excitation rates, although 4$\\%$ of the ANN outputs are discrepant by a factor of 10 more. The results suggest that ANNs will be valuable in extrapolating a dataset of collisional rate coefficients to include high-lying transitions that have not yet been calculated. For the asymmetric top molecules considered in this paper, the favored architecture is a cascade-correlation network that creates 16 hidden neurons during the course of training, with 3 input neurons to characterize the nature of the transition and one output neuron to provide the logarithm of the rate coefficient. ", "introduction": "The interpretation of astrophysical spectra rests heavily upon the availability of relevant atomic and molecular data. Of critical importance are (1) the energies of the various accessible quantum states, (2) the rates of spontaneous radiative decay between states, and (3) the state-to-state rate coefficients for excitation and deexcitation in inelastic collisions. The last of these have proven quite elusive; in many circumstances, our ability to interpret astrophysical spectra is severely limited by the availability of collisional rate coefficients. As a result of new spectroscopic facilities such as the {\\it Spitzer Space Telescope}, the {\\it Herschel Space Observatory}, and the planned Atacama Large Millimetric Array (ALMA), the need for fundamental molecular data has become all the more acute. In modeling the emission from interstellar molecular clouds, in particular, there is a critical need for rate coefficients for the rotational excitation of molecules in collisions with H$_2$. The theoretical calculation of such data is extremely time consuming, involving the computation of a potential energy surface for the system in question, followed by extensive scattering calculations. Even for those molecules for which such calculations have been carried out, the available data are often limited to collisionally-excited collisions involving a relatively small set of low-lying states. Submillimeter and infrared spectra of molecular clouds often reveal the presence of higher-lying transitions for which collisional rate coefficients have not been computed; here, some kind of extrapolation method is needed. For the case of linear molecules, the infinite order sudden (IOS) and related approximations provide an obvious method of extrapolation; for asymmetric molecules, there is no simple analogous method. Given a limited set of transitions of an asymmetric molecule for which collisional rate coefficients have been computed, previous efforts at extrapolation (e.g.\\ Neufeld \\& Melnick 1987, Faure \\& Josselin 2008) have typically adopted an expression with one or more adjustable parameters for how the collisional rate coefficient depends upon the relevant parameters for each transition (energy difference, change in rotational quantum number, etc.) The adjustable parameters are then chosen so as to optimize the fit to those transitions for which collisional rate coefficients are known, and then the expression is used to estimate the values for higher-lying transitions where calculated results are unavailable. For example, the recent study by Faure \\& Josselin (2008) considered the rate coefficients for the excitation of H$_2$O by H$_2$, which will be of crucial importance to the interpretation of many {\\it Spitzer} and {\\it Herschel} spectra. Here, Faure \\& Josselin extrapolated a set of rate coefficients obtained from quasi-classical trajectory calculations for transitions among the lowest 45 rotational states of water. Each such state is characterized by an energy, $E$, and rotational quantum numbers $J$, $K_a$ and $K_c$ that describe the total angular momentum and its projection onto the principal axes of the molecule. For transitions with $\\Delta K_a \\le 2$, $\\Delta K_c \\le 2$ and $\\Delta J \\le 2$ -- termed ``high-propensity\" transitions -- Faure \\& Josselin assumed that rate coefficients depend solely upon $\\Delta K_a$, $\\Delta K_c$ and $\\Delta J$. For all other transitions (``low propensity transitions\"), they assumed the rate coefficients to be {\\it independent} of $\\Delta K_a$, $\\Delta K_c$ and $\\Delta J$ and proportional to $\\exp (-\\Theta \\Delta E/kT)$, where $\\Theta$ is adjusted to optimize the fit to the available data. As an alternative to the methods adopted previously, I have explored the use of an artificial neural network (ANN) to accomplish the extrapolation. In contrast to previous techniques, the use of an ANN does not make any {\\it a priori} assumption about how the transitions might be categorized or how the rate coefficients might depend upon the parameters for each transition. The calculational method is discussed in \\S2 (with the more technical details given in the Appendix). The performance of the ANN in reproducing a known set of collisional rate coefficients in presented in \\S3. A discussion follows in \\S4. ", "conclusions": "1. Artificial neural networks (ANN) can be successfully trained to determine rate coefficients for the collisional excitation of astrophysical molecules. 2. The type of ANN favored here is a cascade correlation network which creates 16 hidden neurons during the course of training. There are three input neurons that inform the network of $\\Delta E$, $\\Delta J$ and $\\Delta \\tau$ for a given transition, and one output neuron that provides the logarithm of the rate coefficient for collisional de-excitation. 3. The performance of such an ANN can be evaluated by training it on a subset of the set of transitions for which collisional rate coefficients are known, and then testing it on the remainder of that set. Since the motivation for this study is to develop a method for estimating the rate coefficients for high-lying transitions for which values have not been computed, the training set involved the lowest-lying transitions among the available data. 4. The performance of the ANN increases slightly if the results from $\\sim 10$ realizations are averaged. 5. In the primary test case considered here, the excitation of ortho-H$_2$ by H$_2$ ($J=0$), the available data -- consisting of 990 rate coefficients -- can be modeled equally well provided that the network is trained on at least $20\\%$ of the available data. 6. The performance of the ANN is summarized in Table 1, which lists the typical output errors in every case considered. Here the RMS and median values of $\\rm log_{10}(ANN\\,\\,output/actual$ value) are listed for the entire test set and separately for low, medium and high values of the actual rate. Overall, for ANN with 16 hidden neurons and 3 input neurons that are trained on at least 20$\\%$ of the available data, the typical RMS value of $\\rm log_{10}(ANN\\,\\,output/actual\\,\\,value)$ is $\\sim 0.4$; for transitions with rate coefficients $\\le 10^{-13} \\rm cm^{-3}\\,s^{-1}$, the typical RMS value is $\\sim 0.5$. The corresponding median values of $\\vert \\rm log_{10}(ANN\\,\\,output/actual\\,\\,value) \\vert$ are $\\sim 0.25$ and $\\sim 0.33$. These median values imply that one-half of the ANN outputs are typically accurate to within factors of 1.8 (or $\\sim 2.1$ in the case of rate coefficients $\\le 10^{-13} \\rm cm^{-3}\\,s^{-1}$.) However, the results shown in Figure 4 indicate that some of the ANN outputs are highly discrepant (with 4$\\%$ in error by more than a factor 10). It should be noted that the errors given here apply to individual state-to-state rate coefficients. When these are used to model molecular line emissions, the errors in the resultant line strengths can be expected to be smaller than in the state-to-state rate coefficients because a given radiative transitions may be pumped by a combination of several different collisionally-induced transitions. \\vskip 0.5 true in \\centerline{\\bf Appendix: Calculational details} The calculation made use of the Fast Artificial Neural Network (FANN) Library, version 2.0.0, an open source software package developed by Steffen Nissen and Evan Nemerson and available at http://fann.sourceforge.net. The cascade training option was adopted for this problem, and was found to result in rapid training. In this algorithm (Fahlman \\& Lebiere 1990), the training starts with an empty network devoid of hidden neurons. Hidden neurons are trained and added to the network one at a time; each new hidden neuron is chosen from a selection of candidate neurons with different initial weights and one of several possible activation functions. In using the FANN library, the input neuron values were rescaled to \\re{small values (relative to unity) to ensure that the input neurons did not saturate}: energies (given as wavenumber in units of cm$^{-1}$) were therefore divided by $10^5$ and rotational quantum numbers were divided by 100. \\re{No other preprocessing was performed. With these rescalings, the input values lay in the ranges [0, 0.14], [--0.03, 0.10], and [--0.16, 0.18] respectively for $\\Delta E$, $\\Delta J$, and $\\Delta \\tau$.} The desired output neuron values were the logarithm \\re{to base 10} of the rate coefficient in units of $\\rm cm^{3}\\,s^{-1}$, \\re{and covered the range [--15.3, --9.9]} The cascade algorithm was used with all the default settings except for one: a linear function was selected as the activation function for the output layer. This choice was necessitated by the fact that the output neuron values lay outside the range of the other activation functions\\footnote{\\re{An alternative solution would have been to rescale the output neuron values.}}. The desired error was set to a very small value to ensure that the ANN made use of the maximum number of hidden neurons specified for each test. \\re{The default settings for the FANN software make six activation functions available to the algorithm, all of which are continuous functions with a domain of [$-\\infty,+\\infty$] and a range of either [0, 1] or [--1, 1]. The network is optimized using the i-RPROP algorithm of Igel \\& H\\\"usken (2000), a variant of the RPROP algorithm introduced by Riedmiller \\& Braun (1993).} Even if the settings and training set are unchanged, the results are slightly different every time the ANN is run. This is because the initial weights are randomly assigned at the start of training. I therefore averaged the output values over a set of $N_r$ runs. Figure 7 shows how the performance of the network depends upon $N_r$, for a network with 3 input neurons and 16 hidden neurons and with $N_{train}=30$. As in Figures 3 -- 5, black curves apply to the entire test set, while blue, green and red histograms apply to transitions with low, medium and high actual rate coefficients; for this purpose, ``low\" applies to value smaller than $10^{-13} \\rm \\, cm^3\\, s^{-1}$, ``medium\" to value in the range $10^{-13} - 10^{-11} \\rm \\, cm^3\\, s^{-1}$, and ``high\"\u00a0to a value greater than $10^{-11} \\rm \\, cm^3\\, s^{-1}$. The magenta curve refers to the training set. Clearly, the performance on the test set is improved if multiple runs are averaged, but for $N_r$ greater than a few, no further improvements are realized. Based upon the behavior exhibited in Figure 7, I selected $N_r = 10$ as the standard parameter for the calculations that yielded the results discussed in \\S3. Figures 8 and 9 show how the performance depends upon the final number of hidden neurons. The color coding in Figure 9 is the same as that in Figure 7. Clearly, if enough hidden neurons are provided, the fit to the training set becomes extremely accurate (lower right panel in Figure 8), with the RMS error for 64 neurons being less than 3$\\%$. However, the performance on the {\\it test} set worsens somewhat once the number of hidden neurons exceeds 16. This behavior indicates that the network is ``overfitting\" the data (e.g.\\ Smith 1993). The following analogy is presented by the use of polynomial fitting functions. Given a function of one variable that has been evaluated at $N$ values, a perfect polynomial fit to those values can always be obtained with a polynomial of order $N-1$. However, such a polynomial may show strong wiggles between (and beyond) the points where the function was evaluated (particularly if the function evaluations have errors). A lower order polynomial may fit the data points imperfectly but may be preferable for the purposes of interpolation and particularly extrapolation. Based upon the behavior exhibited in Figure 8 and 9, I selected $N_{hidden} = 16$ as the standard parameter for the calculations that yielded the results discussed in \\S3." }, "0911/0911.2187_arXiv.txt": { "abstract": " ", "introduction": "Whenever bending of light by gravitational fields is detected in astronomical observations it is generally very weak. Even in the spectacular formation of giant arcs, typical of the so-called strong lensing, photons are deflected by no more than a few arcseconds. This leads to the conclusion that the first order Einstein formula for the deflection angle is sufficient to explain all observed phenomenology% . On the other hand, the existence of very compact objects, such as neutron stars or black holes, is now well-established through many independent astrophysical observations. In the neighbourhood of such objects, electromagnetic radiation is generated and travels through very strong gravitational fields. In such extreme cases, the calculation of the deflection of photons needs to be pushed beyond the first order approximation on which the Einstein formula relies. It must be said that no direct observation of gravitational lensing by black holes or other compact objects has ever been performed up to date. Yet, some peculiarities of the emission spectra of black holes, such as the shape of the $K_\\alpha$ line of the iron, have been interpreted as the final effect of the strong deflection of photons emitted by the accretion disk. The aim of this article is to review the theoretical aspects of gravitational lensing by black holes, and discuss the perspectives for realistic observations. We will first treat lensing by spherically symmetric black holes, in which the formation of infinite sequences of higher order images emerges in the clearest way. We will then consider the effects of the spin of the black hole, with the formation of giant higher order caustics and multiple images. Finally, we will consider the perspectives for observations of black hole lensing, from the detection of secondary images of stellar sources to the interpretation of iron K-$\\alpha$ lines and direct imaging of the shadow of the black hole. ", "conclusions": "" }, "0911/0911.0711_arXiv.txt": { "abstract": "We investigate how the divergence-free property of magnetic fields can be exploited to resolve the azimuthal ambiguity present in solar vector magnetogram data, by using line-of-sight and horizontal heliographic derivative information as approximated from discrete measurements. Using synthetic data we test several methods that each make different assumptions about how the divergence-free property can be used to resolve the ambiguity. We find that the most robust algorithm involves the minimisation of the absolute value of the divergence summed over the entire field of view. Away from disk centre this method requires the sign and magnitude of the line-of-sight derivatives of all three components of the magnetic field vector. ", "introduction": "Measurements of the vector magnetic field in the solar atmosphere are crucial for understanding a wide variety of physical phenomena. Using the linear polarisation of magnetically sensitive spectral lines to infer the component of the field transverse to the line-of-sight results in an ambiguity of 180$^\\circ$ in its direction \\cite{1969PhDT3H}. This ambiguity must be resolved in order to completely determine the magnetic field vector. Several methods are currently in use to resolve the ambiguity (for overviews see \\opencite{metcalfetal06}; \\opencite{lekaetal2008}). In general, each method involves an assumption or approximation that may not be appropriate for observations of solar magnetic fields. A promising approach, which can potentially avoid unrealistic assumptions, is to exploit the divergence-free property of magnetic fields (\\myeg \\opencite{wuai1990}; \\opencite{1993AA...278..279C}; \\opencite{1993AA...279..214L}; \\opencite{1999AA...347.1005B}; \\opencite{lietal07}; \\opencite{cb2007}). One challenge in calculating the divergence is that it requires information about the variation of the field in the direction perpendicular to the solar surface. To this end it is possible to derive from observations the variation of the field along the line-of-sight direction (\\myeg \\opencite{ruizcobodeltooroiniesta92}; \\opencite{1994AA...291..622C}; \\opencite{1995ApJ...439..474M}; \\opencite{1996SoPh..164..169D}; \\opencite{1996SoPh..169...79L}; \\opencite{1998ApJ...494..453W}, \\citeyear{2001ApJ...547.1130W}; \\opencite{2000ApJ...530..977S}; \\opencite{eibeetal2002}; \\opencite{lekametcalf2003}; \\opencite{2005ApJ...631L.167S}, \\citeyear{2007ApJS..169..439S}). The line-of-sight direction is not perpendicular to the solar surface, except at disk centre; therefore some care is required to correctly treat the geometrical perspective. In \\citeauthor{cb2007}~(\\citeyear{cb2007}, henceforth paper~I) we showed that the ambiguity can be resolved with line-of-sight and horizontal heliographic derivative information by using the divergence-free property without additional assumptions about the nature of the field. To resolve the ambiguity away from disk centre we demonstrated that it is sufficient to determine four pieces of information: {\\it i)} the sign of the line-of-sight derivative of the line-of-sight component of the magnetic field; {\\it ii)} the sign and magnitude of the line-of-sight derivative of the magnitude of the transverse component of the field; {\\it iii)} the sign and magnitude of the line-of-sight derivative of the azimuthal angle; and {\\it iv)} the sign and magnitude of the horizontal heliographic derivatives of the magnitude of the transverse component of the field and the azimuthal angle. Paper~I was a theoretical examination in which we assumed that all necessary derivatives could be determined exactly at a given location and that the solar surface could be represented by the heliographic plane. The aim of this paper is to investigate how the divergence-free property can be used to resolve the ambiguity for data where the magnetic field is measured at discrete spatial locations. Using synthetic data we test two methods that have previously appeared in the literature (\\myeg \\opencite{wuai1990}; \\opencite{1993AA...278..279C}; \\opencite{1993AA...279..214L}; \\opencite{1999AA...347.1005B}; \\opencite{lietal07}); both of these are based on the divergence but they make different assumptions about how to use it to resolve the ambiguity. We show that both of these methods can produce significant errors. For this reason, we present a more robust algorithm, the global minimisation method. The outline of this paper is as follows. In Section~\\ref{sec_divb} we describe the derivation of the divergence-free condition for any position on the solar disk in terms of observable quantities and discuss the consequences for ambiguity resolution. In Section~\\ref{sec_synth} we review the synthetic data and metrics that will be used to test the performance of the ambiguity resolution algorithms. In Section~\\ref{algorithms} we test the algorithms and the validity of the assumptions made. In Section~\\ref{sec_conc} we draw conclusions. ", "conclusions": "\\label{sec_conc} We investigate how the divergence-free property of magnetic fields can be exploited to resolve the azimuthal ambiguity that is present in solar vector magnetogram data by using line-of-sight and horizontal heliographic derivative information approximated from discrete measurements. Using synthetic data we test several methods that each make different assumptions about how to use the divergence to resolve the ambiguity. We find that all of the assumptions considered can be violated when the divergence is approximated from discrete measurements (depending on the nature of the field and the resolution of the observations), but some assumptions are more robust than others. The \\citeauthor{wuai1990} criterion expresses the divergence-free condition as an inequality (\\myeg \\opencite{wuai1990}; \\opencite{1993AA...278..279C}; \\opencite{1993AA...279..214L}; \\opencite{lietal07}). We find that methods based on this criterion are very sensitive to the smoothness of the initial configuration of azimuthal angles and the order in which pixels are examined. In addition, the underlying assumption can be incorrect over a significant fraction of the field of view when derivatives are approximated from discrete measurements. The sequential minimisation method (\\opencite{1999AA...347.1005B}) assumes that the magnitude of the divergence is minimised over the pixels used to approximate the divergence at each point. Our tests indicate that this approach is more robust than the \\citeauthor{wuai1990} criterion but can produce errors when the underlying assumption is violated. Moreover, the algorithm can propagate erroneous solutions into regions that do not violate the underlying assumption. We find that the most promising algorithm is the global minimisation method which assumes that the correct configuration of azimuthal angles corresponds to the minimum of the magnitude of the divergence summed over the entire field of view. This method relies on the magnitude of the divergence. Therefore, with this method, the {\\it sign and magnitude} of the line-of-sight derivatives of all three components of the magnetic field vector are required to resolve the azimuthal ambiguity away from disk centre. For the synthetic data employed here, the field is sampled at discrete spatial locations where the exact value of the field is provided. Thus, the effects of noise and unresolved structure in both the horizontal heliographic and line-of-sight directions are neglected. The performance of the global minimisation method is very promising, but before this approach can be reliably applied to real magnetogram data its performance should be tested with more realistic synthetic data that include the effects of noise and unresolved structure (\\myeg \\opencite{lekaetal2008}). \\begin{acks} The authors would like to thank Yuhong Fan for providing one of the synthetic data sets used in this investigation. This work was supported by funding from NASA under contracts NNH05CC75C and NNH09CE60C. \\end{acks}" }, "0911/0911.2008_arXiv.txt": { "abstract": "Analytical relations are derived for the amplitude of astrometric, photometric and radial velocity perturbations caused by a single rotating spot. The relative power of the star spot jitter is estimated and compared with the available data for $\\kappa^1$ Ceti and HD 166435, as well as with numerical simulations for $\\kappa^1$ Ceti and the Sun. A Sun-like star inclined at $i=90\\degr$ at 10 pc is predicted to have a RMS jitter of $0.087$ \\uas\\ in its astrometric position along the equator, and $0.38$ m s$^{-1}$ in radial velocities. If the presence of spots due to stellar activity is the ultimate limiting factor for planet detection, the sensitivity of SIM Lite to Earth-like planets in habitable zones is about an order of magnitude higher that the sensitivity of prospective ultra-precise radial velocity observations of nearby stars. ", "introduction": "With the anticipated launch of the SIM Lite mission in the near future, we are embarking on a long and exciting journey of exoplanet detection by astrometric means. One of the main goals of this mission is the detection of habitable Earth-like planets around nearby stars \\citep{unw}. To date, most of the exoplanet discoveries have been made by the Doppler-shift technique, while the astrometric method has been limited to the use of the FGS on the Hubble Space Telescope \\citep{ben} and to ground-based CCD observations of low-mass stars with giant, super-Jupiter, planetary companions \\citep{sha}. In achieving the strategic goal of confident detection of rocky, Earth-sized planets in the habitable zone, the prospective astrometric and spectroscopic ultra-precise measurements will encounter a number of limitations of technical and astrophysical nature. For the Doppler-shift technique, many of these limitations will be dealt with by further improvement in the instrumentation or refinement of the observational procedure \\citep{may}. However, the presence of astrophysical noise due to stellar magnetic activity emerges as the ultimate bound on the sensitivity of planet detection techniques, and the only remedy suggested thus far, is selection of particularly inactive, slowly rotating stars. Indeed, a very small fraction of stars in the high-precision HARPS program of exoplanet search exhibit radial velocity (RV) scatter of less than 0.5 m s$^{-1}$. Although this type of variability is probably driven by the rotation of bright and dark structures on the surface (star spots and plage areas), the frequency power spectrum of such perturbations can be fairly flat, extending to frequencies much higher or lower than the rotation, as shown by \\citet{cat} for the Sun. Arguments have been presented \\citep[e.g., ][]{eri} that star spots also result in very large astrometric noise of $\\sim 10$ $\\mu$AU, which should thwart discovery of habitable Earth analogs. The aim of this paper is to quantify the effects of rotating spots in astrometric photometric and RV measurements more accurately, taking into account the limb darkening, geometric projection and differential rotation, and to assess the expected vulnerability of the RV and astrometric methods to such perturbations. We do this by direct analysis as well as by numerical simulation, and support our findings by the data for the Sun and two rapidly rotating stars. ", "conclusions": "\\label{con.sec} Our results for the Sun are in good agreement with the approximate relations by \\citet{eri}, who estimated a positional standard deviation of 0.7 $\\mu$AU. At the same time, their conclusion that \"for most spectral types the astrometric jitter is expected to be of the order of 10 micro-AU or greater\" is misleading, because it is largely based on overestimated values of photometric variability from ground-based observations, and it does not differentiate the luminosity classes of giants and dwarfs. It can not be concluded that the Sun is exceptionally inactive compared to its peers just because the ultra-precise solar irradiance data, such as PMOD or SOHO, reveal a much smaller scatter than the inferior photometric data for other stars. We investigated the indices of chromospheric activity ($\\log R^\\prime_{\\rm HK}$) and available rotation periods for some 80 SIM targets, and found that half of them should rotate with the same rate as the Sun, or slower. \\citet{hal} presented a detailed study of variability of solar-type stars and its relation to the index of chromospheric activity $\\log R'_{\\rm HK}$, based on 14 years of photometric and spectroscopic observations. They found that the Sun at $\\log R'_{\\rm HK}=-4.96$ is not more variable than its F-G peers at the low end of the activity distribution. Given that most Solar-type stars exhibit similar low levels of chromospheric activity \\citep{gra}, we expect that finding stars with levels of jitter similar to, or lower than the Sun, should not be a problem. Low-jitter, stable stars are common and plentiful, which augurs well for the prospects of finding small, rocky planets with {\\it Kepler} \\citep{bat}. \\begin{deluxetable}{lrrr} \\tabletypesize{\\scriptsize} \\tablecaption{Observable signals and star spot jitters for an Earth-like planet orbiting a typical dwarf star at 10 pc. \\label{snr.tab}} \\tablewidth{0pt} \\startdata Star type \\dotfill & Sun & F5V & K5V \\\\ Rotation period, d \\dotfill & 25.4 & 18 & 30 \\\\ Astrometric signal, \\uas\\ \\dotfill & 0.30 & 0.23 & 0.45 \\\\ RV signal, m s$^{-1}$ \\dotfill & 0.089 & 0.078 & 0.109 \\\\ Astrometric jitter, \\uas\\ \\dotfill & 0.087 & 0.113 & 0.063\\\\ RV jitter, m s$^{-1}$ \\dotfill & 0.38 & 0.69 & 0.23 \\\\ Astrometric SNR \\dotfill & 3.4 & 2.0 & 7.1 \\\\ RV SNR \\dotfill & 0.23 & 0.11 & 0.47 \\\\ \\enddata \\end{deluxetable} The SIM Lite Astrometric Observatory (formerly known as the Space Interferometry Mission) will achieve a single-measurement accuracy of 1 \\uas\\ or better in the differential regime of observation \\citep{unw, sim}. Several previous studies have addressed SIM's exoplanet detection and orbital characterization capabilities \\citep[][and references therein]{catpa}. The \"Tier 1\" program includes $\\sim 60$ nearby stars for which the astrometric signature of a terrestrial habitable planet is large enough to be confidently measured by SIM. The recently completed double-blind test \\citep{tra} demonstrated that Earth-like planets around nearby stars can be discovered and measured even in complex planetary systems. In this paper, we are concerned with the more general question of the ultimate limit to planet detectability set by the activity-related jitter. Assuming that the instrumentation progresses to levels where observational noise becomes insignificant, which technique holds the best prospects for detection of habitable Earth-like planets? Table \\ref{snr.tab} summarizes the expected RMS jitter and the exoplanet signal for three typical nearby stars. In all cases, the solar spot filling factor ($r^2$) is assumed. The signal-to-noise ratios (SNR) per observation include only the star spot jitter. Note that the astrometric SNR in this case is independent of the distance to the host star, because both the signal and the star spot jitter are inversely proportional to distance. As a measure of the relative sensitivity of the two methods, the ratio of the SNR values for a given star is independent of the planet mass or the filling factor to first-order approximation. The physical radius of the star and the period of rotation are the two parameters with the largest impact on the relative sensitivity, but their combined effect is rather modest for normal stars, as is seen in the Table. Therefore, in the ultimate limit of exoplanet detection defined by intrinsic astrophysical perturbations, the astrometric method is at least an order of magnitude more sensitive than the Doppler technique for most nearby solar-type stars." }, "0911/0911.1301_arXiv.txt": { "abstract": "High contrast coronagraphic imaging of the immediate surrounding of stars requires exquisite control of low-order wavefront aberrations, such as tip-tilt (pointing) and focus. We propose an accurate, efficient and easy to implement technique to measure such aberrations in coronagraphs which use a focal plane mask to block starlight. The Coronagraphic Low Order Wavefront Sensor (CLOWFS) produces a defocused image of a reflective focal plane ring to measure low order aberrations. Even for small levels of wavefront aberration, the proposed scheme produces large intensity signals which can be easily measured, and therefore does not require highly accurate calibration of either the detector or optical elements. The CLOWFS achieves nearly optimal sensitivity and is immune from non-common path errors. This technique is especially well suited for high performance low inner working angle (IWA) coronagraphs. On phase-induced amplitude apodization (PIAA) type coronagraphs, it can unambiguously recover aberrations which originate from either side of the beam shaping introduced by the PIAA optics. We show that the proposed CLOWFS can measure sub-milliarcsecond telescope pointing errors several orders of magnitude faster than would be possible in the coronagraphic science focal plane alone, and can also accurately calibrate residual coronagraphic leaks due to residual low order aberrations. We have demonstrated $\\approx 10^{-3} \\lambda/D$ pointing stability in a laboratory demonstration of the CLOWFS on a PIAA type coronagraph. ", "introduction": "High performance coronagraphs with small inner working angle (IWA) are unavoidably very sensitive to small pointing errors and other low order aberrations \\citep{lloy05,shak05,siva05,beli06,guyo06}. This property is due to the fact that the wavefront of a source at small angular distance (typically between 1 and 2 $\\lambda/D$) from the optical axis is \"similar\" (in the linear algebra sense of the term) to an on-axis wavefront with a small ($\\ll \\lambda/D$) pointing error. If the coronagraph must \"transmit\" the former, it will also transmit a significant part of the latter, and therefore be extremely sensitive to pointing errors. This behavior is indeed verified by performance comparison between coronagraph concepts \\citep{guyo06}. \\begin{figure*}[htb] \\includegraphics[scale=0.35]{f1.eps} % \\caption{\\label{fig:lowfsprinciple} Optical layout of a coronagraphic low order wavefront sensor system, shown here with a PIAA coronagraph. See text for details.} \\end{figure*} For high performance coronagraphs, such as the Phase-Induced Amplitude Apodization (PIAA) coronagraph \\citep{guyo03,guyo05} used as example in this paper, the coronagraph should ideally be designed to balance IWA against stellar angular diameter, which sets a fundamental limit on the achievable coronagraph performance. Such coronagraphs are therefore pushed to become sensitive to pointing errors corresponding to the angular size of nearby stars, roughly 1 milliarcsecond (mas), and are also highly sensitive to other low order wavefront errors such as focus and astigmatism. Milliarcsecond-level pointing error can increase stellar leakage in the coronagraph to the point where a planet would be lost in photon noise. Even smaller errors can create, if not independantly measured, a signal which is very similar to a planet's image. A robust and accurate measurement of low order aberrations (especially tip-tilt errors, which are easily generated by telescope pointing errors and vibrations) is therefore essential for high contrast coronagraphic observations at small angular separation. The science focal plane after the coronagraph is unfortunately \"blind\" to small levels of low order aberrations, which can only be seen when already too large to maintain high contrast in the coronagraphic science image. A better option is to monitor pointing errors by using starlight which would otherwise be rejected by the coronagraph. This scheme was successfully implemented on the LYOT project coronagraph \\citep{oppe04,digb06}. Alternatively, the measurement could be performed independently from the coronagraph optical train (for example, the wavefront sensor in the adaptive optics system upstream of the coronagraph). We propose in this paper an improved solution to obtain accurate measurement of several low-order aberrations including pointing: the coronagraphic low-order wavefront sensor (CLOWFS). The wavefront control requirements for a PIAA coronagraph are first clearly defined in \\S\\ref{sec:requ}. The CLOWFS principle is presented in \\S\\ref{sec:principle}. Wavefront reconstruction algorithms and CLOWFS sensitivity are discussed in \\S\\ref{sec:wfreconstr}. The aberration sensitivity of a PIAA coronagraph equipped with a CLOWFS is discussed in \\S\\ref{sec:aberrsens}, and the results of a laboratory demonstration on a PIAA Coronagraph system are shown in \\S\\ref{sec:labexp}. ", "conclusions": "The CLOWFS design presented in this paper can efficiently measure low order aberrations \"for free\", as it uses light that would otherwise be discarded by the coronagraph. Both the hardware configuration and software algorithms presented are easy to implement and their performance is robust against calibration errors, chromaticity, non-common path errors and small errors/aberrations in the optical components. The CLOWFS pointing measurement can also lead to improved astrometric accuracy for the position of faint companions, as the star position on the image is usually difficult to measure in coronagraphic images \\citep{digb06}. In this paper, we have studied a CLOWFS design on a low-IWA PIAA coronagraph. Although coronagraphs with larger IWA can tolerate larger amount of low order aberrations (see for example Kuchner \\& Traub 2002; Kuchner et al. 2005), they require these aberration to be measured ahead of the coronagraphed beam because they are \"blind\" to aberrations until they are large enough to produce significant coronagraph leaks. The CLOWFS would therefore be very useful to any high contrast coronagraph, and the optical design presented in this paper can readily be used on any coronagraph where a focal plane mask physically blocks starlight. The CLOWFS can also be applied to phase mask coronagraphs. For such coronagraphs (see for example Roddier \\& Roddier 1997; Rouan et al. 2000; Palacios 2005), where starlight is diffracted outside the pupil by the focal plane mask, a modified Lyot stop is placed on the pupil plane to reflect starlight to the CLOWFS. In addition, using a pattern matching algorithm, the CLOWFS can estimate low-order wavefront aberrations accurately and quickly even in non-linear region. These new CLOWFS designs and their performances will be presented in an upcoming paper." }, "0911/0911.1592_arXiv.txt": { "abstract": "Stacking analysis is a means of detecting faint sources using a priori position information to estimate an aggregate signal from individually undetected objects. Confusion severely limits the effectiveness of stacking in deep surveys with limited angular resolution, particularly at far infrared to submillimeter wavelengths, and causes a bias in stacking results. Deblending corrects measured fluxes for confusion from adjacent sources; however, we find that standard deblending methods only reduce the bias by roughly a factor of two while tripling the variance. We present an improved algorithm for simultaneous stacking and deblending that greatly reduces bias in the flux estimate with nearly minimum variance. When confusion from neighboring sources is the dominant error, our method improves upon RMS error by at least a factor of three and as much as an order of magnitude compared to other algorithms. This improvement will be useful for Herschel and other telescopes working in a source confused, low signal to noise regime. ", "introduction": "At every wavelength of astronomy, there is a need to extract faint signals buried in noisy images. For example, a certain class of objects may be lurking undetected below the noise threshold of a data set at a particular wavelength. However, if accurate positions for the objects are known a priori from detections at other wavelengths, then this prior information may be exploited to break through the noise floor and achieve a detection, at least in a statistical sense, for the objects in question. Stacking is an effective analysis technique in cases where noisy images at one wavelength are complemented by catalogs of known objects that have been detected in the same region of sky at other wavelengths. The noisy data at known object positions are averaged, and the resulting aggregate flux measurement has an enhanced signal to noise ratio due to averaging, which reduces the noise by a factor $1/\\sqrt{R}$, where $R$ is the number of target objects in the stack. As an early application of this method in x-ray astronomy, stacking analysis was used to study x-ray emission from stars \\citep{1985ApJ...289..279C}; subsequently, stacking became a standard x-ray analysis technique that, among others, has been applied to normal galaxies \\citep{2001AJ....122....1B}, Lyman Break Galaxies (LBGs) \\citep{2001ApJ...558L...5B} and radio sources \\citep{2003MNRAS.345..939G}. At optical to near infrared wavelengths stacking has also become widely adopted, and has been used for example to study galactic halos \\citep{Zibetti2004} and intergalactic stars \\citep{Zibetti2005}, as well as the cosmic infrared background \\citep{2006A&A...451..417D}, star forming galaxies \\citep{2004ApJ...610..745L, 2007ApJ...671..278G} and Extremely Red Objects (EROs) \\citep{1997AJ....113..474H}. At submillimeter wavelengths, stacking has been employed to study the submillimeter background \\citep{2000MNRAS.318..535P, 2004ApJS..154..118S, 2006ApJ...647...74W, 2006ApJ...644..769D, 2009arXiv0904.1205M, 2009arXiv0904.0028G}, optical/IR color selected galaxies \\citep{2004ApJ...605..645W, 2005ApJ...631L..13D, 2007MNRAS.381.1154T, 2009MNRAS.394....3D, 2009arXiv0904.0028G}, and radio detected galaxies \\citep{2008MNRAS.385.2225S}. Similarly, at radio wavelengths stacking has been widely adopted, for example to study LBGs \\citep{2007ApJ...660L..77I, 2008ApJ...689..883C}, star forming BzK galaxies \\citep{2005ApJ...631L..13D, 2007MNRAS.381.1154T, 2009MNRAS.394....3D}, Distant Red Galaxies (DRGs) \\citep{2007ApJ...660L..77I}, EROs \\citep{2009MNRAS.394....3D} and quasars \\citep{2007ApJ...654...99W}. It has been widely recognized that confusion noise limits astronomical measurements particularly at long wavelengths \\citep{1974ApJ...188..279C, 2001AJ....121.1207H} and confusion poses special complications for stacking analyses. When multiple sources crowd a resolution element of the data, then blending of sources must be accounted for in the stacking procedure. Blending is a particular problem in the submillimeter regime, where at present images suffer from poor resolution compared to many other astronomical observations. Methods have been used for deblending in submillimeter data, however a systematic assessment of the effectiveness of these techniques has not been previously reported. This paper addresses the problem of deblending the various unresolved sources of submillimeter emission in order to effectively stack objects in the confusion dominated, low signal to noise, regime. We present an algorithm that accomplishes simultaneous stacking and deblending of these data and demonstrate its effectiveness through Monte Carlo simulation. We also demonstrate that the standard approaches to deblending are in fact subject to substantial bias and scatter when applied to real data. ", "conclusions": "The simultaneous stacking and deblending algorithm presented here demonstrated more than a factor of three improvement in RMS error over other deblending algorithms at all signal to noise ratios, and by an order of magnitude at large signal to noise ratios. This improvement translates into a greater sensitivity for stacking analyses that will be applicable to present and future far infrared and submillimeter surveys. For example, the Herschel SPIRE instrument will reach 1$\\sigma$ survey depths of 3 mJy in 10-16 hours at 18-36$''$ spatial resolution \\citep{2008SPIE.7010E...4G}. At this depth, and with 1,000 known source positions, stacking analyses can reach a depth of 0.1 mJy when limited only by statistical error. This depth would be sufficient to detect stacking signals from color selected galaxies such as BzKs, DRGs and EROs that have been detected in stacking at 870 $\\mu$m with fluxes in the 0.2-0.4 mJy range with the ground based LABOCA instrument \\citep{2009arXiv0904.0028G}. However, with the factor of three poorer sensitivity caused by standard deblending algorithms, these objects would remain out of reach to stacking analysis with Herschel, or their reported stacked fluxes would be significantly biased." }, "0911/0911.1889_arXiv.txt": { "abstract": "We present new measurements of the energy spectra of cosmic-ray (CR) nuclei from the second flight of the balloon-borne experiment Cosmic Ray Energetics And Mass (CREAM). The instrument included different particle detectors to provide redundant charge identification and measure the energy of CRs up to several hundred TeV. The measured individual energy spectra of C, O, Ne, Mg, Si, and Fe are presented up to $\\sim 10^{14}$ eV. The spectral shape looks nearly the same for these primary elements and it can be fitted to an $E^{-2.66 \\pm 0.04}$ power law in energy. Moreover, a new measurement of the absolute intensity of nitrogen in the 100-800 GeV/$n$ energy range with smaller errors than previous observations, clearly indicates a hardening of the spectrum at high energy. The relative abundance of N/O at the top of the atmosphere is measured to be $0.080 \\pm 0.025\\, $(stat.)$\\, \\pm 0.025\\,$(sys.) at $\\sim\\,$800 GeV/$n$, in good agreement with a recent result from the first CREAM flight. ", "introduction": "Experimental studies of charged cosmic rays (CRs) are focused on the understanding of the acceleration mechanism of high-energy CRs, identification of their sources, and clarification of their interactions with the interstellar medium (ISM). The origin of CRs is still under debate, although direct measurements with instruments on stratospheric balloons or in space have provided information on their elemental composition and energy spectra, and indirect detection by ground-based experiments has extended the measurements up to the end of the observed all-particle spectrum. \\\\ It is generally accepted that CRs are accelerated in blast waves of supernova remnants (SNRs), which are the only galactic candidates known with sufficient energy output to sustain the CR flux \\citep{Bell, hillas}. Recent observation of emission of TeV gamma-rays from SNR RX J1713.7-3946 by the HESS Cherenkov telescope array proved that high-energy charged particles are accelerated in SNR shocks up to energies beyond 100 TeV \\citep{HESS}. This result is consistent with expectations of the class of theoretical models that predict the existence of a rigidity-dependent limit, above which the diffusive shock acceleration becomes inefficient. The maximum energy attainable by a nucleus of charge $Z$ may range from $Z\\times$10$^{14}$ eV to $\\sim$ $Z\\times$10$^{17}$ eV depending on the model and types of supernovae considered \\citep{LC, Berezhko, parizot}. In this scenario, the energy spectra of elements exhibit a $Z$-dependent cutoff. As a consequence, the CR elemental composition is expected to change as a function of energy, marked by a depletion of low-$Z$ nuclei in the several hundred TeV region. This mechanism has been proposed as a possible explanation of the steepening (``knee'') in the CR energy spectrum (described by a power law: $dN$/$dE \\propto E^{-\\gamma}$) with a change of the spectral index from $\\gamma\\approx\\,$2.7 to $\\gamma\\approx\\,$3.1 observed at energies around 4 PeV. An alternative approach is adopted by models that relate the knee to leakage of CRs from the Galaxy. In this case, the knee is expected to occur at lower energies for light elements as compared to heavy nuclei, due to the rigidity-dependence of the Larmor radius of CR particles propagating in the galactic magnetic field \\citep{horandel2}.\\\\ A detailed understanding of the propagation of CRs during their wanderings inside the Galaxy and their interactions with the ISM is needed to infer the injection spectra of individual elements at the source from the observed spectra at Earth. CR nuclei can be divided into primaries (i.e., H, He, C, O, Fe, etc.) that come from CR sources, and secondaries (such as Li, Be, B, Sc, Ti, V, etc.), which are products of the interactions of primary nuclei with the ISM. The amount of material traversed by CRs between injection and observation can be derived from the measured ratio of secondary-to-primary nuclei, such as the boron-to-carbon ratio (B/C) or the ratio of sub-iron elements to iron. Likewise, their confinement time in the Galaxy can be determined through measurements of long-lived radioactive secondary nuclei \\citep{longair, yanasak}. Observations from space-based experiments like CRN on the Spacelab2 mission of the Space Shuttle \\citep{CRN2} and C2 \\citep{HEAO} and Heavy Nuclei Experiment (HNE) \\citep{binns} onboard the {\\em HEAO 3} satellite have shown that the energy spectra of the secondary nuclei are steeper than those of the primaries. The escape pathlength of CRs from the Galaxy depends on the magnetic rigidity as $R^{-\\delta}$, with the parameter $\\delta\\approx\\,$0.6. This implies that the spectra observed at Earth are steeper than the injection spectra, i.e., the spectral index at the source is smaller by the value $\\delta$. \\\\ Those pioneering measurements were statistically limited at energies of order $\\sim$ 10$^{11}$ eV. Accurate direct measurements of the energy spectra of individual elements into the knee region are needed to discriminate between different astrophysical models proposed to explain the acceleration and propagation mechanisms. Recently, a new generation of balloon-borne experiments has performed accurate measurements of CRs in the previously experimentally unexplored TeV region \\citep{ATIC, TRACER}. Among these, the Cosmic Ray Energetics And Mass (CREAM) experiment was designed to measure directly the elemental composition and the energy spectra of CRs from hydrogen to iron over the energy range 10$^{11}$-10$^{15}$ eV in a series of flights \\citep{seo}. Since 2004, four instruments were successfully flown on long-duration balloons in Antarctica. The instrument configurations varied slightly in each mission due to various detector upgrades. The first flight produced important results extending the measurement of the relative abundances of CR secondary nuclei B and N close to the highest energies ($\\sim$ 1.5 TeV/$n$) allowed by the irreducible background generated by the residual atmospheric overburden at balloon altitudes. The data clearly indicate that the escape pathlength of CRs from the Galaxy has a power-law rigidity-dependence, with the parameter $\\delta$ in the range 0.5-0.6 \\citep{Ahn2008}. In this work, we present new measurements of the energy spectra of the major primary CR nuclei from carbon to iron, up to a few TeV/$n$ made with the second CREAM flight (CREAM-II). A new measurement of the nitrogen intensity with unprecedented statistics is also presented up to 800 GeV/$n$. The measured N/O ratio confirms the results of CREAM-I. The procedure used to analyze the data and reconstruct the CR energy spectra is described in detail together with an assessment of the systematic uncertainties.\\\\ ", "conclusions": "CREAM-II carried out measurements of high-$Z$ CR nuclei with an excellent charge resolution and a reliable energy determination. The use of a dual layer of silicon sensors allowed us to unambiguously identify each individual element up to iron and to reduce the background of misidentified nuclei. We demonstrated the feasibility of measuring the particle energy up to hundreds of TeV by using a thin ionization sampling calorimeter, with sufficient resolution to reconstruct accurately the CR spectra. \\\\ The energy spectra of the major primary heavy nuclei from C to Fe were measured up to $\\sim\\,$10$^{14}$ eV and found to agree well with earlier direct measurements. All the spectra follow a power law in energy with a remarkably similar spectral index $\\bar{\\gamma} = -2.66 \\pm 0.04$, as is plausible if they have the same origin and share the same acceleration and propagation processes. \\\\ A new measurement of the nitrogen intensity in an energy region so far experimentally unexplored indicates a harder spectrum than at lower energies, supporting the idea that nitrogen has secondary as well as primary contributions. These results provide new clues for understanding the CR acceleration and propagation mechanisms, but further accurate observations with high statistics beyond 10$^{14}$ eV are needed to finally unravel the mystery of CR origin." }, "0911/0911.0288_arXiv.txt": { "abstract": "We consider the problem of characterisation of burst sources detected with the Laser Interferometer Space Antenna (LISA) using the multi-modal nested sampling algorithm, {\\sc MultiNest}. We use {\\sc MultiNest} as a tool to search for modelled bursts from cosmic string cusps, and compute the Bayesian evidence associated with the cosmic string model. As an alternative burst model, we consider sine-Gaussian burst signals, and show how the evidence ratio can be used to choose between these two alternatives. We present results from an application of {\\sc MultiNest} to the last round of the Mock LISA Data Challenge, in which we were able to successfully detect and characterise all three of the cosmic string burst sources present in the release data set. We also present results of independent trials and show that {\\sc MultiNest} can detect cosmic string signals with signal-to-noise ratio (SNR) as low as $\\sim 7$ and sine-Gaussian signals with SNR as low as $\\sim8$. In both cases, we show that the threshold at which the sources become detectable coincides with the SNR at which the evidence ratio begins to favour the correct model over the alternative. ", "introduction": "The development of data analysis algorithms for gravitational wave detectors on the ground and for the proposed space-based gravitational wave detector, the Laser Interferometer Space Antenna (LISA)~\\cite{lisa}, is an active area of current research. LISA data analysis activities are being encouraged by the Mock LISA Data Challenges~\\cite{mldc} (MLDCs). Prior to the most recent round, round 3, which finished in April 2009, the MLDCs had included data sets containing individual and multiple white-dwarf binaries, a realisation of the whole galaxy of compact binaries, single and multiple non-spinning supermassive black hole (SMBH) binaries, either isolated or on top of a galactic confusion background and isolated extreme-mass-ratio inspiral (EMRI) sources in purely instrumental noise. Round 3 included a realisation of the galactic binaries that included chirping systems, a single data set containing multiple overlapping EMRIs and data sets containing three new types of source --- inspirals of spinning supermassive black hole binaries, bursts from cosmic string cusps and a stochastic background of gravitational radiation~\\cite{mldc}. Several of these sources, including the EMRI signals, the spinning black hole binaries and the cosmic string bursts, have highly multi-modal likelihood surfaces which can cause problems for algorithms such as Markov Chain Monte Carlo (MCMC), as these may become stuck in secondary maxima rather than finding the primary mode~\\cite{EtfAG,neilemri,BBGPa,BBGPb}. Recently~\\cite{MNnospin} we applied a new algorithm to the problem of gravitational wave data analysis, {\\sc MultiNest}~\\cite{feroz08,multinest}, which is a nested sampling algorithm, optimized for problems with highly multi-modal posteriors. We demonstrated that {\\sc MultiNest} could be used as a search tool and to recover posterior probability distributions for the case of data sets containing multiple signals from non-spinning SMBH binary inspirals. We used the algorithm as a search tool in MLDC round 3 to analyse two of the data sets --- the spinning SMBH inspirals (challenge 3.2) and the cosmic string bursts (challenge 3.4). We will discuss the second of these applications in this paper. The nested sampling algorithm~\\cite{Skilling04} was developed as a tool for evaluating the Bayesian evidence. It employs a set of live points, each of which represents a particular set of parameters in the multi-dimensional search space. These points move as the algorithm progresses, and climb together through nested contours of increasing likelihood. At each step the algorithm works by finding a point of higher likelihood than the lowest likelihood point in the live point set and then replacing the lowest likelihood point with the new point. The difficulty is to sample, efficiently, new points of higher likelihood from the remaining prior volume. {\\sc MultiNest} achieves this using an ellipsoidal rejection sampling scheme% . It has demonstrated orders of magnitude improvement over standard methods in several applications in cosmology and particle physics~\\cite{2008arXiv08100781F,2008arXiv08111199F,Feroz:2008wr,2008JHEP12024T}. It also proved to be very effective in a gravitational wave context for the test case mentioned above~\\cite{MNnospin}. {\\sc MultiNest} can be used as a search tool and returns the posterior probability distributions as a by-product. We employed the algorithm successfully to analyse MLDC challenge 3.4, accurately finding all three of the cosmic string burst sources present in the data set. This will be discussed in more detail in Section~\\ref{sec:mldcres}. The cosmic string cusp is a special type of burst source, since the waveform is known and can be modelled, so a matched filtering search is possible and that is what {\\sc MultiNest} does in computing the likelihoods across the parameter space. However, there may be bursts of gravitational waves in the LISA data stream from other sources and the question arises as to whether we would be able to detect them and if we can distinguish cosmic string bursts from bursts due to these other sources. The evidence that {\\sc MultiNest} computes is one tool that can be used to address model selection. Evidence has been used for gravitational wave model selection to test the theory of relativity in ground based observations~\\cite{Veitch:2008wd}, and, in a LISA context, to distinguish between an empty data set and one containing a signal~\\cite{littenberg09}. Calculation of an evidence ratio requires a second model for the burst as an alternative to compare against. We chose to use a sine-Gaussian waveform as a generic alternative burst model, as this is one of the burst models commonly used in LIGO data analysis (see for instance~\\cite{LSCUsingSG}). We find that the evidence is a powerful tool for characterising bursts --- for the majority of detectable bursts, the evidence ratio strongly favours the true model over the alternative. While the sine-Gaussian model does not necessarily well describe all possible un-modelled bursts, these results suggest that the evidence can be used to correctly identify any cosmic string bursts that are present in the LISA data stream. This paper is organised as follows. In Section~\\ref{sec:methods} we describe the methods that we employ in this analysis --- we describe Bayesian inference, nested sampling, the {\\sc MultiNest} algorithm, the two burst waveform models we use and the detector noise spectral density. In Section~\\ref{sec:res} we present our results, including a description of our methods and results for the analysis of MLDC challenge 3.4 (in Section~\\ref{sec:mldcres}); an estimate of the signal-to-noise ratio (SNR) required for detection of cosmic string and sine-Gaussian bursts using the {\\sc MultiNest} algorithm; and a calculation of the evidence ratio of the two models for data sets containing each type of source. We finish in Section~\\ref{sec:discuss} with a summary and discussion of directions for further research. ", "conclusions": "\\label{sec:discuss} We have considered the use of the multi-modal nested sampling algorithm \\MN~for detection and characterisation of cosmic string burst sources in LISA data. As a search tool, the algorithm was able successfully to find the three cosmic string bursts that were present in the MLDC challenge data set. These sources, and the five sources in the MLDC training data, were correctly identified in the sense that the full signal-to-noise ratio of the injected source was recovered, and a posterior distribution for the parameters obtained. The maximum likelihood and maximum a-posteriori parameters were not particularly close to the true parameters of the injected signals, but this was a consequence of the intrinsic degeneracies in the cosmic string model parameter space and in all cases the true parameters were consistent with the recovered posterior distributions. In controlled studies, we found that the SNR threshold required for detection of the cosmic string bursts was $\\sim 7$--$11$, depending on the burst parameters. Bursts with a low break-frequency require a higher SNR to detect than those with high break frequencies. We also explored the detection of sine-Gaussian bursts and in that case the SNR required for detection was slightly higher, being typically $8$--$11$, with sources having frequency close to Nyquist being more difficult to detect. \\MN~is designed to evaluate the evidence of the data under a certain hypothesis, and this can be used to compare possible models for the burst sources. LISA may detect bursts from several different sources, and it is important for scientific interpretation that the nature of the burst be correctly identified. We used the Bayesian evidence as a tool to choose between two different models for a LISA burst source --- the cosmic string model and the sine-Gaussian model, which was chosen to represent a generic burst. The Bayesian evidence works very well as a discriminator between these two models. The evidence ratio begins to clearly favour the correct model over the alternative at the same SNR that the sources become loud enough to detect in the first place. The usefulness of \\MN~as a search tool in this problem is a further illustration of the potential utility of this algorithm for LISA data analysis, as previously demonstrated in a search for non-spinning SMBH binaries~\\cite{MNnospin}. Other algorithms based on Markov Chain Monte Carlo techniques have also been applied to the search for cosmic strings~\\cite{CornishCS}. Both approaches performed equally well as search tools in the last round of the MLDC. We are now exploring the application of \\MN~to searches for other LISA sources, including spinning massive black hole binaries and extreme-mass-ratio inspirals. \\MN~was not designed primarily as a search algorithm, but as a tool for evidence evaluation and this work has demonstrated the utility of the Bayesian evidence as a tool for model selection in a LISA context. Other problems where the evidence ratio approach could be applied include choosing between relativity and alternative theories of gravity as explanations for the gravitational waves observed by LISA, or choosing between different models for a gravitational wave background present in the LISA data set. The Bayesian evidence was previously used in a LIGO context as a tool to choose between alternative theories of gravity~\\cite{Veitch:2008wd} and in a LISA context to distinguish a data set containing a source from one containing purely instrumental noise~\\cite{littenberg09}. \\MN~provides a more efficient way to compute the evidence and this should be explored in more detail in the future. In the context of interpretation of LISA burst events, what we have considered here is only part of the picture. We have shown that we are able to correctly choose between two particular models for a burst, and this can easily be extended to include other burst models. However, LISA might also detect bursts from unmodelled sources. In that case, algorithms such as \\MN~which rely on matched filtering would find the best fit parameters within the model space, but a higher intrinsic SNR of the source would be required for detection. In such a situation, we would like to be able to say that the source was probably not from a model of particular type, e.g., not a cosmic string burst. There are several clues which would provide an indication that this was the case. The sine-Gaussian model is sufficiently generic that we would expect it, in general, to provide a better match to unmodelled bursts than the cosmic string model, which has a very specific form. Therefore, we could say that if the evidence ratio favoured the cosmic string model over the sine-Gaussian model it was highly likely that the burst was in fact a cosmic string and not something else. Similarly, if we found that several of the alternative models had almost equal evidence, but the SNR was quite high, it would be indicative that the burst was not described by any of the models. We have seen that at relatively moderate SNRs, when the signal is described by one of the models, the evidence clearly favours the true model over an alternative. If we found that two models gave almost equally good descriptions of the source, it would suggest that the burst was not fully described by either of them. A third clue would come from the shape of the posterior for the source parameters. The cosmic string waveform space contains many degeneracies, but these can be characterised theoretically for a given choice of source parameters. If the signal was not from a cosmic string, we might find that the structure of the posterior was modified. Finally, some techniques have been developed for the Bayesian reconstruction of generic bursts~\\cite{roever09} which could also be applied in a LISA context. The usefulness of these various approaches can be explored further by analysing data sets into which numerical supernovae burst waveforms have been injected. While the necessary mass of the progenitor is probably unphysically high for a supernova to produce a burst in the LISA frequency band, such waveforms provide examples of unmodelled burst signals on which to test analysis techniques. The final LISA analysis will employ a family of burst models to characterize any detected events. The work described here demonstrates that the Bayesian evidence will be a useful tool for choosing between such models, and \\MN~is a useful tool for computing those evidences." }, "0911/0911.5522_arXiv.txt": { "abstract": "Comet 133P/Elst-Pizarro is the first-known and currently best-characterised member of the main-belt comets, a recently-identified class of objects that exhibit cometary activity but which are dynamically indistinguishable from main-belt asteroids. We report here on the results of a multi-year monitoring campaign from 2003 to 2008, and present observations of the return of activity in 2007. We find a pattern of activity consistent with the seasonal activity modulation hypothesis proposed by Hsieh et al. (2004, AJ, 127, 2997). Additionally, recomputation of phase function parameters using data in which 133P was inactive yields new IAU parameters of $H_R=15.49\\pm0.05$~mag and $G_R=0.04\\pm0.05$, and linear parameters of $m_R(1,1,0)=15.80\\pm0.05$~mag and $\\beta=0.041\\pm0.005$~mag~deg$^{-1}$. Comparison between predicted magnitudes using these new parameters and the comet's actual brightnesses during its 2002 and 2007 active periods reveals the presence of unresolved coma during both episodes, on the order of $\\sim$0.20 of the nucleus cross-section in 2002 and $\\sim$0.25 in 2007. Multifilter observations during 133P's 2007 active outburst yield mean nucleus colours of $B-V=0.65\\pm0.03$, $V-R=0.36\\pm0.01$, and $R-I=0.32\\pm0.01$, with no indication of significant rotational variation, and similar colours for the trail. Finally, while 133P's trail appears shorter and weaker in 2007 than in 2002, other measures of activity strength such as dust velocity and coma contamination of nucleus photometry are found to remain approximately constant. We attribute changes in trail strength to the timing of observations and projection effects, thus finding no evidence of any substantial decrease in activity strength between 2002 and 2007. ", "introduction": "Discovered on 1996 August 7 \\citep{els96}, Comet 133P/Elst-Pizarro (also designated 7968 Elst-Pizarro; hereafter 133P) orbits in the main asteroid belt ($a=3.156$~AU, $e=0.165$, $i=1.39^{\\circ}$). It has a Tisserand parameter (with respect to Jupiter) of $T_J=3.184$, while classical comets have $T_J<3$ \\citep{vag73,kre80}. In 2005, two more objects displaying cometary activity that are likewise dynamically indistinguishable from main-belt asteroids were identified: P/2005 U1 (Read) \\citep{rea05} and 176P/LINEAR (also known as asteroid 118401 (1999 RE$_{70}$)) \\citep{hsi06c}. Their discoveries led to the designation of a new cometary class --- the main-belt comets (MBCs) --- among which 133P is also classified \\citep{hsi06b}. A fourth MBC, P/2008 R1 (Garradd), has also since been discovered \\citep{gar08,jew09}. Despite the initial excitement over the discovery of the cometary nature of 133P in 1996, no physical studies or monitoring reports were published in the refereed literature until the comet's activity was re-observed in 2002 \\citep{hsi04,low05}. Consequently, little is known about 133P's active behaviour in that intervening period. Since knowledge of the timing of active episodes can constrain hypotheses concerning the source of the activity, we report the results of our own monitoring campaign, which began following 133P's active outburst in 2002 and which culminated in observations of renewed activity in 133P in 2007. ", "conclusions": "} \\subsection{Monitoring Campaign\\label{monitoring}} For all monitoring observations, individual $R$-band images (aligned on the object's photocenter using linear interpolation) from each night were combined into single composite images (Fig.~\\ref{images_133p}). For reference, we also show composite images from 133P's 2002 active phase \\citep[Figs.~\\ref{images_133p}a--\\ref{images_133p}d;][]{hsi04}. Activity is marginally visible in images from 2007 May 19, 2007 August 18, and 2007 September 12 (Figs.~\\ref{images_133p}p, \\ref{images_133p}r, \\ref{images_133p}s), while the comet's characteristic dust trail is clearly visible in the image from 2007 July 17 (Fig.~\\ref{images_133p}q). We find no evidence of activity in images from 2003 September 22 through 2007 March 21 (Figs.~\\ref{images_133p}e--\\ref{images_133p}o) and from 2008 July 1 (Fig.~\\ref{images_133p}t). In all images, even those obtained while 133P was active, the FWHM of the object's surface brightness profile is consistent with the typical FWHM seeing at the time of night when those images were obtained, implying that little or no coma is present. In Figure~\\ref{actv133p}, we mark the positions where we observed 133P to be active or where others reported it to be active, as well as positions where we observed it to be inactive, on a plan view of its orbit. The figure shows that reports of activity in 133P are approximately confined to the quadrant following perihelion, with the earliest detection of activity occurring shortly before perihelion at a true anomaly of $\\nu\\approx350^{\\circ}$ and the latest detection occurring at $\\nu\\approx90^{\\circ}$. This activity profile is consistent with the hypothesis of seasonal activity modulation described in \\citet{hsi04} and \\citet{hsi06a}, whereby 133P's activity is driven by the sublimation of a localised patch of exposed volatile material confined to either the ``northern'' or ``southern'' hemisphere of the body. Assuming non-zero obliquity, activity then only occurs during the portion of the orbit when that active site receives enough solar heating to drive sublimation, i.e., during that hemisphere's ``summer''. We note that our observations of 133P on 2008 July 1 at the NTT showed it to be inactive despite the object being observed to be active at nearly the same orbital position in 2002. We attribute this discrepancy to a combination of the low signal-to-noise of this observation and the expected extremely weak activity of 133P at that point in its orbit \\citep{hsi04}. \\subsection{Photometric Activity Detection and Measurement\\label{activitydetection}} When no coma is clearly visible for an object, an alternate method for detecting activity is examination of its photometric behaviour: i.e., determining whether it is consistent with an inactive object of a fixed size, or whether it shows anomalous brightening over a certain portion of its orbit. This type of analysis led to the discovery of activity in 95P/(2060) Chiron \\citep{tho88,bus88,mee89,har90}. In applying this approach to 133P, we recall that \\citet{hsi04} originally derived linear and IAU $H$,$G$ phase function solutions for 133P using data taken in 2002 when the object was visibly emitting dust. In the case of that data set, 133P's activity was judged to contribute negligibly to nucleus photometry (as no significant coma was detected) and was thus assumed to affect phase function derivations similarly negligibly. Having since accumulated a substantial set of observations while 133P was entirely inactive, though, we can now assess the validity of this neglect by deriving new phase function solutions and comparing the results to those of \\citet{hsi04}. We caution that, unlike the data used by \\citet{hsi04}, the photometric data used in this follow-up analysis (2003 Sep 22 to 2007 Mar 21) all consist of ``snapshot observations'', which are short sequences of exposures at unknown rotational phases, instead of full lightcurves. This caveat is significant because rotation of the body is expected to cause deviations in measured brightness by as much as 0.2~mag from the comet's true mean brightness at a given time \\citep{hsi04}. Given a sufficiently large data set, however, we expect that the average of these fluctuations will approach zero, allowing us to derive reasonably accurate phase function solutions without necessarily knowing the rotational phase at which each individual photometry point was obtained. Nonetheless, the lack of rotational phase information for our snapshot observations remains a source of uncertainty. We compute the reduced magnitude, $m_R(1,1,\\alpha)$, of 133P at the time of each observation using \\begin{equation} m_R(1,1,\\alpha) = m_{mid}(R,\\Delta,\\alpha) - 5\\log(R\\Delta) \\end{equation} where $R$ is the heliocentric distance of the object in AU, $\\Delta$ is the object's geocentric distance in AU, and $m_{mid}(R,\\Delta,\\alpha)$ is the estimated $R$-band magnitude at the midpoint of the full photometric range of the rotational lightcurve (Table~\\ref{obs_elstpiz}). For observations of full lightcurves, $m_{mid}$ is determined by simply plotting the data and locating the midpoint between the maximum and minimum values of the lightcurve. For snapshot observations, $m_{mid}$ is generally taken to be the mean of the available photometry data with large error bars applied to reflect rotational phase uncertainties, assuming a full possible photometric range of 0.40~mag. We fit reduced magnitude values to both a linear phase function and an IAU phase function, finding best-fit values of $m_R(1,1,0)=15.80\\pm0.07$~mag and $\\beta=0.041\\pm0.005$~mag~deg$^{-1}$ for the linear phase function where \\begin{equation} m_R(1,1,\\alpha) = m_R(1,1,0) + \\beta\\alpha . \\end{equation} We also find best-fit values of $H_R=15.49\\pm0.05$~mag and $G_R=0.04\\pm0.05$ for the IAU phase function as defined in \\citet{bow89}. Photometry obtained at phase angles of $\\alpha<5^{\\circ}$, where an opposition surge effect is expected, are included in the derivation of the IAU phase function but omitted from the derivation of the linear phase function. We plot our best-fit solutions in Figure~\\ref{phaselaws}. A modest amount of scatter around our solutions is present, as expected, but in all cases the deviations from the best-fit phase functions are consistent with expected brightness fluctuations due to 133P's rotation. Due to the uncertainty of the active status of 133P on 2008 July 01 (\\S\\ref{monitoring}), the photometry from that night is plotted but was not included in the computation of the best-fit phase functions. While the slope parameters of both newly-derived functions are consistent with the parameters computed in \\citet{hsi04} ($\\beta=0.044\\pm0.007$~mag~deg$^{-1}$; $G_R=0.026\\pm0.1$), both newly-derived absolute magnitudes are $\\sim$0.2~mag fainter than their previously derived values ($m_R(1,1,0)=15.61\\pm0.01$~mag; $H_R=15.3\\pm0.1$~mag), strongly suggesting that the previously derived parameters were affected by contamination by 133P's dust emission. This contamination is assumed to consist of a combination of coma and the portion of the dust trail (as projected in the plane of the sky) contained within the seeing disc. This suggestion of dust contamination is reinforced by Figure~\\ref{phaselaws} where we note that photometry from both 133P's 2002 and 2007 active phases are consistently brighter than expected from our new phase function solutions. Because most of the data points from 2002 and 2007 are mean magnitudes derived from fully sampled lightcurves, brightness fluctuations due to rotation cannot account for the discrepancies. Assuming that the discrepancy between an observed magnitude, $m_{mid}$, and expected magnitude, $m_{exp}$, is due to dust contamination, the scattering surface area of the dust, $A_d$, is given by \\begin{equation} A_d = A_n\\left({A_d\\over A_n}\\right) = A_n\\left(10^{0.4(m_{exp} - m_{mid})} - 1\\right) \\label{eqadust} \\end{equation} where $A_n=\\pi r_e^2=1.13\\times10^7$~m$^2$ is the scattering cross-section of the nucleus \\citep{hsi09}, and albedos of the nucleus and dust are assumed to be equal. Assuming optically thin dust, the total dust mass, $M_d$, can then be estimated from \\begin{equation} M_d \\sim {4\\over 3}\\pi\\rho a_d^3 \\left({A_d\\over\\pi a_d^2}\\right) \\label{eqmdust} \\end{equation} where we adopt typical dust grain radii of $a_d=10$~$\\mu$m and a bulk grain density of $\\rho_d=1300$~kg~m$^{-3}$ \\citep[{\\it cf}.][]{hsi04}. For reference, we also compute $Af\\rho$ \\citep[{\\it cf}.][]{ahe84} for each set of observations where the parameter is given by \\begin{equation} Af\\rho = {(2R\\Delta)^2\\over \\rho} 10^{0.4[m_{\\odot}-m_R(R,\\Delta,0)]} \\label{eqafrho} \\end{equation} where $R$ is in AU, $\\Delta$ is in cm, $\\rho$ is the physical radius in cm of a $4\\farcs0$-radius photometry aperture at the distance of the comet, and $m_R(R,\\Delta,0)$ is the phase-angle-corrected $R$-band magnitude of the comet measured using a $4\\farcs0$-radius aperture, which we calculate using \\begin{equation} m_R(R,\\Delta,0) = m_{mid}(R,\\Delta,\\alpha) + 2.5\\log\\left[(1-G)\\Phi_1(\\alpha)+G\\Phi_2(\\alpha)\\right] \\end{equation} where $\\Phi_1$ and $\\Phi_2$ are given by \\begin{equation} \\Phi_1=\\exp\\left[-3.33\\left(\\tan{\\alpha\\over 2}\\right)^{0.63}\\right] \\end{equation} \\begin{equation} \\Phi_2=\\exp\\left[-1.87\\left(\\tan{\\alpha\\over 2}\\right)^{1.22}\\right] \\end{equation} \\citep{bow89}. Using Equations~\\ref{eqadust}, \\ref{eqmdust}, and \\ref{eqafrho}, we compute $A_d$, $M_d$, and $Af\\rho$ for each set of observations from 2002 and 2007 during which 133P was observed to be active, and tabulate the results in Table~\\ref{dustcontrib}. We find that for data from 2002, dust contamination is approximately constant with a scattering surface area of $\\sim$0.20$A_n$ and a dust mass of $M_d\\sim4\\times10^4$~kg contained within $\\sim$3~arcsec ($\\sim$4500~km in August-November; $\\sim$6500~km in December) photometry apertures. The relatively constant amount of dust over this time period explains why we were able to derive reasonably accurate slope parameters for 133P from our 2002 data despite arriving at incorrect results for the comet's absolute magnitude due to the dust contamination. In data from July 2007, we find that 133P's inferred dust coma has a strength comparable to that observed in 2002, having a scattering surface area equivalent to $\\sim$0.25$A_n$ and a dust mass of $M_d\\sim5\\times10^4$~kg contained within $\\sim$4~arcsec ($\\sim$4700~km) photometry apertures. The slightly larger amount of inferred dust in 2007 could indicate a higher rate of dust production, but could also be due to different viewing geometries, given that 133P was close to opposition when observed in July 2007. At this position, the antisolar vector for 133P points very nearly directly behind the object as seen from Earth, causing more of the dust trail to be located within the seeing disc of the comet as projected on the sky. Given the limitations of our observations, however, we are unable to disentangle this possible projection effect from any intrinsic increase in dust production. Additionally, 133P's apparent brightness could also have been enhanced by an opposition surge effect from the dust in its coma, though we unfortunately lack observational constraints for quantifying this effect. Given these various possible contributing factors to 133P's enhanced brightness on 2007 July 17 and 20, we are unable to determine whether 133P was more active on these dates compared to 2002 August 19 through 2002 November 7. We can conclude, however, that coma contamination is present in nucleus photometry performed for 133P during both observing periods, and that the measured magnitude enhancements suggest at least comparable levels of activity in each case. The remainder of our photometry from 133P's 2007 active phase is derived from incomplete lightcurve information, and as such, coma estimates at these times have much larger uncertainties than at other times. We find no definitive evidence of a coma on 2007 May 19, but find that the inferred comae on 2007 Aug 18 and 2007 Sep 12 are far stronger ($\\sim$0.65$A_n$) than in any other observations, a rather unexpected discovery given the minimal amount of time elapsed since our 2007 July observations. We suggest that the large inferred dust contribution to nucleus photometry in August and September could be at least partly due to geometric effects. As can be seen in Figures~\\ref{images_133p}q-\\ref{images_133p}s, the orientation of the projection of the dust trail appears to change over this period of time. We caution that poor seeing during our August and September observations and the small aperture (1.3~m) of the telescope used to obtain these data mean that the observed morphology (namely, the near-disappearance of the dust trail) cannot be considered entirely reliable. If the observed morphology is believed, however, much of the precipitous increase in 133P's apparent coma strength between July and August could be due to the dust trail becoming almost directly aligned behind the nucleus in August and September, thus becoming unavoidably included within our photometry apertures. We can account for this viewing geometry effect by integrating the scattering surface area of the visible dust trail measured in July data (discussed below; \\S\\ref{dusttrail}) and then assuming that it all falls within the photometry aperture used to measure the nucleus magnitudes in August and September. The net increase in dust scattering surface area implied by photometry between 2007 July 20 and 2007 August 18 is $\\sim$0.40$A_n$. The integrated scattering surface area of the dust trail on 2007 July 20 over the first 30~arcsec from the nucleus (the trail becomes too faint to measure reliably beyond this point), however, is $\\sim$0.20$A_n$, accounting for only about half of the observed increase in dust contamination between July and August. The remainder of the observed increase could be partly due to distant material in the dust trail that was too diffuse to detect in trail form in July data, but nevertheless contributed positively to nucleus photometry when projected directly behind the nucleus in August and September. It seems unlikely, however, that half of the dust in the trail could go undetected in our July data, and as such, we surmise that at least part of the increase must in fact be due to a real increase in dust production, which of course would certainly be plausible at this early stage in 133P's active phase. \\subsection{The Lightcurve Revisited\\label{ltcurverevisit}} \\subsubsection{Search for Rotational Colour Variations\\label{colorvariations}} During our 2007 NTT run when 133P was active, we observed the comet in continuously cycling filters ($VRI$ on 2007 July 17 and $BVRI$ on 2007 July 20). Observations were made in this way to allow us to obtain deep imaging of 133P in multiple filters and also construct simultaneous lightcurves in each filter. These lightcurves then allowed us to search for surface colour inhomogeneities that, for example, may constrain the position of the localised active site hypothesized by \\citet{hsi04}. These lightcurves, phased to a rotational period of $P_{rot}=3.471$~hr \\citep{hsi04}, are plotted in Figure~\\ref{nttltcurves}. To then assess colour variation as a function of rotational phase for each filter pair, we use linear interpolation to obtain the magnitudes of the object in the second filter at times of observations in the first filter and then plot the differences (Fig.~\\ref{nttcolors}), again phased to $P_{rot}=3.471$~hr. We find mean nucleus colours of $B-V=0.65\\pm0.03$~mag, $V-R=0.36\\pm0.01$~mag, and $R-I=0.32\\pm0.01$~mag. These values are somewhat different from the mean colours found for 133P by \\citet{hsi04}, but are within the range of individual values measured in that work. We regard the colour measurements presented here to be more accurate since our repeated multifilter observations of 133P here allowed us to account for both rotational magnitude variations (via lightcurve interpolation) and minor extinction variability (using field stars as references for making differential photometric corrections). The single sets of multifilter observations used to make 133P's previous colour measurements did not permit either of these corrective measures. Upon examining individual colour measurements, we find no conclusive evidence of rotational colour inhomogeneity. We find maximum colour variations of only $\\Delta(B-V)=0.11\\pm0.13$~mag, $\\Delta(V-R)=0.06\\pm0.07$~mag, and $\\Delta(R-I)=0.08\\pm0.08$~mag, where the non-systematic distribution of even these small variations indicates that they are most likely due to ordinary measurement uncertainties. We note that this result does not rule out the possibility that 133P's active area exhibits a different colour signature than inactive surface material. First, the coma that is likely present (\\S\\ref{activitydetection}) should act to obscure colour variations on the nucleus surface, with the precise amount of obscuration varying with rotational phase as the ratio of the nucleus's scattering cross-section to the coma's cross-section changes. Furthermore, under the seasonal heating hypothesis \\citep{hsi04,hsi06a}, the active site is in fact expected to be illuminated by the Sun at all rotational phases when near perihelion (assumed to be close to solstice) when these observations were made. The nucleus orientation at this time allows the active site to receive maximal solar heating but also means that the active site is always in the line of sight as viewed from Earth. We suggest that more favourable conditions for detecting colour inhomogeneities will occur around 133P's next pre-perihelion equinox (likely near $\\nu\\sim270^{\\circ}$). Based on prior observations (\\S\\ref{monitoring}), the nucleus should be largely coma-free over this portion of the orbit, and at equinox, the active site should pass into and out of the line of sight as the nucleus rotates, maximising any colour variations. We therefore encourage additional rotationally-resolved colour measurements of 133P between late 2011 and early 2012. \\subsubsection{Implications for 133P's Pole Orientation} For reference, we remove the estimated dust contamination from both our 2002 and 2007 lightcurve data, and overplot the two sets of lightcurves (Fig.~\\ref{ltcurves0207}). Each of the two sets of data are phased self-consistently to $P_{rot}=3.471$~hr, though given the great difficulty of phasing data together that are separated by almost 5 years to such a short rotational period, the 2002 and 2007 data are simply aligned by eye. Due to the two-peaked nature of 133P's lightcurve, though, there is an ambiguity in performing this alignment. In one case (Fig.~\\ref{ltcurves0207}a), the data can be aligned such that the lightcurve shape and photometric range appear largely unchanged between the two observation epochs. In the second case (Fig.~\\ref{ltcurves0207}b), the data can be aligned such that the photometric range of lightcurve appears to decline to $\\Delta m_R\\sim0.25$~mag in 2007 from $\\Delta m_R\\sim0.35$~mag in 2002. In the latter case, it should be recalled that the coma contribution to the data plotted has already been subtracted, and as such the change in photometric range cannot be attributed to differences in the amount of coma. Unfortunately, due to the incomplete sampling of the lightcurve in 2007, it is not possible to resolve the ambiguity between these two cases. This ambiguity is significant because of the implications of photometric range behaviour for the orientation of 133P's rotational pole. To gain more insight as to how the photometric range of 133P should change depending on pole orientation and observing geometry, we simulate its lightcurve behaviour using the model presented in \\citet{lac07}. We assume a simple prolate ellipsoidal shape for the nucleus of 133P and render it at various observing geometries and rotational phases. At each rotational phase, the light reflected back to the observer is integrated to generate lightcurve points. The 2002 September coma-corrected photometric range for 133P was measured to be $\\Delta m_R=0.35$~mag, and so we use a nucleus axis ratio of $a/b=10^{0.4\\Delta m_R}=1.39$ (it should be noted that this is a lower limit due to the unknown projection angle at the time). We use a Lommel-Seeliger ``lunar'' scattering function \\citep[{\\it cf}.][]{fai05} which has no free parameters and is appropriate for simulating the low albedo \\citep[$p_R=0.05\\pm0.02$;][]{hsi09} surface of 133P. To simplify the geometry, we neglect the small orbital inclination ($i=1.4\\degr$) of 133P and assume that it is coplanar with the Earth (i.e., $i=0\\degr$). The seasonal heating hypothesis implies that 133P is at solstice when close to perihelion, i.e. has a true anomaly at solstice of $\\nu_\\mathrm{sol}\\approx0\\degr$, and also requires that the object have non-zero obliquity ($\\varepsilon\\neq0\\degr$). In principle, $\\nu_\\mathrm{sol}$ could potentially have any value from $\\nu_\\mathrm{sol}\\approx0\\degr$ to $\\nu_\\mathrm{sol}\\approx45\\degr$, since the temperature of the hemisphere where 133P's active site is located will begin to rise due to solar heating before the spin axis direction is actually aligned with the Sun. The seasonal heating hypothesis is inconsistent, however, for pole orientations for which $\\nu_\\mathrm{sol}\\approx90\\degr$, or $\\varepsilon=0\\degr$. We simulate the lightcurve behaviour of 133P for $\\nu_\\mathrm{sol}=0\\degr$, $\\nu_\\mathrm{sol}=40\\degr$ and $\\nu_\\mathrm{sol}=90\\degr$. The first and third pole orientations are limiting cases that are consistent and inconsistent with the seasonal heating hypothesis, respectively. The intermediate geometry, in which solstice is reached approximately half-way through the active portion of the orbit, is meant to test how sensitive we are to the exact longitude of the pole. Because we assume zero orbital inclination, each case sets the ecliptic longitude of pole, and the ecliptic latitude is defined by the choice of obliquity. We simulate obliquities of $\\varepsilon=0\\degr$, $\\varepsilon=10\\degr$, $\\varepsilon=20\\degr$ and $\\varepsilon=30\\degr$. Only $\\varepsilon=0\\degr$ is inconsistent with the seasonal hypothesis. Rendered samples of 133P, where we assume $\\varepsilon=30\\degr$, are shown in Figures~\\ref{ltcurve_nu0}, \\ref{ltcurve_nu40} and \\ref{ltcurve_nu90}. Figure~\\ref{rangevsgeom} shows the expected photometric range in 2002 September and 2007 July for each pole orientation. As expected, the photometric range changes in opposite directions for $\\nu_\\mathrm{sol}=0\\degr$ and $\\nu_\\mathrm{sol}=90\\degr$, whereas the intermediate pole orientation ($\\nu_\\mathrm{sol}=40\\degr$) produces only a small change between the two epochs. The absolute value of $\\Delta m_R$ in the figure depends on the assumed axis ratio and is unimportant in this analysis, in which we are primarily concerned with relative changes. The key feature is the variation of the range between the two epochs. The two possible scenarios indicated by the data ({\\it cf}. Fig.~\\ref{ltcurves0207}) are where (a) both the 2002 and 2007 photometric ranges are similar ($\\Delta m_R\\sim0.35$ mag), or (b) the 2002 photometric range ($\\Delta m_R\\sim 0.35$ mag) is larger than the 2007 range ($\\Delta m_R\\sim0.25$ mag). Inspection of Figure~\\ref{rangevsgeom} shows that the first scenario is consistent with low obliquity ($\\varepsilon\\lesssim10\\degr$) and any of the considered pole orientations. The second scenario is only consistent with a solstice around $\\nu_\\mathrm{sol}=0\\degr$ and significant obliquity ($\\varepsilon\\gtrsim20\\degr$). Both scenarios rule out a pole orientation where $\\nu_\\mathrm{sol}=90\\degr$ if there is also significant obliquity. Clearly, additional and more complete lightcurve observations at different points in 133P's orbit are needed to clarify how 133P's photometric range varies with orbit position, constrain the object's pole orientation, and determine whether the seasonal heating hypothesis remains plausible. Given our current data, we can neither confirm nor reject the plausibility of seasonal activity modulation as described by \\citet{hsi04}. While the pattern of activity of 133P along its orbit appears consistent with the seasonal heating hypothesis, the discovery of an incompatible pole solution could indicate that activity is in fact modulated by factors other than obliquity, {\\it e.g.}, shadowing of the active site by crater walls or other local topographic features. In Figure~\\ref{rangevsorbit}, we use our model to forecast the photometric range behaviour of 133P over 1.5 orbits from its perihelion passage in 2007 August and to its aphelion passage in 2016 January. We plot solutions for four pole positions, two consistent with the seasonal heating hypothesis ( $\\varepsilon=20\\degr$, $\\nu_\\mathrm{sol}=0\\degr$, and $\\varepsilon=20\\degr$, $\\nu_\\mathrm{sol}=40\\degr$) and two inconsistent with that hypothesis ($\\varepsilon=20\\degr$, $\\nu_\\mathrm{sol}=90\\degr$, and $\\varepsilon=0\\degr$). The observability of 133P during this period is also indicated in the figure, and should assist in planning observations that are best-suited for discriminating between the various pole orientations that we consider here. \\subsection{The Dust Trail Revisited\\label{dusttrail}} To produce deep composite images from our 2007 NTT data, we use linear interpolation to shift the multiple images obtained in each filter to align the photocenters of the nucleus in each image, and sum the resulting shifted images. To measure the surface brightness profiles of the dust trail in these composite images, we then rotate the images to make the trail horizontal in the image frames, and measure the net flux in rectangular apertures placed along the length of the trail \\citep[{\\it cf}.][]{hsi04}. The dimensions of these equally-sized apertures are set to lengths (along the direction of the trail) of 5 pixels, and widths (perpendicular to the trail) of 6 pixels (approximately equal to the FWHM of the trail cross-section on each night). The net fluxes in these apertures are then converted to net fluxes per linear arcsec (as measured along the length of the trail) and normalised with respect to the total net flux of the nucleus. We plot the resulting surface brightness profiles for both 2007 Jul 17 and 2007 Jul 20 in Figure~\\ref{trailcolors07}. From these plots, we see that the trail profile does not change significantly between the two nights. We also note that there are minimal differences in the profiles of the trail as observed in different filters, indicating that the colours of the dust along the trail are consistently similar to those of the nucleus. To quantify this observation, we measure the surface brightness of the trail as observed on 2007 July 20 in each filter in a single aperture approximately 5 arcsec (15 pixels, or $\\sim$6000~km) in length and 1 arcsec (3 pixels or $\\sim$1200~km) in width placed along the trail. Seeking to minimise the effect of the nucleus on our trail photometry, we place the nearest edge of this aperture $\\sim$3.0~arcsec from the nucleus photocentre. We find surface brightnesses of $\\Sigma_B=24.88\\pm0.17$~mag~arcsec$^{-2}$, $\\Sigma_V=24.35\\pm0.05$~mag~arcsec$^{-2}$, $\\Sigma_R=24.04\\pm0.05$~mag~arcsec$^{-2}$, and $\\Sigma_I=23.70\\pm0.07$~mag~arcsec$^{-2}$, giving colours of $B-V=0.53\\pm0.18$~mag~arcsec$^{-2}$, $V-R=0.31\\pm0.07$~mag~arcsec$^{-2}$, and $R-I=0.34\\pm0.09$~mag~arcsec$^{-2}$, consistent with the colours of the nucleus found in \\S\\ref{ltcurverevisit}. We also wish to know how trail morphology changes between 133P's 2002 and 2007 active episodes. The most obvious difference between the two observing epochs is that the dust trail of 133P is significantly shorter in our 2007 data than in 2002 (despite composite images from each epoch being of approximately equivalent effective exposure time), extending only $\\sim$30 arcsec from the nucleus in 2007 observations, compared to nearly 3 arcmin in 2002 \\citep{hsi04}. In terms of trail width, the observed mean FWHM of the trail on 2002 Sep 07 over the first 10~arcsec of the trail, as measured from the edge of the nucleus's seeing disc (taken to be 2.5$\\times$ the FWHM seeing), was measured to be $\\theta_o=1\\farcs3$. This observed value corresponds to an intrinsic FWHM of $\\theta_i=0\\farcs9$ ($\\sim$1300~km in the plane of the sky), which is computed using \\begin{equation} \\theta_i = (\\theta_o^2 - \\theta_s^2)^{1/2} \\label{intrinsicwidth} \\end{equation} where the FWHM seeing was $\\theta_s=0\\farcs9$ on 2002 Sep 07. For comparison, the observed FWHM of the trail on 2007 July 17 was $\\theta_o=1\\farcs9$, corresponding to $\\theta_i=1\\farcs3$ ($\\sim$1500~km in the plane of the sky), where $\\theta_s=1\\farcs4$. Given that viewing geometries (parametrized by out-of-plane viewing angles, $\\alpha_{pl}$) in 2002 and 2007 were comparable, we therefore find that the computed intrinsic width of the dust trail is approximately equal in both our 2002 and 2007 observations. As \\citet{hsi04} found the primary factor controlling 133P's trail width to be particle ejection velocity, this result suggests that sublimation took place with comparable intensity in both 2002 and 2007. In order to further compare 133P's activity level in 2002 and 2007, we measure the profile of 133P's trail in $R$-band data from 2002 September 07 using the procedure described above, i.e., using rectangular apertures placed along the length of the trail with lengths of 5 pixels and widths of 6 pixels each. We then compare the resulting profile to the mean $R$-band trail profile from 2007 (Fig.~\\ref{trailcomparison}), finding that the trail is noticeably weaker in 2007 than it was in 2002. The difference in trail strength in 2002 and 2007 could be due to several factors. The simplest explanation is that the activity was actually weaker in 2007 due to depletion of exposed volatile material on 133P by the previous outburst. This explanation, however, is at odds with our findings of comparable dust ejection velocities for the two observing epochs (above), and comparable dust enhancement of the nucleus brightness (\\S\\ref{activitydetection}). A more likely explanation is that by the time our 2007 NTT observations were made, 133P was no more than 4 months into its current active phase, whereas it had been active for about a year by the time it was observed on 2002 September 07. Thus, 133P may simply have not yet reached its peak level of activity by the time we observed it with the NTT in 2007. The position of 133P near opposition on 2007 July 17 and 20 also meant that a dust trail pointed in the antisolar direction would be highly projected in the sky, which would additionally explain why the trail appeared to be so much shorter in 2007 than in 2002." }, "0911/0911.0131_arXiv.txt": { "abstract": "We propose here a robust scheme to infer the physical parameters of compact stars from their f-mode gravitational wave signals. We first show that the frequency and the damping rate of f-mode oscillation of compact stars can be expressed in terms of universal functions of stellar mass and moment of inertia. By employing the universality in the f-mode one can then infer accurate values of the mass, the moment of inertia and the radius of a compact star. In addition, we demonstrate that our new scheme works well for both realistic neutron stars and quark stars, and hence provides a unifying way to infer the physical parameters of compact stars. ", "introduction": "Introduction} Ever since the first prediction of neutron stars (NS) by \\citet{Baade1934a}, NS have become a major test bed of various theories of dense matter and nuclear matter. The extreme high pressure and density achievable inside NS offers an ideal and unique environment to infer the equation of state (EOS) of nuclear matter from astronomical observations. On the other hand, the possible existence of quark stars (QS) would provide a strong support to a new form of matter, namely, the quark matter \\citep{Itoh70, Bodmer71, Witten}. Moreover, the size and the mass of QS could also cast light on the properties of quark matter. Thus, extracting useful information and constraints about nuclear and particle physics from astronomical observations of compact stars like NS and QS has become an active direction in astrophysics \\citep[see, e.g.,][for reviews] {lattimer2007nso,haensel2007nse}. Since the density in the core of a NS (or QS) can be several times the normal nuclear density, the EOS there is rather uncertain. Various physical phenomena including pion condensation, hyperon formation and deconfined quark matter have been considered by theoretical nuclear physicists and sophisticated techniques of many-body calculations have been developed to yield a bunch of EOS intended for the description of the deep interior of compact stars. It is standard practice for astrophysicists to construct compact stars with these EOS for nuclear matter and establish various relationships among different physical quantities which could be obtained from astronomical observations. Motivated by the possibilities of constraining EOS of nuclear matter and extracting physical parameters of compact stars as well, researchers in astrophysics and nuclear physics have been actively seeking for EOS-dependent and EOS-independent relationships by examining the physical characteristics of compact stars constructed with different EOS \\citep[see, e.g.,][]{Lattimer:2001,Bejger:2002p8392,Lattimer:2005p7082, bejger2005cdm,lattimer2007nso,haensel2007nse}. For example, \\citet{Lattimer:2001} compared the structure of compact stars of various kinds and discovered several empirical relationships connecting different physical characteristics of a star, such as the mass, the radius, the moment of inertia, and the mass distribution function as well. Such relationships can be applied to infer the physical attributes of a compact star, including its EOS, from astronomical observations. The main objective of the present paper is to unveil the universality embedded in the pulsation frequencies of non-rotating compact stars and to exploit these findings to infer the physical characteristics of NS or QS (e.g., mass, radius, moment of inertia and EOS) from their oscillation spectra. The pulsations of compact stars are an interesting and popular topic in its own right, for gravitational waves (GW) can be generated in such processes \\citep[see][for a seminal exposition of the topic]{Thorne}. It is indisputable that the detection of GW would be a major milestone in general relativity and astrophysics. Several Earth-based GW interferometric detectors such as LIGO, VIRGO, GEO600 and TAMA300 have been operating. While the current detectors are still not sensitive enough to detect GW directly, interesting upper limits have been placed on either the GW strains or event rate for several potential astrophysical sources \\citep[see, e.g., ][for the results obtained from the latest science runs of LIGO]{LIGO_S4_BHringdown,LIGO_S5_LowMass, LIGO_S5_pulsars}. Owing to the energy carried away by GW, pulsations of compact stars are damped harmonic oscillations, which are analyzed in terms of quasi-normal modes (QNM) \\citep[see, e.g.][]{Press_1971,Leaver_1986,rmp,Kokkotas_rev}. Each QNM is characterized by a complex eigenfrequency $\\omega=\\omega_\\r+ i\\omega_\\i$ and has a time dependence $\\exp(i\\omega t)$, displaying exponential decay with a damping time $\\tau\\equiv 1/\\omega_\\i$. Several attempts have been made to relate $\\omega_\\r$ and $\\omega_\\i$ (or $\\tau$) to the mass $M$ and the radius $R$ of a compact star and some universal relationships which are, to certain degree of accuracy, EOS-independent have been found \\citep{Andersson1996,Andersson1998,Ferrari,Ferrari_prd,Tsui05:1}. However, most of these relationships fail to describe QNM of QS, which are stiff and self-bound. On the other hand, \\citet{Lattimer:2005p7082} found an empirical relation approximately expressing the moment of inertia in terms of the mass and the radius. Such a relationship is universal as long as the EOS under consideration is not too soft (e.g., hyperon matter) or too stiff (e.g., quark matter). They proposed to use the universality to estimate the radius of a compact star from its mass and moment of inertia, with the latter two quantities being measurable for double pulsar systems by considering the spin-orbit coupling effect in general relativity \\citep{Lattimer:2005p7082}. Motivated by the discovery of \\citet{Lattimer:2005p7082}, we propose here to investigate the correspondence between the QNM frequency (both $\\omega_\\r$ and $\\omega_\\i$) of the (fundamental) f-mode oscillations, $M$ and the moment of inertia $I$ of compact stars. We show that there exist EOS-independent universal relationships between them, namely equations (\\ref{lau_r}) and (\\ref{lau_i}), which are accurate up to a few percent or better and work nicely for QS as well. Based on equations (\\ref{lau_r}) and (\\ref{lau_i}), a feasible scheme is in turn established to determine the mass and the moment of inertia of a compact star once its f-mode frequency is found from GW observations. As the radius $R$ of a NS (or QS) is approximately expressible in terms of the mass and the moment of inertia, our scheme can also yield good estimate of the radius and hence imposing constraints on the EOS \\citep{Bejger:2002p8392,Lattimer:2005p7082}. The outline of the paper is as follows. Section 2 is a brief review on the the behavior of f-mode oscillations of compact stars \\citep{Tsui05:1} and the universal relationships among the mass, the radius and the moment of inertia of compact stars \\citep{Bejger:2002p8392,Lattimer:2005p7082}. In Section 3 we introduce an accurate universal relation between the scaled frequency $M \\omega$ and the physical quantity $\\sqrt{M^3/I}$. In Section 4 we establish a feasible and robust scheme to apply our findings reported in Section 3 to invert the mass, the moment of inertia, and in turn the radius of a compact star if its f-mode GW signal is detected. We finally summarize our paper in Section 5 with a brief discussion. Unless otherwise noted, we adopt geometric units with $G=c=1$ throughout the whole paper, and use kilometers as the unit of lengths. ", "conclusions": "Conclusion and discussions} In this paper we study the universality embedded in f-mode pulsations of compact stars and propose to use it to determine the physical parameters (including mass, radius and moment of inertia) of a compact star from which the f-mode gravitational wave signal is detected. We first establish a pair of empirical equations (\\ref{lau_r}) and (\\ref{lau_i}) that can predict the frequency and the damping rate of f-modes from $M$ and $I$ with good accuracy for both NS and QS. We then apply such discovery to develop an inversion scheme to infer the mass, radius and moment of inertia of a compact star from its f-mode. In particular, the scheme is shown to work well for both NS and QS. The universality discovered in the present paper takes moment of inertia into the consideration, replacing the role of radius used in previous attempts \\citep{Andersson1996,Andersson1998,Ferrari,Ferrari_prd,Tsui05:1}. It is worth noting that the radius of a star is sensitive to the low-density part of its EOS, whereas f-mode frequencies are expected to be dominated by dynamical behavior in the high-density regime. It is therefore physically appealing to use moment of inertia, which measures global mass distribution, to study f-mode oscillations. As clearly shown in Table~\\ref{uni_quality}, the replacement of $R$ with $I$ indeed leads to a much improved universal behavior. As it is arguable that the accuracy of the inversion scheme directly reflects the quality of the universal behavior upon which the scheme is based, the universality discovered in the present paper leads to an accurate inversion scheme. Finally, it is worthy of mentioning that the first gravitational-wave search sensitive to the f-modes has recently been carried out by the LIGO detectors \\citep{abbott-2008sgr}. While the current detectors have not yet detected gravitational waves directly, it is plausible that the advanced LIGO detectors, which have a factor of 10 improvement on the sensitivity, might detect the f-mode gravitational wave signals from compact stars. The inversion scheme proposed in this work will provide a unifying way to infer the mass, radius and moment of inertia of both NS and QS accurately. Of course in reality the situation is often complicated by the presence of magnetic field and rotation, in the future we shall consider the impacts of these factors on our proposal." }, "0911/0911.0313_arXiv.txt": { "abstract": "{In this paper we present the first system test in which we demonstrate the concept of using an array of Distributed Read Out Imaging Devices (DROIDs) for optical photon detection.} {After the successful S-Cam 3 detector the next step in the development of a cryogenic optical photon counting imaging spectrometer under the S-Cam project is to increase the field of view using DROIDs. With this modification the field of view of the camera has been increased by a factor of 5 in area, while keeping the number of readout channels the same.} {The test has been performed using the flexible S-Cam 3 system and exchanging the 10x12 Superconducting Tunnel Junction array for a 3x20 DROID array. The extra data reduction needed with DROIDs is performed offline.} {We show that, although the responsivity (number of tunnelled quasiparticles per unit of absorbed photon energy, e-/eV) of the current array is too low for direct astronomical applications, the imaging quality is already good enough for pattern detection, and will improve further with increasing responsivity.} {The obtained knowledge can be used to optimise the system for the use of DROIDs.} ", "introduction": "With the S-Cam project the Advanced Studies \\& Technology Preparation Division of the European Space Agency is developing a series of prototype cryogenic detectors to be used as optical photon counting imaging spectrometers for ground-based astronomy. S-Cam uses Superconducting Tunnel Junctions (STJs) (\\cite{Friedrich:2006}; \\cite{Prober:2006}; \\cite{Peacock:1996}) as its detector technology. The merit of this and other cryogenic detectors (\\cite{Romani:1999}) is that they combine single photon detection with sub-microsecond time resolution and intrinsic wavelength resolution, imaging and good detection efficiency in a single device. STJs consist of 2 superconducting layers separated by a thin insulating layer acting as a tunnel barrier. With the absorption of a photon in the superconducting layer a large quantity (several thousands) of Cooper pairs are broken into quasiparticles which can tunnel across the barrier and, under the influence of an applied bias voltage, produce a measurable current pulse. The number of created quasiparticles is given by: $N(E_{0})=\\frac{E_{0}}{\\varepsilon}$, with $N(E_{0})$ the number of created quasiparticles, $E_{0}$ the energy of the absorbed photon and $\\varepsilon=1.75\\Delta_{g}$ the mean energy needed to create a quasiparticle (\\cite{Kurakado:1981}) with $\\Delta_{g}$ the energy gap of the superconducting material. As shown the number of created quasiparticles, and hence the amplitude of corresponding tunnel current, is proportional to the energy of the absorbed photon, thus providing the detector with its spectrographic capabilities. The theoretical limit for the intrinsic energy resolution is given by: $\\Delta E= 2.355\\sqrt{\\varepsilon E_{0} (F+G)}$, where $F$ is the Fano factor (\\cite{Fano:1947}), equal to $F=0.2$ (\\cite{Kurakado:1981}; \\cite{Rando:1991}), and $G=1+\\frac{1}{}$ (\\cite{Mears:1993}) accounts for the statistical variations in the tunnel process, with $$ the average number of tunnels of a single quasiparticle. The energy gap, $\\Delta_{g}$, of the superconducting material is proportional to its critical temperature, $T_{c}$), the temperature at which the phase changes from superconducting to normal metal. For a BCS type superconductor (usually an elemental superconducting material which follows the theory developed by \\cite{Bardeen:1957}), $\\Delta_{g}=1.764k_{b}T_{c}$. A lower energy gap of the superconducting material will therefore increase the number of created quasiparticles and provide better spectrographic capabilities, but it also puts increasing constraints on the operating temperature ($T_{op}$). This needs to be well below the critical temperature of the superconducting layer ($T_{op}\\approx0.1T_{c}$) in order to sufficiently reduce the thermally excited quasiparticle population. For a more extended overview of the STJ technology the reader is referred to \\cite{Peacock:1996}. Each STJ needs to be read out using a dedicated electronics chain which limits the maximum number of pixels that can be read out in a practical application (\\cite{Martin:2006}). To overcome this limitation the Distributed Read Out Imaging Device (DROID) (\\cite{Kraus:1989}) is being developed. A DROID consists of a superconducting absorber strip with STJs on either end, see fig. \\ref{fig:Figure 1}. The photon is absorbed in the absorber and the created quasiparticles diffuse towards the STJs where they tunnel. The sum of the tunnel signals of both STJs is a measure for the energy of the absorbed photon while the ratio is a measure for the absorption position. Depending on the position resolution of the DROID it can replace a number of single STJs and reduce the number of read out channels for a given sensitive area (\\cite{Hijmering:2008}). Within the S-Cam project three prototype cameras have already successfully been used on telescopes such as the William Herschel Telescope (WHT) on La Palma and the Optical Ground Station (OGS) on Tenerife (\\cite{Martin:2004}). S-Cam 1 (\\cite{Verhoeve:2002}) and 2 (\\cite{Rando:2000}) were based on a $6\\times6$ pixel array ($25\\times25 \\mu m^{2}$ pixels) with a wavelength resolving power of 6. S-Cam 3 (\\cite{Martin:2007}; \\cite{Martin:2006}) was based on a $10\\times12$ pixel array ($35\\times35 \\mu m^{2}$ pixels), increasing the field of view on the WHT from $4\\arcsec\\times4\\arcsec$ to $10\\arcsec\\times12\\arcsec$. Also the covered wavelength range, operating temperature and resolving power ($\\sim14@500nm$) has been enhanced with S-Cam 3. The applicability of this type of detector has been proven in different observation campaigns in which several types of astronomical objects have been observed. The high time resolution spectrally resolved S-Cam data has provided strong constraints on the geometry of eclipsing binaries (\\cite{Perryman:2001}; \\cite{de Bruijne1:2002}; \\cite{Martin:2003}). Precise timing of the Crab pulsar light curve has shown that the optical pulses lead the radio pulses by $273\\pm65\\mu s$ (\\cite{Perryman:1999}; \\cite{Oosterbroek:2006}). The spectral information provided by the STJs has enabled the direct determination of quasar redshifts (\\cite{de Bruijne2:2002}) and stellar temperatures (\\cite{Reynolds:2003}). The next step is to increase the field of view further with the use of DROIDs. Here we present the results of the first system test using a $3\\times20$ DROID array as a detector. ", "conclusions": "We have successfully demonstrated operation of an array of DROIDs as a photon counting and imaging detector in an astronomical instrument. Although the responsivity of the array was too low for practical use, the resolving power and imaging capabilities, which will improve further with increasing responsivity, are already adequate. From this first system test we have obtained a good understanding on how to further optimize the system for photon detection with DROIDs." }, "0911/0911.2316_arXiv.txt": { "abstract": "We have resimulated the six galaxy-sized haloes of the Aquarius Project including metal-dependent cooling, star formation and supernova feedback. This allows us to study not only how dark matter haloes respond to galaxy formation, but also how this response is affected by details of halo assembly history. In agreement with previous work, we find baryon condensation to lead to increased dark matter concentration. Dark matter density profiles differ substantially in shape from halo to halo when baryons are included, but in all cases the velocity dispersion decreases monotonically with radius. Some haloes show an approximately constant dark matter velocity anisotropy with $ \\beta \\approx 0.1-02$, while others retain the anisotropy structure of their baryon-free versions. Most of our haloes become approximately oblate in their inner regions, although a few retain the shape of their dissipationless counterparts. Pseudo-phase-space densities are described by a power law in radius of altered slope when baryons are included. The shape and concentration of the dark matter density profiles are not well reproduced by published adiabatic contraction models. The significant spread we find in the density and kinematic structure of our haloes appears related to differences in their formation histories. Such differences already affect the final structure in baryon-free simulations, but they are reinforced by the inclusion of baryons, and new features are produced. The details of galaxy formation need to be better understood before the inner dark matter structure of galaxies can be used to constrain cosmological models or the nature of dark matter. ", "introduction": "Dissipationless cosmological simulations have contributed substantially to our understanding of the structure and evolution of cold dark matter (CDM) haloes, showing them to have triaxial shapes (Frenk et al. 1998; Dubinski \\& Carlberg 1991; Jing \\& Suto 2002; Hayashi, Navarro \\& Springel 2007) and density profiles with inner cusps (Dubinski \\& Carlberg 1991, Navarro, Frenk \\& White 1996; Moore et al. 1999; Diemand et al. 2005) and an approximately universal shape, independent of mass or cosmological parameters (Navarro, Frenk \\& White 1997). This universal cuspy profile appears in conflict with several observations, for example, the slow increase of rotation velocity with radius at the centre of low surface brightness galaxies (Flores \\& Primack 1994; McGaugh \\& de Blok 1998; Moore et al 1999; Salucci, Yegorova \\& Drory 2008) and the relatively weak concentration of galaxy clusters inferred from strong lensing (Sand et al. 2008). Additional evidence supporting central dark matter densities lower than predicted by CDM models comes from the difficulty in simultaneously matching observed luminosity functions and the zero-point of the Tully-Fisher relation (e.g. Dutton, van den Bosch \\& Courteau 2008). Recently, Navarro et al. (2008, hereafter N08) analysed in detail the density profiles of very high-resolution dark matter-only simulations of six Milky Way mass haloes from the Aquarius Project (Springel et al. 2008). They found the innermost cusps in these haloes to be weaker than claimed in some earlier work, and showed how halo-to-halo variations in radial structure reflect the detailed formation histories of individual haloes (see also Vogelsberger et al. 2009). Galaxy formation might reinforce such history-specific features in the dark matter distribution, since baryons are subject to dissipative processes such as cooling, star formation and feedback in addition to gravity. The condensation of baryons within dark matter haloes modifies both their dynamics and their structure, but exactly how this happens is still quite uncertain. The simplest model assumes that the dark halo is compressed radially and adiabatically by the added mass at its centre (Blumenthal et al. 1986, hereafter B86; see Eggen et al. 1962, Zeldovich et al. 1980 and Barnes \\& White 1984 for previous applications of this formalism). Recent simulations have found the very simple adiabatic compression (AC) scheme of B86 to overestimate the effect (e.g. Gnedin et al. 2004 and Abadi et al. 2009, hereafter G04 and A09, respectively; Pedrosa, Tissera \\& Scannapieco 2010) but it is nevertheless often used in simplified modelling of galaxy formation (e.g. Mo, Mao \\& White 1998). More sophisticated AC models, based, for example, on the formalism of Young (1980) have claimed to reproduce better the contraction and the final shape of haloes (G04; Sellwood \\& McGaugh 2005). The common outcome of these schemes is nevertheless an increase in the dark matter density in the central regions, exacerbating the problem of reconciling $\\Lambda$CDM models with observation. Simulations of galaxy formation which follow both dark matter and baryons in their proper cosmological context are the primary tool for studying how galaxy assembly affects dark matter haloes. There are many papers devoted to this subject. Improving numerical algorithms and increasing computer power have led to continual progress in understanding the complex interplay between these two components. Previous analyses of the evolution of dark haloes in hydrodynamical simulations have reported an increase both of the central mass concentration and of the central velocity dispersion, which no longer shows the 'temperature inversion' characteristic of dissipationless CDM haloes (e.g. Katz \\& Gunn 1991; Evrard, Summers \\& Davis 1994; Navarro \\& White 1994; Tissera \\& Dominguez-Tenreiro 1998; O\\~norbe et al. 2008; Romano-Diaz et al. 2008). Recently, Pedrosa, Tissera \\& Scannapieco (2009, 2010) investigated how baryonic assembly history affects the dark matter distribution. They found the SN feedback process to play a key role by regulating star formation activity and ejecting material not only from the main system, but also from infalling satellites. It seems clear that the final distribution of baryons at the centre of a halo is insufficient to determine halo response to galaxy assembly, thus contradicting the AC hypothesis. Typical dark matter haloes have long been reported to be triaxial with major to minor axis ratios often exceeding two in CDM-only simulations (e.g. Barnes \\& Efstathiou 1987; Frenk et al. 1988; Dubinski \\& Carlberg 1991; Jing \\& Suto 2002; Hayashi, Navarro \\& Springel 2007). However, their shape changes, becoming more nearly oblate as baryons condense within them (Katz \\& Gunn 1991; Evrard et al. 1994; Tissera \\& Dom\\'{\\i}nguez-Teneiro 1998; Kazantzidis et al 2004; Debattista et al. 2008; A09). These results suggest the need for a comprehensive analysis of halo structure in simulations which include both the appropriate cosmological context and a realistic description of the physics of baryon condensation. It is clearly necessary to analyse a number of different galaxy-sized haloes in order to explore how differing assembly histories affect the structure of the final haloes. In this paper, we study a set of high-resolution resimulations of the six galaxy-mass haloes of the Aquarius Project (Springel et al. 2008). These were carried out with a version of {\\small GADGET-3} which includes a multi-phase treatment of metal-dependent cooling, star formation and SN feedback (Scannapieco et al. 2005, 2006). The original haloes were selected from a cosmological CDM-only simulation with no restriction on merger history, except that implied by eliminating objects with high-mass close neighbours. This Aquarius halo set is well-suited to study how formation history and baryonic condensation together determine the final structure of dark matter haloes. Our ability to isolate these effects is aided by comparing results from our hydrodynamical simulations (hereafter SPH runs) with corresponding results from the CDM-only simulations (hereafter DM runs) as reported by N08. Properties of the galaxies in these haloes are studied in Scannapieco et al. (2009) and will be further analysed in a forthcoming paper. This paper is organized as follows. In Section 2, we describe the numerical experiments. In Section 3, we analyse the dark matter density profiles. Section 4, describes the velocity dispersion structure of our haloes. In Section 5, we study the effects of baryons on the pseudo-phase-space density profile, while Section 6 discusses halo shapes. Section 7 compares the change in halo circular velocity profile between the SPH and the DM runs with the predictions of AC models. Finally, in Section 8, we summarize our main findings. ", "conclusions": "We have analysed the dark matter distributions in six galaxy-sized haloes belonging to the Aquarius Project, comparing results from the original dark matter only simulations to those from re-simulations including baryonic processes (S09). In agreement with previous work, we find that dark matter haloes become more concentrated when baryons condense at their centres, but that the characteristics of the contraction do not correlate in a simple way with the total amount of baryons. Our main contribution here is to analyse similar mass haloes with a variety of formation histories using high resolution simulations with and without baryons. Our results show that the response of haloes to the presence of baryons is sensitive to the details of halo assembly. This set of six galaxy-sized haloes provides the opportunity to study which properties can be considered common to such haloes and which depend significantly on their particular formation history. Our findings can be summarized as follows: \\begin{itemize} \\item In the regions dominated by baryons, haloes become significantly more concentrated than their dark matter only counterparts. The level of concentration varies significantly from object to object, however, it is not simply related to the total baryonic mass accumulated in the central galaxy. For $r < r_{-2}$, the dark matter density profiles of many of our simulations are nearly isothermal except, possibly, very close to the centre. \\item The velocity dispersion structure is modified in all haloes, with velocity dispersion increasing monotonically to small radii in all cases. The temperature inversion of NFW profiles is no longer present once the galaxy has formed. In some systems, the tangential dispersion increases more than the radial dispersion (but not all), causing them to become more nearly isotropic. \\item The $\\beta$-$\\gamma$ relation proposed by Hansen \\& Moore (2006) is obeyed at most over a restricted radial range for some of our haloes. It works best in the central region of those systems where the isothermal behaviour extends over a relatively small range. The departures from their predictions become large when the profile is isothermal or has constant anisotropy over an extended radial range. \\item Pseudo-phase-space density no longer follows the same power law in radius as in Bertschinger's (1985) similarity solution once galaxy formation is included. The profile differs from one halo to the next and in no halo is it consistent with the purely adiabatic contraction of the dark matter only case. \\item As in previous work, the condensation of baryons makes the central regions of all our haloes less aspherical and more nearly oblate, although in two cases the changes are small. One of these has no significant disk component and the other has a bulge with substantial nett rotation. \\item None of the simple adiabatic contraction models proposed in earlier work is able to describe how the radial density profiles of our haloes are modified by baryon condensation. The scheme suggested by Abadi et al (2009; see also Pedrosa et al. (2010)) is the most successful of those we consider, although it can significantly over- or underestimate the effects in individual cases. For the same haloes, the central mass is overestimated by more than $50\\%$. The circular velocity curves of our galaxy formation simulations are not as centrally peaked as predicted by such adiabatic contraction models, but are still more rapidly rising than in many observed spirals. Since none of our simulated galaxies is disk-dominated it is unclear whether this is a problem. \\end{itemize} Our analysis show that haloes of similar virial mass respond in different ways to baryon condensation depending on the details of their assembly history. Consequently, a much more detailed understanding of galaxy formation is needed before we can make reliable predictions for the detailed structure of the dark matter haloes surrounding galaxies." }, "0911/0911.0852_arXiv.txt": { "abstract": "We present archival high spatial resolution VLA and VLBA data of the nuclei of seven of the nearest and brightest Seyfert galaxies in the Southern Hemisphere. At VLA resolution ($\\sim 0.1$ arcsec), the nucleus of the Seyfert galaxies is unresolved, with the exception of MCG-5-23-16 and NGC\\,7469 showing a core-jet structure. Three Seyfert nuclei are surrounded by diffuse radio emission related to star-forming regions. VLBA observations with parsec-scale resolution pointed out that in MRK\\,1239 the nucleus is clearly resolved in two components separated by $\\sim$ 30 pc, while the nucleus of NGC\\,3783 is unresolved. Further comparison between VLA and VLBA data of these two sources shows that the flux density at parsec scales is only 20\\% of that measured by the VLA. This suggests that the radio emission is not concentrated in a single central component, as in elliptical radio galaxies, and an additional low-surface brightness component must be present. A comparison between Seyfert nuclei with different radio spectra points out that the ``presence'' of undetected flux on milli-arcsecond scale is common in steep-spectrum objects, while in flat-spectrum objects essentially all the radio emission is recovered. In the steep-spectrum objects, the nature of this ``missing'' flux is likely due to non-thermal AGN-related radiation, perhaps from a jet that gets disrupted in Seyfert galaxies because of the denser environment of their spiral hosts. ", "introduction": "Only a small fraction ($\\sim$ 10\\%) of the population of active galactic nuclei (AGN) possess a powerful radio emission ($L_{\\rm 1.4\\,GHz} > 10^{23}$ W/Hz), as found in radio galaxies/quasars and blazars. Seyfert galaxies are part of the ``radio-quiet'' AGN population, with radio luminosity $L_{\\rm 1.4\\,GHz} \\leq 10^{20 - 23}$ W/Hz. Despite their weak radio emission, Seyfert nuclei are very nearby, allowing the study in detail of the radio properties of their central engine.\\\\ Radio observations with arcsecond resolution of several Seyfert samples \\citep[see e.g.][]{ulvestad84, morganti99, thean00} showed that a large fraction of Seyferts have resolved structures, with hint of jets and/or extended emission, the latter usually related to star-forming regions. Several objects, such as NGC\\,1052 \\citep{wrobler84}, NGC\\,1068 \\citep{ulvestad87}, NGC\\,7674 \\citep{momjian03}, and MRK\\,3 \\citep{kukula99} have been found to display a radio morphology with core, collimated jets and hot spots, similar to those found in radio-loud galaxies. However, powerful radio sources have linear structures reaching hundreds of kpc or even Mpc scales, while in Seyferts the radio emission is confined to a few kpc or even sub-kpc scales.\\\\ When observed with milli-arcsecond resolution, the pc-scale structure of Seyfert nuclei is usually resolved in several components \\citep[e.g. NGC\\,3079,][]{trotter98}, resembling a jet structure \\citep[e.g. NGC\\,4151,][]{ulvestad98, nagar01} and sometimes with the presence of extended emission \\citep[e.g. NGC\\,5793,][]{haghi00}. The comparison between arcsec and milli-arcsec radio properties pinpointed a frequent misalignment between pc and kpc-scale jets, suggesting either a change in jet ejection axis, or a bending due to pressure gradients in the ambient medium \\citep{middelberg04}. \\\\ An intriguing characteristic shown by a large number of Seyfert nuclei is that the radio emission arising from their pc-scale structure is often much fainter than that derived from observations with lower resolution, even in the case the nucleus is unresolved. This result suggests that in Seyfert nuclei the radio emission is not concentrated in the central region, as found in powerful radio galaxies, but it extends on scales of tens or hundreds pc \\citep[see e.g.][]{sadler95}. However, not all the Seyfert nuclei have missing flux on parsec scales, as in the case of MRK\\,530 \\citep{lal04}, indicating that the radio emission mainly arises from the central compact component, without evidence of extended, low-surface brightness features.\\\\ In this paper, we present the results of multi-frequency archival VLA and/or VLBA data of a sample of some of the nearest and brightest Seyfert galaxies, taken from the infrared high spatial resolution studies conducted by \\citet{prieto09} and \\citet{reunanen09}. For these sources, no complete and/or unambiguous information on their radio properties could be found in the literature. The comparison between these data with those at pc-scale resolution available from the literature allows a better determination of the physical conditions of the radio emission at kpc and pc scales. \\begin{figure*} \\begin{center} \\special{psfile=mn09_1328_fig1a.eps voffset=-175 hoffset=0 vscale=80 hscale=80} \\special{psfile=mn09_1328_fig1b.eps voffset=-175 hoffset=180 vscale=80 hscale=80} \\special{psfile=mn09_1328_fig1c.eps voffset=-175 hoffset=360 vscale=80 hscale=80} \\special{psfile=mn09_1328_fig1d.eps voffset=-345 hoffset=0 vscale=80 hscale=80} \\special{psfile=mn09_1328_fig1e.eps voffset=-345 hoffset=180 vscale=80 hscale=80} \\special{psfile=mn09_1328_fig1f.eps voffset=-345 hoffset=360 vscale=80 hscale=80} \\special{psfile=mn09_1328_fig1g.eps voffset=-515 hoffset=0 vscale=80 hscale=80} \\vspace{18cm} \\caption{VLT-VISIR images at 11.8 $\\mu$m of the Seyfert nuclei studied in this paper. Images adapted from \\citet{reunanen09}. Contours are at 3$\\sigma$, 5, 7, 11 and 19 levels. The triangle in NGC\\,7582 represents the ionization cone.} \\label{ir} \\end{center} \\end{figure*} Throughout this paper, we assume the following cosmology: $H_{0} = 71\\, {\\rm km\\, s^{-1}\\, Mpc^{-1}}$, $\\Omega_{\\rm M} = 0.27$, and $\\Omega_{\\Lambda} = 0.73$, in a flat Universe. The spectral index is defined as $S {\\rm (\\nu)} \\propto \\nu^{- \\alpha}$. ", "conclusions": "We presented the analysis of multi-frequency archival VLA and VLBA data of 7 close and bright Seyfert nuclei. The conclusions we can draw from this investigation are:\\\\ \\begin{itemize} \\item At VLA resolution, FWHM $\\sim 0^{\\prime\\prime}.1$, the nucleus of the Seyfert galaxies is unresolved, with the only exception of MGC-5-23-16 and NGC\\,7469 which show a core-jet structure. At VLBA resolution, equivalent to a few parsecs, the nucleus of MRK\\,1239 is resolved in two components, while in NGC\\,3783 it is still unresolved. \\\\ \\item The Seyfert galaxies in this study with known circumnuclear star-forming regions in the IR, namely NGC\\,1097, NGC\\,7469 and NGC\\,7582 \\citep{reunanen09} present a radio counterpart for a few ($\\sim$10\\%) of these regions. Most are not detected in radio, indicating that those detected may be characterised by higher supernovae rate. Their steep radio spectral index is in line with this idea. Conversely, none of the other Seyfert nuclei present any circumnuclear radio emission. The only exception is NGC\\,5506 that shows a radio halo surrounding the nucleus. \\\\ \\item A comparison between arcsecond and milli-arsecond resolution in MRK\\,1239 and NGC\\,3783, pointed out that almost 80\\% of the radio flux density detected in VLA observations is not recovered at pc-scale resolution. This suggests the presence of a diffuse component on scales of a few tens of parsecs, undetected with the VLBA. Similar situation is found in NGC\\,1068, NGC\\,4151, NGC\\,5506 and NGC\\,7469. The nature of this ``missing'' flux components is likely due to synchrotron AGN-related emission. This difference between the flux density measured on arcsecond and milli-arcsecond resolution images is not found in the case of elliptical radio galaxies, but it appears to be a common phenomenon in Seyfert galaxies with a steep spectrum, mostly hosted in spirals. If of synchrotron origin, this emission may be spilt off from a jet that may get easily distorted and/or disrupted by the dense interstellar medium in the nucleus of spirals.\\\\ \\item A comparison between Seyfert nuclei with different spectral properties points out that in flat-spectrum nuclei, almost all the flux density is recovered on milli-arcsecond scale. This indicates that in flat-spectrum objects the radio emission is essentially concentrated in the compact core, without evidence of jet-like structure even on milli-arcsecond scale, while in steep-spectrum objects a significant fraction of the radio emission arises from low-surface brightness, extended features. \\end{itemize}" }, "0911/0911.2585_arXiv.txt": { "abstract": "A solution of the large discrepancy existing between inclusive and exclusive measurements of the ${}^8{\\rm Li}+{}^4{\\rm He}\\to{}^{11}{\\rm B}+n$ reaction cross section at $E_{cm} <3$ MeV is evaluated. This problem has profound astrophysical relevance for this reaction is of great interest in Big-Bang and r-process nucleosynthesis. By means of a novel technique, a comprehensive study of all existing ${}^8{\\rm Li}+{}^4{\\rm He}\\to{}^{11}{\\rm B}+n$ cross section data is carried out, setting up a consistent picture in which all the inclusive measurements provide the reliable value of the cross section. New unambiguous signatures of the strong branch pattern non-uniformities, near the threshold of higher ${}^{11}{\\rm B}$ excited levels, are presented and their possible origin, in terms of the cluster structure of the involved excited states of ${}^{11}{\\rm B}$ and ${}^{12}{\\rm B}$ nuclei, is discussed. ", "introduction": " ", "conclusions": "" }, "0911/0911.4195_arXiv.txt": { "abstract": "A set of HI sources extracted from the north Galactic polar region by the ongoing ALFALFA survey has properties that are consistent with the interpretation that they are associated with isolated minihalos in the outskirts of the Local Group (LG). Unlike objects detected by previous surveys, such as the Compact High Velocity Clouds of Braun \\& Burton (1999), the HI clouds found by ALFALFA do not violate any structural requirements or halo scaling laws of the $\\Lambda$CDM structure paradigm, nor would they have been detected by extant HI surveys of nearby galaxy groups other than the LG. At a distance of $d$ Mpc, their HI masses range between $5\\times 10^4d^2$ and $10^6d^2$ \\msun ~and their HI radii between $<0.4d$ and $1.6d$ kpc. If they are parts of gravitationally bound halos, the total masses would be on order of $10^8$--$10^9$ \\msun, their baryonic content would be signifcantly smaller than the cosmic fraction of 0.16 and present in a ionized gas phase of mass well exceeding that of the neutral phase. This study does not however prove that the minihalo interpretation is unique. Among possible alternatives would be that the clouds are shreds of the Leading Arm of the Magellanic Stream. ", "introduction": " ", "conclusions": "" }, "0911/0911.1315_arXiv.txt": { "abstract": "WMAP5 and related data have greatly restricted the range of acceptable cosmologies, by providing precise likelihood ellypses on the the $w_0$--$w_a$ plane. We discuss first how such ellypses can be numerically rebuilt, and present then a map of constant--$w$ models whose spectra, at various redshift, are expected to coincide with acceptable models within $\\sim 1\\, \\%$. ", "introduction": "\\label{intro} One of the main puzzles of cosmology is why a model as $\\Lambda$CDM, implying so many conceptual problems, apparently fits all linear data in such unrivalled fashion \\citep{astier,riess,komatsu}. It is then important that the fine tuning paradox of $\\Lambda$CDM is eased in cosmologies where Dark Energy (DE) is a self--interacting scalar field $\\phi$ (dDE cosmologies), with no likelihood downgrade \\citep{colombo,lavacca}. Unfortunately, however, only cosmological observables can provide information on the form of the self--interaction potential, even though several researchers incline to priviledge potentials allowing tracking solutions. When aiming to obtain information on the physical potential, the basic observable is however the evolution of the DE scale parameter, $w(a)$. Here $a = 1/(1+z)$ is the scale factor of a spatially flat metric \\begin{equation} ds^2 = c^2 dt^2 - a^2(t) d\\ell^2~, \\label{metric} \\end{equation} while $z$ is the redshift. The analysis of available data, made by the WMAP team \\citep{komatsu}, was able to constrain the coefficients $w_0$ and $w_a$ in the expression \\begin{equation} w(a) = w_0 + (1-a)\\, w_a~, \\label{wa} \\end{equation} putting again into evidence that a model with $w \\equiv -1$ is close to top likelihood, but also stressing a preference for the phantom area ($w < -1$), which is hardly consistent with current tracking potentials. The main tool, to go beyond these constraints on the $w(a)$ law, will be tomographic shear analyses \\citep{refregier}, able to reconstruct the spectrum of density fluctuations at various $z$'s, with a precision approaching 1$\\, \\%$ \\citep{huterer}. This work is therefore focused on the relevance of spectral predictions, and aims at providing a tool to ease the determination of model spectra. Within this context, it is essential to outline the spectral equivalence criterion (SEC). It has been noted since a few years \\citep{white,linderW} that the density fluctuation spectra $P(k) = \\langle |\\delta \\rho/\\rho |^2 \\rangle$, up to $k=3\\, h$Mpc$^{-1}$, essentially depend on the distance from the LSB (Last Scattering Band). This was first verified at $z=0$ \\citep{linder} and then tested at higher $z$ \\citep{CaMaB}. $N$--body simulations show that the SEC is however true when model parameters are suitably tuned. Let us be more specific on this point: When a given $w(a)$ is considered, it is easy to determine the comoving distance of the LSB. Keeping then the same values of $\\Omega_{b,c}$ and $h$, we can seek a constant--$w$ cosmology with the same distance from the LSB; \\citep{linder} tested that the spectra of these two cosmologies coincide (within $1\\, \\%$). They also saw that, when $z \\neq 0$ values are explored, spectral discrepancies are mostly greater, in the range of a few percents. \\begin{figure}% \\vskip -2.4truecm \\begin{center} \\includegraphics[width=10.cm]{wmap5.w.eps} \\end{center} \\vskip -.6truecm \\caption{1-- and 2--$\\sigma$ curves, yielding the marginalized model likelihood on the $w_0$--$w_a$ plane, as obtained from the Lambda NASA site. The reproduction device is detailed in the text. Also 0.5-- and 1.5--$\\sigma$ curves are provided. Crosses indicates models for which the SEC was explicitly tested at various $z$'s. } \\vskip -.3truecm \\label{elly} \\end{figure} The SEC however works also at $z \\neq 0$, provided that the distance between such $z$ and the LSB is evaluated and a constant--$w$ auxiliary model is defined, with an equal distance between $z$ and the LSB. The assigned cosmology and the auxiliary model must also have equal $\\Omega_{b,c}$, $h$ {\\it at $z=0$}, while $\\sigma_8(z=0)$ must be tuned in order that, at $z$, the {\\it r.m.s.} density fluctuations of the two models, on the $8\\, h^{-1}$Mpc scale, coincide \\citep{CaMaB}. Let us outline, in particular, that the SEC does not require that the auxiliary model shares the values of $\\Omega_{b,c}$ and $h$ at the assigned $z \\neq 0$. In fact, by multiplying the critical density definition by $\\Omega_{c,b}$, one has \\begin{equation} \\omega_{c,b} \\propto \\Omega_{c,b} H^2 = (8\\pi G/3) \\Omega_{c,b} \\rho_{cr} = (8\\pi G/3) \\rho_{c,b} \\propto a^{-3} \\end{equation} so that the assigned cosmology and the auxiliary model share the values of $\\omega_{c,b}$ at any $z$, and this is enough. This comes as no surprise, however: most linear feature, {\\it e.g.} BAO's, essentially depend just on $\\omega_{b,c}$. This recipe, however, prescribes a different constant--$w$ at any $z$ and is somehow curious that, starting from the assigned $w(a)$, one builds a $w_{eff}(a)$ law, completely different from it. In a similar way, one can build a $\\sigma_{8;(z=0)}$ dependence from $z$, in order that auxiliary models be fairly normalized at any $z$. Examples of $w_{eff}(z) $ and $\\sigma_{8;(z=0)}(z)$ are provided by \\citep{CaMaB}. The scope of this work, instead, amounts to applying the SEC to models with $w(a)$ of the form (\\ref{wa}) and consistent with data, exploring decreasing likelihood ellypsoids. The plan of the paper is as follows: In the next Section we shall discuss how one can rebuild the likelihood ellypsoids on the $w_0$--$w_a$ plane. This Section defines the parameter $t$ used in the plots which are one of the results of this work. In Section 3 we shall then discuss such plots. Finally, Section 4 will be devoted to drawing our conclusions. ", "conclusions": "\\label{conclusions} When tomographic cosmic shear data will become available, an inspection on DE nature will surely start from comparing them with the predictions of constant--$w$ cosmologies. As soon as data will become more refined, it will be possible to bin them, discriminating among the $w$ values best fitting data at various redshift. It is quite possible that these inspections yield $w$ values compatible with a constant, all through the redshift range explored. As it is possible that such value is compatible with $-1$, so vanifying the efforts to improve our knowledge on DE nature. Let us suppose that, instead, data analysis is consistent with the same $\\omega_{b,c}$ values in all bins, but require different $w$'s in different bins. \\begin{figure}% \\begin{center} \\includegraphics[width=10cm]{tw2.s1.eps} \\end{center} \\caption{Bias in observational $w$ values (named here $w_{eff}$), in respect to the physical value of the state parameter, at each redshift. The points in the curves of the upper frame refer to different models at 1--$\\sigma$ from the top--likelihood cosmology, discriminated~by different $t$ values. In the lower frame the ratio $w_{eff}/w$ is explicitly shown as a function of $t$~.} \\label{tw21} \\end{figure} \\begin{figure}% \\begin{center} \\includegraphics[width=10cm]{tw2.s2.eps} \\end{center} \\caption{As Figure \\ref{tw21} for models at 2--$\\sigma$ from the top--likelihood cosmology.} \\label{tw22} \\end{figure} We wish to outline here a major danger that data analysis could meet in this welcome case. As a matter of fact, one could be tempted to conclude that the $w(z)$ dependence found is the physical scale dependence of the DE state parameter. Unfortunately, this could be badly untrue. In fact, cosmic shear spectra can be easily translated into fluctuation spectra, so that observational values of $w(z)$ would correspond to $w_{eff}(z)$, not to the physical $w(z).$ How different the two behaviors can be is already represented by Figures 2--5 in this paper. But Figures \\ref{tw21} and \\ref{tw22} further illustrate this point. They show what is $w_{eff}$ when $w$ has a given value, for models laying on the 1--$\\sigma$ or 2--$\\sigma$ curves. Different colors in each plot correspond to different $z$'s. In the lower frame, the $w_{eff}/w$ ratio is also shown, as a function of $t$. Even in the most furtunate cases, provided that the {\\it true} cosmology is not a constant--$w$ one, discrepancies are hardly below 10$\\, \\%$ and are greatest at $z$ approaching zero. Another point concerning model fitting is that a systematic mapping of (reasonable) dDE cosmologies, in order to fit future data, is apparently unnecessary. At each $z$, in fact, there will be a constant--$w$ model able to fit any dDE cosmology. dDE will then be revealed by the $w$ dependence on $z$, as above outlined. At present, well approximated analytical expressions of $P(k)$, at different $z$, are provided by the so--called {\\it Halofit} formulae, holding for $\\Lambda$CDM cosmologies \\citep{smith}. Some attempt to generalize {\\it Halofit} to constant $w \\neq -1$ were also performed \\citep{mcdonald}, but they do not cover the desired parameter ranges. Our conclusion is that it will be important to extend {\\it Halofit} to constant--$w$ cosmologies, for the whole range of (reasonable) cosmological parameters; this will enable us to fit future cosmic shear data without any substantial restriction on the $w(a)$ behavior." }, "0911/0911.1409_arXiv.txt": { "abstract": "We observed CO $J=3-2$ emission from the ``water fountain\" sources, which exhibit high-velocity collimated stellar jets traced by \\h2o maser emission, with the Atacama Submillimeter Telescope Experiment (ASTE) 10~m telescope. We detected the CO emission from two sources, \\i1634 and \\j1828. The \\i1634 CO emission exhibits a spectrum that is well fit to a Gaussian profile, rather than to a parabolic profile, with a velocity width (FWHM) of $158 \\pm 6$ \\kms \\ and an intensity peak at \\vlsr$=50 \\pm2$ \\kms. The mass loss rate of the star is estimated to be $ \\sim 2.9\\times 10^{-5}M_{ \\odot}$~yr$^{-1}$. Our morpho-kinematic models suggest that the CO emission is optically thin and associated with a bipolar outflow rather than with a (cold and relatively small) torus. The \\j1828 CO emission has a velocity width (FWHM) of $3.0 \\pm 0.2$ \\kms, smaller than typically seen in AGB envelopes. The narrow velocity width of the CO emission suggests that it originates from either an interstellar molecular cloud or a slowly-rotating circumstellar envelope that harbors the \\h2o maser source. ", "introduction": "The process in which mass outflow shapes planetary nebulae (PNe) is still unresolved and needs to be elucidated. One of the striking features of some PNe is the bipolarity in their morphological and kinematical structures, some of which clearly exhibit high velocity bipolar jets ($V_{\\mbox jet}>$100 \\kms). Such bipolar jets have been found at the beginning of the post asymptotic giant branch (AGB) phase and even at the end of the AGB phase (e.g. \\cite{sah98,ima02,ima07b}). In particular, a small fraction of stellar jets have velocities higher than a typical expansion velocity of a circumstellar envelope (CSE) found in 1612~MHz OH maser emission (\\vjet $ \\gtrsim$30\\kms), and are traced by \\h2o \\ maser emission. Such high velocity jets are called ``water fountains\" \\citep{ima07b}. As of 2009, 13 such objects have been identified \\citep{wal09,sua09}. Some spatio-kinematical structures of the jets have been elucidated by very long baseline interferometry (VLBI) observations of the \\h2o masers \\citep{ima02,ima05,ima07,ima07b, bob07,cla09}. Interestingly, all of the estimated apparent dynamical ages of the jets are about 100~yr or shorter, which is consistent with the rare detection of water fountain sources. This implies that these jets should have just been launched at the final evolutionary stage of the dying stars. Note that the \\h2o maser emission is excited when the bipolar tips of the stellar jet strike into the ambient CSE that was produced from the previous spherically symmetric stellar mass loss. Therefore, in order to more reliably estimate the stellar mass loss rate, the dynamical age, and the whole morphological and kinematical structure of the jet, a mapping observation of thermal emission such as CO emission is crucial. The first detection of CO emission from a water fountain source was reported with the Arizona Radio Observatory 10~m telescope towards \\i1634 \\citep{he08}. Usually it is difficult to detect such CO emission towards the water fountain sources because they are located close to the Galactic plane with heavy contamination from the interstellar CO emission or they are too distant ($D \\gtrsim$2~kpc). Here we present results of the first systematic CO $J=3-2$ emission observations towards the water fountain sources with the Atacama Submillimeter Telescope Experiment (ASTE) 10~m telescope. In addition, we also show results of CO $J=1-0$ emission observations towards W~43A and \\j1828 with the Nobeyama 45~m telescope. In section 2 and 3, we describe in detail the observations and results, respectively. In section 4, we discuss the mass loss rate and the morphology and kinematics of \\i1634, whose CO $J=3-2$ emission is detected. ", "conclusions": "Using ASTE, we have surveyed CO $J=3-2$ emission from nine water fountain sources and detected the CO emission towards I16342 and \\j1828. I16342 may have a mass loss rate of $ \\dot{M}_{ \\mbox{ \\scriptsize{gas}}} \\approx 5.6\\times10^{-5} \\:M_{ \\odot}$~yr$^{-1}$, with most emission attributed to a bipolar outflow rather than a high velocity expanding torus or a biconical flow with a wide opening angle. On the other hand, to further clarify the intrinsic CO detection from \\j1828 with a very narrow velocity width ($\\sim 3$\\kms), we need interferometric observations of this object. Because the water fountain sources are located at large distances, higher sensitivity is essential in future observations. \\bigskip We deeply appreciate members of the ASTE team for their careful observations, preparation and kind operation support. The ASTE project is driven by Nobeyama Radio Observatory (NRO), a branch of National Astronomical Observatory of Japan (NAOJ), in collaboration with University of Chile, and Japanese institutes including University of Tokyo, Nagoya University, Osaka Prefecture University, Ibaraki University, and Hokkaido University. Observations with ASTE were in part carried out remotely from Japan by using NTT's GEMnet2 and its partner R\\&E (Research and Education) networks, which are based on AccessNova collaboration of University of Chile, NTT Laboratories, and NAOJ. We also acknowledge the referee, Albert Zijlstra, for careful reading and providing several fruitful comments for paper improvement. HI and SD have been financially supported by Grant-in-Aid for Scientific Research from Japan Society for Promotion Science (20540234). JN was supported by the Research Grants Council of the Hong Kong under grants HKU703308P, and by the financial support from the Seed Funding Programme for Basic Research in HKU (200802159006)." }, "0911/0911.1912_arXiv.txt": { "abstract": "\\normalsize The inner 10~pc of our galaxy contains many counterpart candidates of the very high energy (VHE; $> 100$~GeV) \\gr\\ point source \\hgcfull. Within the point spread function of the \\hess\\ measurement, at least three objects are capable of accelerating particles to very high energies and beyond, and of providing the observed \\gr\\ flux. Previous attempts to address this source confusion were hampered by the fact that the projected distances between those objects were of the order of the error circle radius of the emission centroid (34'', dominated by the pointing uncertainty of the \\hess\\ instrument). Here we present \\hess\\ data of the Galactic Centre region, recorded with an improved control of the instrument pointing compared to \\hess\\ standard pointing procedures. Stars observed during \\gr\\ observations by optical guiding cameras mounted on each \\hess\\ telescope are used for off-line pointing calibration, thereby decreasing the systematic pointing uncertainties from 20'' to 6'' per axis. The position of \\hgc\\ is obtained by fitting a multi-Gaussian profile to the background-subtracted \\gr\\ count map. A spatial comparison of the best-fit position of \\hgc\\ with the position and morphology of candidate counterparts is performed. The position is, within a total error circle radius of 13'', coincident with the position of the supermassive black hole \\astar\\ and the recently discovered pulsar wind nebula candidate \\pwn. It is significantly displaced from the centroid of the supernova remnant \\aeast, excluding this object with high probability as the dominant source of the VHE \\gr\\ emission. \\vspace{5mm}\\\\ Key words: Galaxy: centre -- ISM: individual: Sgr~A~East -- ISM: individual: Sgr~A* -- ISM: individual: G~359.95-0.04 -- gamma-rays: observations ", "introduction": "\\label{Introduction} Since the discovery of the strong compact radio source \\astar\\ \\citep{Balick:1974aa}, the Galactic Centre (GC), as the closest galactic nucleus, has served as a unique laboratory for investigating the astrophysics of galactic nuclei in general. The radio picture \\citep{LaRosa00} of the central few 100~pc around the centre of the Milky Way exhibits a complex and very active region, with numerous sources of non-thermal radiation, making this region a prime target for observations at very high energies (VHE; $> 100$~GeV). Indeed several Imaging Atmospheric Cherenkov Telescopes (IACTs) have detected a source of VHE \\grs\\ in the direction of the GC \\citep{Aharonian:2004wa,Albert:2005kh,Kosack:2004ri,Tsuchiya:2004wv}. The \\hess\\ instrument \\citep[see][and references therein]{crab} provides the to date most precise VHE data on this source, henceforth called \\hgcfull. As shown with deep observations in 2004, \\hgc\\ is a point source for \\hess\\ (rms spatial extension $< 1\\farcm 2$ at 95\\% CL), and is within $7''\\pm 14''_{\\mathrm{stat}}\\pm 28''_{\\mathrm{sys}}$ positionally coincident with the bright radio source \\astar\\ \\citep{Aharonian:2006wh}. The measured energy spectrum does not fit Dark Matter (DM) model spectra -- at least for the most popular models of DM annihilation --, ruling out the bulk of the TeV emission soley to be of a DM origin \\citep{Aharonian:2006wh}. Of all possible astrophysics counterparts, the $3\\times 10^6\\ \\mathrm{M}_{\\astrosun}$ supermassive black hole (SMBH) coincident with the \\astar\\ radio position is a compelling candidate. Various models predict VHE emission from this object, produced either close to the SMBH itself \\citep{Aharonian:2005ti}, within an ${\\cal O}(10)$~pc zone around \\astar\\ due to the interaction of run-away protons with the ambient medium \\citep{Aharonian:2005b,Liu2006a,Wang:2009}, or by electrons accelerated in termination shocks driven by winds emerging from within a couple of Schwarzschild radii \\citep{Atoyan2004}. \\astar\\ is a source of bright and frequent X-ray and infrared flares. Detection of quasi-periodic oscillations (QPOs) on time scales of 100-2250~s has been claimed \\citep[e.g.][]{Baganoff2001,Genzel:2003,Porquet2003}. Recently, however, observations with the Keck II telescope could not confirm the existence of such QPOs \\citep{Meyer:2008}. No hint for variability, flaring activity, or QPOs has been found in the VHE \\gr\\ lightcurve in 93~h live time of \\hess\\ data collected during the years 2004-2006 \\citep{GCSpectrum}. Moreover, during a campaign of simultaneous \\hess\\ and Chandra observations of \\astar\\ in 2005, a major X-ray flare of 1600~s duration was observed. Although the X-ray flux increased to $\\approx 9$ times the quiescent level, no evidence for flaring activity was detected in the VHE lightcurve \\citep{Aharonian:2008yb}. This result makes it highly unlikely that X-ray and VHE emission originate from the same source region, and puts constraints on models predicting correlated flaring. Besides \\astar\\ and its immediate vicinity, there are at least two other production site candidates for VHE emission. The first one is the radio-bright, shell-like supernova remnant (SNR) \\aeast, which surrounds partially \\astar. SNRs have been shown to be efficient particle accelerators \\citep[see e.g.][]{Helder:2009fm}, and the presence of an ${\\cal O}$(mG) magnetic field \\citep{Yusef96} makes \\aeast\\ a compelling candidate for particle acceleration to very high energies \\citep{Crocker2005}. The second one is the recently detected pulsar wind nebula (PWN) candidate \\pwn\\ \\citep{Wang:2005ya}. Despite of its faint X-ray flux, it may plausibly emit TeV \\grs\\ at an energy flux level compatible with the \\hess\\ observations \\citep{Hinton:2006zk}, assuming that \\pwn\\ is located at the same distance as \\astar. A firm identification of \\hgc\\ is particularly hampered by the -- compared to radio or X-ray instruments -- modest angular resolution of the current generation of Cherenkov telescopes ($\\leq 5'$ for a single \\gr\\ at TeV energies), which gives rise to source confusion in this densely populated region of the galaxy. Adopting a distance to the GC of 8.33~kpc \\citep{Gillessen:2009}, the \\hess\\ source size upper limit encloses a region of about 2.9~pc radius. Comparing this number to the projected distance of \\astar\\ to the radio maximum of \\aeast\\ and the X-ray maximum of \\pwn\\ (3.7~pc and 0.4~pc, respectively), it becomes clear that a precise position measurement of the centre-of-gravity of \\hgc\\ can help to shed light on the nature of this source. Although previous \\hess\\ position measurements have been unprecedentedly precise, the relatively large -- compared to statistical errors -- systematic errors due to pointing uncertainties of the \\hess\\ array rendered the identification of the major contributing source of the VHE emission, and especially a clear statement on the role of \\aeast, difficult. In this paper a refined measurement of \\hgc's emission centroid is reported. Using improved telescope pointing control, the systematic error of the measurement is decreased by a factor of three compared to previous results, and the total error on the centroid position is reduced to $13''$ (68\\% containment radius), compared to $34''$ in \\cite{Aharonian:2006wh}. ", "conclusions": "\\begin{figure} \\includegraphics[width=0.49\\textwidth]{./figures/Fig2.eps} \\caption{90~cm VLA radio flux density map \\citep{LaRosa00} of the innermost 20~pc of the GC, showing emission from the SNR \\aeast. Black contours denote radio flux levels of 2, 4, and 6~Jy/beam. The centre of the SNR \\citep{Green:2009qf} is marked by the white square, and the positions of \\astar\\ \\citep{saga_radio} and \\pwn\\ \\citep[head position,][]{Wang:2005ya} are given by the cross hairs and the black triangle, respectively. The 68\\% CL total error contour of the best-fit centroid position of \\hgc\\ is given by the white circle. The dashed white circle shows the same contour for the previously reported \\hess\\ measurement \\citep{Aharonian:2006wh}. The white and black dashed-dotted lines show the 95\\% CL upper limit contour of the source extension and the 68\\% containment region of the \\hess\\ PSF, respectively. The white stars marked \\emph{A} and \\emph{B} denote the position of the radio maximum and the best-fit position for the radio emission after smoothing with the PSF of the \\hess\\ instrument, respectively. } \\label{fig:RadioMap} \\end{figure} Fig.~\\ref{fig:RadioMap} shows a VLA 90~cm image of the innermost 20~pc region of the GC, centred on \\astar. The shell-like radio structure of the SNR \\aeast\\ is clearly visible. The best-fit position of \\hgc\\ lies in a region where the radio emission is comparatively low, and is shown as a 68\\%~CL total error contour, computed from the summed (in quadrature) statistical and systematic best-fit position errors. As can be seen from the figure, the centroid of the VHE source is coincident with the positions of \\astar\\ and \\pwn, but inconsistent with the regions of intense radio emission from \\aeast. Two rather independent approaches to derive a quantitative statement about the compatibility of \\hgc's best-fit position with \\aeast\\ are detailed in the following. The white star labelled $A$ in Fig.~\\ref{fig:RadioMap} denotes the position of \\aeast's radio maximum. Comparing the 68\\%~CL radius of the observed VHE centroid to the angular distance between $A$ and the best-fit position, a coincidence of the two positions is ruled out with 7.1 standard deviations. By the same arguments, VHE point emission from the centre of the SNR \\citep{Green:2009qf}, indicated as a white square in Fig.~\\ref{fig:RadioMap}, is ruled out with 4.7 standard deviations. Instead of point emission, extended emission can be considered, e.g. by assuming that the hypothetical VHE emission from \\aeast\\ follows closely the morphology of the radio flux. The centroid of such emission would be detected at the coordinates marked $B$ in Fig.~\\ref{fig:RadioMap}. This position was derived by fitting the radio map -- smoothed with the \\hess\\ PSF -- with the technique used above for the VHE \\gr\\ data. Following the methods used for position $A$, the radio fit position is 5.4 standard deviations away from the best-fit VHE centroid position. The above results are obtained assuming that the VHE emission and radio morphology are correlated. Since the best-fit position does not coincide with a region of intense radio emission (see Fig.~ \\ref{fig:RadioMap}), relaxing this assumption leads to more conservative estimates of the association probability. A priori, it would appear conservative to assume that the centroid of TeV emission associated with \\aeast\\ might appear anywhere within the boundaries, with equal probability. With this assumption one can calculate the probability that the VHE emission is produced inside \\aeast, but is only by chance positionally coincident with \\astar\\ and \\pwn, which themselves are plausible emitters of VHE radiation and thus viable counterpart candidates of \\hgc. Defining the 2~Jy/beam radio contour of \\aeast\\ as the SNR boundary, which encloses the best-fit position of the emission centroid, a chance probability of $9\\times 10^{-5}$ is derived (corresponding to 3.9 standard deviations). This number does slightly change depending on the assumed size of the SNR boundary. It is clear, however, that even with this conservative approach an association of \\aeast\\ with the observed VHE \\gr\\ emission is rather unlikely. The exclusion of \\aeast\\ as the main contributor to the VHE emission is a major step towards an identification of \\hgc. Despite the fact that \\hgc\\ is a non-variable \\gr\\ source, both \\pwn\\ and \\astar\\ are compelling counterpart candidates, as models exist (see section \\ref{Introduction}), which can explain a steady $\\gamma$-ray flux and variable X-ray emission from \\astar\\ at the same time. More information is needed to discriminate between these two objects. Due to their enhanced angular resolution and sensitivity, and their extended energy range, proposed future VHE \\gr\\ observatories such as CTA or AGIS could shed light on the open question of which of these sources dominates the production of the VHE \\gr\\ emission from the gravitational centre of our galaxy." }, "0911/0911.2700_arXiv.txt": { "abstract": "We make an inventory of the baryonic and gravitating mass in structures ranging from the smallest galaxies to rich clusters of galaxies. We find that the fraction of baryons converted to stars reaches a maximum between ${M}_{500} = 10^{12}$ and $10^{13}\\;\\mathrm{M}_{\\sun}$, suggesting that star formation is most efficient in bright galaxies in groups. The fraction of baryons detected in all forms deviates monotonically from the cosmic baryon fraction as a function of mass. On the largest scales of clusters, most of the expected baryons are detected, while in the smallest dwarf galaxies, fewer than 1\\% are detected. Where these missing baryons reside is unclear. ", "introduction": "The early universe was a highly uniform and homogeneous mix of dark and baryonic matter with baryon fraction ${f_b = 0.17 \\pm 0.01}$ \\citep{WMAP5} . If this primordial mix persists as individual gravitationally bound structures emerge during the course of cosmic evolution, then the baryonic mass of any given object would be $M_b = f_b M_{tot}$. Indeed, combining this with the baryon density constraint from big bang nucleosynthesis \\citep{BBN} provides one important argument that the density parameter is less than unity \\citep{White93,OS}. Here we reverse the logic and ask what fraction of the expected baryons are actually detected. We present an inventory of the baryonic and gravitating masses of cosmic structures spanning a dozen decades in detected baryonic mass. The mass of baryons known in each system correlates well with the total mass, but not as a simple proportion. The implication is that most of the baryons associated with individual dark matter halos are now missing. While some are in the intergalactic medium \\citep{OVI}, a complete accounting of where the baryons now reside, and how they relate to their parent structures, remains wanting. ", "conclusions": "The stellar fraction reaches a maximum between ${M}_{500} = 10^{12}$ and $10^{13}\\;\\mathrm{M}_{\\sun}$ (Fig.~\\ref{fdfstV}). This is broadly consistent with previous results based on counting statistics \\citep{yang}. However, there may be some offset in the peak mass scale relating to the long standing dichotomy between Tully-Fisher and luminosity function based normalizations of the halo mass function. Notably, the efficiency of star formation appears to increase monotonically with mass for individual galaxies \\citep{baldry}. After the transition to cluster halos containing many galaxies, the efficiency declines again. Perhaps the most striking aspect of Fig.~\\ref{fdfstV} is that the fraction of detected baryons falls short of the cosmic fraction at all scales. In no system is it unity, as would be expected if we had a complete accounting of all the baryons associated with a given bound system. Where are all these missing baryons? An obvious possibility is that the missing baryons are present, and simply inhabit their dark matter halos in some undetected form. Indeed, one might expect that not all baryons would have time to cool into the observed cold gas and stellar component of galaxies. In this case, many baryons might remain mixed in with the dark halo. The notion that we are simply not seeing many or even most of the baryons in individual galaxies is profoundly unsatisfactory. Direct searches for hot halo gas have turned up nothing substantial: halo baryon reservoirs fail to explain the observed deficit by two orders of magnitude \\citep{bregman,andbreg}. In clusters, the hot gas is detected, and constitutes the majority of the baryons. Clusters fall short of the cosmic baryon fraction by a modest amount \\citep{mccarthy,giodini} which might be readily explicable \\citep{crain}. However, the fraction of missing baryons is highly significant on the scales of individual galaxies. In dwarfs, fewer than 1\\% of the baryons expected from the cosmic fraction are detected. The detected baryon fraction varies systematically with scale. It is a matter of taste whether one chooses to describe this scale as one of circular velocity, mass, or potential well depth. It is tempting to attribute this correlation to feedback processes being more effective in objects with smaller potential wells \\citep{DS}. However, the details are heinously complicated \\citep{MayerMoore,crazy} and not well understood. We would naively expect feedback to be a messy process resulting in lots of scatter in any correlations that might result. Instead, the observed relation is remarkably tight. In principle, the entire range $0 \\le f_d \\le 1$ is accessible at each mass, yet only a very particular value is observed. Moreover, the current potential well is the result of the hierarchical assembly of many smaller building blocks. Left to itself, a small dark matter halo will have a small $f_d$. If incorporated into a larger halo, $f_d$ goes up. How does each building block know to bring the right amount of baryons to the final halo? We do not see a satisfactory solution to this missing baryon problem at present. Considerable work remains to be done to obtain a complete understanding of the universe and its contents." }, "0911/0911.2470_arXiv.txt": { "abstract": "Classical Cepheid variable stars have been important indicators of extragalactic distance and Galactic evolution for over a century. The \\spitzer{} Space Telescope has opened the possibility of extending the study of Cepheids into the mid- and far-infrared, where interstellar extinction is reduced. We have obtained photometry from images of a sample of Galactic Cepheids with the IRAC and MIPS instruments on \\spitzer. Here we present the first mid-infrared period--luminosity relations for Classical Cepheids in the Galaxy, and the first ever Cepheid period--luminosity relations at 24 and 70~\\micron. We compare these relations with theoretical predictions, and with period--luminosity relations obtained in recent studies of the Large Magellanic Cloud. We find a significant period--color relation for the $[3.6]-[8.0]$ IRAC color. Other mid-infrared colors for both Cepheids and non-variable supergiants are strongly affected by variable molecular spectral features, in particular deep CO absorption bands. We do not find strong evidence for mid-infrared excess caused by warm ($\\sim 500$~K) circumstellar dust. We discuss the possibility that recent detections with near-infrared interferometers of circumstellar shells around $\\delta$~Cep, $\\ell$~Car, Polaris, Y~Oph and RS~Pup may be a signature of shocked gas emission in a dust-poor wind associated to pulsation-driven mass loss. ", "introduction": "Although it was over 100 years ago that Henrietta Leavitt discovered the Cepheid Period--Luminosity relation (PL, \\citealt{leavitt1908}), or ``Leavitt Law'', few tools in astronomy have such enduring importance. Classical Cepheid variable stars are fundamental calibrators of the extragalactic distance scale and in addition their observed properties are a benchmark for stellar evolution models of intermediate mass stars. It is now widely accepted that distances derived from SN Ia are still significantly influenced by the classical Cepheid calibration: \\citet{riess2005} found a change of 15\\% in the value of $H_0$ when only modern, high-quality SN Ia data and HST Cepheids were used. Recent works have extended optical and near-infrared PL and period--color (PC) relations to the mid-infrared (mid-IR), where there is less interstellar extinction, with observations of Large Magellanic Cloud (LMC) Cepheids (\\citealt{freedman2008, ngeow2008, ngeow2009, madore2009a}) with the \\spitzer{} Space Telescope \\citep{werner2004}. Evolutionary and pulsation properties of intermediate-mass stars in the core He-burning phase, like Cepheids, play a crucial role in several long-standing astrophysical problems. They are transition objects between stellar structures ending up their evolution either as white dwarfs or as core collapse supernovae. Therefore, they are not only the most popular primary distance indicators, but are also crucial to understanding the chemical evolution of stellar systems hosting a substantial fraction of young stars, e.g the Galactic disk and the dwarf irregular galaxies in the Local Group and in the Local Volume ($d \\la 10$~Mpc). Although these objects are fundamental for stellar evolution \\citep{bono2000, beaulieu2001}, stellar pulsation \\citep{bono1999, marconi2005}, and Galactic chemical evolution models \\citep{pedicelli2009, spitoni2009}, current predictions are still hampered by several problems. The most outstanding issue is the ``Cepheid mass discrepancy'' between pulsation masses of classical Cepheids and their evolutionary masses. Evidence was brought forward more than 30 years ago by \\citet{fricke1972} who found that pulsation masses were from 1.5 to 2 times smaller than the evolutionary masses. This conundrum was partially solved \\citep{moskalik1992} by the new sets of radiative opacities released by the Opacity Project \\citep{seaton1994} and by OPAL \\citep{rogers1992}. However, several recent investigations focused on Galactic Cepheids \\citep{bono2001a, caputo2005} and Magellanic Cloud Cepheids \\citep{beaulieu2001, bono2002, keller2006} suggest that such a discrepancy still amounts to 10--15\\%. Measured masses of Galactic binary Cepheids (e.g. \\citealt{evans2008}) are also smaller than predicted by evolutionary models neglecting core convective overshooting during central hydrogen burning phases. The relative importance of the main factors affecting Cepheid model mass estimates (extra-mixing, rotation, radiative opacity, mass loss and binarity) is still debated. Even though commonly used semi-empirical relations \\citep{reimers1975, dejager1997} do not predict enough mass loss to solve the Cepheid mass discrepancy problem, mass loss may indeed be the key culprit among the physical mechanisms suggested to explain the mass discrepancy problem. The semi-empirical mass loss relation derived by \\citet{reimers1975} is clearly inadequate to correctly estimate the mass loss for certain evolutionary phases of giant stars (see e.g. \\citealt{willson2000}). A plausible increase in the typically adopted Reimers wind free parameter ($0.2 \\le \\eta \\le 0.4$) does not account for the entire range in mass covered by cluster horizontal branch (central helium burning) stars \\citep{yong2000, castellani2005, serenelli2005}. There is no good reason why a similar discrepancy should not affect intermediate-mass stars. Mass loss is often betrayed by the presence of dust shells, detectable as infrared excesses, or by stellar winds which produce blue-shifted absorption dips in the ultraviolet. Empirical estimates of mass loss rates based on infrared (IRAS) and ultraviolet (IUE spectra) observations for a large sample of Galactic Cepheids suggest mass-loss rates ranging from 10$^{-10}$ to 10$^{-7}$~\\msun~yr$^{-1}$ \\citep{deasy1988}. However, evidence for mass loss rates high enough to affect evolution is very rare, and it is not clear that mass loss is a wide-spread phenomenon. \\citet{mcalary1986} used IRAS photometry \\citep{beichman1985} and found evidence of very cool dust ($T_d \\la 50$~K) around two classical Cepheids (\\rspup{} and SU~Cas) known to be associated with reflection nebul\\ae{}. Due to the very large IRAS beamsize (as large as $\\sim 5$~arcmin), however, it was very difficult to separate local dust emission from ``Galactic cirrus'' background emission. More recently, $K$ band near-IR interferometric observations \\citep{merand2006, merand2007, kervella2006, kervella2008, kervella2009} have detected circumstellar emission around five Classical Cepheids: \\deltacep{}, $\\ell$~Car, Polaris, Y~Oph and RS~Pup. While the nature of the material responsible for this emission remains mysterious, this is a tantalizing suggestion that mass loss activity may be present around these nearby Cepheids. A similar conclusion was also reached by \\citet{neilson2009}, that using OGLE (Optical Gravitational Lensing Experiment; \\citealt{udalski1999}) and \\spitzer{} data for Magellanic Cepheids, found evidence of a wide range of mass loss rates. To investigate this possibility, we have obtained \\spitzer{} observations of a sample of Galactic Classical Cepheids. All stars were observed with both the \\spitzer{} Infrared Array Camera (IRAC, \\citealt{fazio2004}) and the Multiband Infrared Photometer for \\spitzer{} (MIPS, \\citealt{rieke2004}). The aim of the IRAC observations was to characterize Galactic Cepheid colors in the mid-IR, where the emission from the photosphere is still dominant, and search for infrared excess related to warm ($\\sim 500$~K) circumstellar dust. The MIPS observations were intended to investigate the presence of extended emission from cool ($\\la 100$~K) dust, taking advantage of the higher angular resolution of \\spitzer{} (5~arcsec at 24~\\micron). In this paper we present the photometry of our Cepheid sample in the IRAC and MIPS bands, discussing their PL and period-color (PC) relations. The results of our search for extended emission in IRAC and MIPS images will be presented in a separate paper (Barmby et al. in preparation). The criteria for our sample selection are laid out in section~\\ref{sec-sample}, and the observations are described in section~\\ref{sec-obs}. The techniques adopted to measure the source photometry in all IRAC and MIPS bands are discussed in detail in section~\\ref{sec-phot}. In section~\\ref{sec-leavitt} we derive the Leavitt Law and PC relations in the IRAC and MIPS bands, and we compare them with similar relations obtained for the LMC. We study the intrinsic mid-IR colors of Cepheids in section~\\ref{sec-excess}, where we set limits on infrared excess at \\spitzer{} wavelengths. Our results are discussed in section~\\ref{sec-discussion} and summarized in section~\\ref{sec-summary}. ", "conclusions": "\\label{sec-discussion} The reliability of Classical Cepheids as standard candles is of paramount importance for astronomy. The recent advances in infrared space astronomy have shifted the focus of obtaining accurate PL relations to wavelengths longer than the visible, where interstellar extinction is reduced. \\citet{madore2009b} and \\citet{freedman2009} have demonstrated how PL relations obtained at IRAC wavelengths can be effectively used to measure the distance of nearby galaxies. This work is a further step in the characterization of PL relations in all \\spitzer{} photometric bands, including two MIPS bands at 24 and 70~\\micron, by using a sample of Classical Cepheids in the Galaxy. In order to provide a detailed comparison between theory and observations we adopted the large set of nonlinear, convective models computed by \\citet{bono1999, marconi2005} and by \\citet{fiorentino2007}. For Galactic Cepheids we chose a scaled-solar chemical composition (helium, $Y=0.28$; metals, $Z=0.02$) and accounted for fundamental mode pulsators. Moreover, we covered a broad range of stellar masses ($3.5 \\le M/M_\\odot \\le 11.5$), and to account for current uncertainties affecting the size of the helium core \\citep{bono2006} we adopted two different mass-luminosity relations based on canonical and non-canonical (the latter incorporating a range of main sequence core convective overshoot) evolutionary models. Theoretical predictions were transformed into the observational plane using scaled-solar atmospheres based on {\\tt ATLAS9} models \\citep{castelli2003}, and multiplied with the IRAC band-passes. The slopes and zero points of these model Leavitt laws are listed in the last two columns of Table~\\ref{tab-PL}. While the zero points are in general agreement with all our fits, the slopes are significantly shallower than our best fits obtained with both ``new'' and ``old'' distances (the disagreement is however reduced in the fits adopting the ``new'' distances). The PL relations derived with the exclusive use of the astrometric distances (that do not rely on the $p(P)$ relation), are however in excellent agreement with the theoretical relations. This result may suggest that the period dependence of the $p-$factor currently adopted in IRSB distances may still need further refinement. Because of the small number of Cepheids and the limited range of periods for which accurate parallaxes are available, we could not attempt to invert the problem and estimate the $p(P)$ dependence with our data. Comparison between our Galactic PL relations and the relations derived for the LMC by \\citet{madore2009a} and \\citet{ngeow2009} leads to contradictory results. The PL relations obtained with both the ``new'' and ``old'' IRSB distances are steeper than the LMC relations, while the Galactic PL relations we obtain with the astrometric distances are significantly shallower. To add to the confusion, it should be noted that the slopes derived by \\citet{ngeow2009} and \\citet{madore2009a} disagree by more than their respective uncertainties, despite using stars from the same galaxy\\footnote{The difference between the two LMC results can be an issue of crowding and binarity in the \\citet{ngeow2009} sample (selected without individual image inspection, and containing a larger fraction of fainter short period Cepheids) that could result in shallower slopes.}. The contradictory results in our Galactic PL relation slopes shows once more the effects of the systematics introduced by the IRSB distances and their dependence on the $p(P)$ relation. Based on the more reliable astrometric distances we should conclude that the slope of the PL relation could be shallower at Galactic metallicity than in the LMC, but even this result is not statistically significant, due to the larger uncertainty in the PL slope. In conclusion, the insufficient reliability of the IRSB distances, and the small number of stars for which astrometric distances are available, prevents us from resolving the dependence of the PL relations from metallicity. This problem has already been noticed at optical and near-IR wavelengths by other authors (see e.g. \\citealt{storm2004, fouque2007, romaniello2008}). \\citet{riess2009} also noted that the precision of current datasets does not allow measurement of the dependence of the PL relation on metallicity; those authors measured statistically consistent slopes at 1.6 micron for the Milky Way, the LMC and other galaxies. We also cannot determine a reliable wavelength dependence of the PL slope, as all our values are within 1$\\sigma$ from the average slope value. This is consistent with the expected flattening of the PL slope wavelength dependence in the mid-IR shown by the models. If we compare our slope values with Figure~4 in \\citet{freedman2008}, however, we note a similar trend. Our result for 4.5~\\micron{} is marginally higher (by $\\sim 0.1$) than the slope at the 3.6 and 8.0~\\micron, as also shown by the LMC Cepheid fit (and by the models). This anomaly in the 4.5~\\micron{} (and, to a minor extent, 5.8~\\micron) slope is related to the presence of the variable CO band, which increases the amplitude variation in these bandpasses. For the first time we have measured the PL relation in the MIPS 24 and 70~\\micron{} bands. The 24~\\micron{} relation, in particular, is important for the determination of Cepheid distances in the Galactic plane from datasets such as MIPSGAL \\citep{carey2009}, where high extinction from ISM dust may complicate their detection at IRAC bands. The determination of a PC relation is in principle not limited by the uncertainty in the distances affecting our PL relations. Figure~\\ref{fig-pc}, however, clearly shows that such a relation can be derived with sufficient accuracy only in the $[3.6]-[8.0]$ color. Other colors show a scatter that is increasingly larger with the period, as high as $\\sim 0.1$~mag, one order of magnitude larger than our color accuracy of $\\sim 0.02$. The fact that this scatter was not found by \\citet{ngeow2009} may again be a consequence of the difference in the median period between our Galactic and their LMC sample. Based on our data and on numerical modeling of one representative Cepheid ($\\zeta$~Gem), we conclude that the color scatter affecting long period Cepheids is a consequence of variable stellar CO absorption, rather than infrared excess from circumstellar dust. Circumstellar dust at $\\sim 500$~K (the temperature where dust thermal emission is maximum in the IRAC bands) would produce the largest excess in the $[3.6]-[8.0]$ color, which we do not observe, with the possible exception of a weak detection ($\\sim 0.1$~mag) around Polaris. Our $\\zeta$~Gem models show that scatter in the IRAC 4.5 and 5.8~\\micron{} bands can be induced by variations in effective temperature and/or the propagation of shocks through the Cepheid atmospheres. A more comprehensive modeling effort, including Cepheids with different periods and pulsation modes, is however required to confirm and quantify this effect. Not finding an infrared excess in most, if not all, stars is somewhat unexpected, given the detection of circumstellar shells by near-IR interferometers. Figure~8 in \\citet{merand2007} shows a number of targets, including the long period Cepheids Y~Oph and $\\ell$~Car (both part of our sample), with a significant (4--5\\%) excess in the K band. We do not measure any excess of this magnitude in these or other stars, which seems to rule out the presence of circumstellar dust in the shells detected around these and other Cepheids. The interferometric determination of the radius of these circumstellar shells, however, indicate values as low as $\\sim 2$ stellar radii \\citep{merand2006}. For a Cepheid with $T_{\\textrm{eff}} \\sim 5$,000--6,000~K, dust equilibrium temperature at $\\sim 2$~R$_*$ would be $T_d \\simeq T_*/\\sqrt{R_d/R_*} \\la 3$,500~K. This is well above the sublimation temperature of any known astronomical dust. Our results can be reconciled with the interferometric detection of circumstellar shells close to the star if the emission is not due to dust, but rather to some strong molecular line emission. Given that the observations in \\citet{merand2006, merand2007} and \\citet{kervella2006} were made in the $K$ band, a possible candidate for this emission is shocked $H_2$. The presence of $H_2$ shocked emission lines could further contribute to the scatter in the IRAC 4.5~\\micron{} band, for certain densities and shock velocities \\citep{smith2006}. This hypothesis supports the idea that Cepheid stars indeed have a strong stellar wind associated with a pulsation driven mass-loss mechanism, as suggested by \\citet{merand2007}. Unlike the case of mass loss in red giants, AGB and supergiant stars (showing a dust-driven wind), this stellar wind would be largely dustless, due to the higher temperature of the star. This hypothesis needs to be tested by near-IR multi-epoch spectral monitoring of these stars. This hypothesis does not imply that a Cepheid stellar wind is completely devoid of dust, that may be condensing at larger radii, in quantities below the detection limits of our IRAC observations (see $\\ell$~Car mid-IR detection of mid-IR circumstellar emission with the VLT MIDI and VISIR instruments, \\citealt{kervella2009}). Cold dust may also be collected by the outflow, at much larger distances (thousands of AU), from the ISM, as in the case of the well known nebula around RS~Pup \\citep{kervella2009}. We find evidence supporting this mechanism in our IRAC and MIPS images. Extended emission at 5.8, 8.0, 24 and 70~\\micron{} around $\\delta$~Cep indicate that this star may be losing mass due to a strong wind pushing into the local interstellar medium, which is leading to the formation of a 70~\\micron{} bow shock detected at large distance from the star ($\\sim 10$,000~AU). A detailed analysis of this phenomenon is being published elsewhere (Marengo et al., in preparation). In a more comprehensive paper \\citep{barmby2009} we will discuss the presence of spatially resolved extended emission at 24 and 70~\\micron{} around more targets, and quantify the occurrence of mass loss in the Cepheid phase that can be inferred from our imaging data. We have derived the PL and PC relations in \\spitzer/IRAC bands for a sample of Galactic Cepheids. These relations are critically dependent on the choice of the period dependence of the $p$-factors used for the distance determination with the IRSB method, even though the uncertainties in our fits prevent the assessment of the best $p(P)$ relation choice. The best agreement with theoretical PL relations, however, is obtained when only distances obtained by astrometric methods are used. We do not detect statistically significant variations between the slope and zero points of the PL relations between our Galactic sample and LMC relations obtained by \\citet{madore2009a}, despite the difference in metallicity. We find that the intrinsic variations in the 4.5 and 5.8~\\micron{} fluxes are larger for long period Cepheids. These variations (of the order of $\\sim 0.1$~mag) are related to deep CO absorption, dependent on the stellar $T_{eff}$ and hydrodynamic effects associated to the stellar pulsations. We do not find significant infrared excess related to warm circumstellar dust, except for a weak excess detected at IRAC wavelengths for Polaris, and at 70~\\micron{} for SZ~Tau. This may rule out the presence of extensive dust driven mass loss in the Cepheid phase, but leaves open the possibility of pulsation-driven mass loss from a dust-poor wind, as suggested by recent interferometric observations." }, "0911/0911.4019.txt": { "abstract": "{}{}{}{}{} % 5 {} token are mandatory \\abstract % % context heading (optional) % % {} leave it empty if necessary % {XXX} % % aims heading (mandatory) % {YYY} % % methods heading (mandatory) % {ZZZ} % % results heading (mandatory) % {ZZZ} % % conclusions heading (optional), leave it empty if necessary {The planetary nebula \\TS\\ (also called \\png\\ or \\sbs), with its record-holding low oxygen abundance and its double degenerate close binary core (period 3.9 h), is an exceptional object located in the Galactic halo. We have secured observational data in a complete wavelength range in order to pin down the abundances of half a dozen elements in the nebula. The abundances are obtained via detailed photoionization modelling taking into account all the observational constraints (including geometry and aperture effects) using the pseudo-3D photoionization code Cloudy$\\_$3D. The spectral energy distribution of the ionizing radiation is taken from appropriate model atmospheres. Incidentally, from the new observational constraints, we find that both stellar components contribute to the ionization: the ``cool'' one provides the bulk of hydrogen ionization, and the ``hot'' one is responsible for the presence of the most highly charged ions, which explains why previous attempts to model the nebula experienced difficulties. The nebular abundances of C, N, O, and Ne are found to be respectively, 1/3.5, 1/4.2, 1/70, and 1/11 of the Solar value, with uncertainties of a factor 2. Thus the extreme O deficiency of this object is confirmed. The abundances of S and Ar are less than 1/30 of Solar. The abundance of He relative to H is 0.089$\\pm$0.009. Standard models of stellar evolution and nucleosynthesis cannot explain the abundance pattern observed in the nebula. To obtain an extreme oxygen deficiency in a star whose progenitor has an initial mass of about 1\\,\\msun\\ requires an additional mixing process, which can be induced by stellar rotation and/or by the presence of the close companion. We have computed a stellar model with initial mass of 1\\,\\msun, appropriate metallicity, and initial rotation of 100\\kms, and find that rotation greatly improves the agreement between the predicted and observed abundances. } ", "introduction": "\\label{sec:introduction} \\sbs\\ was discovered in the second Byurakan Sky Survey and first classified as a cataclysmic variable \\citep{1999PASP..111.1099S}. \\cite{2001A&A...370..456T} discussed in detail the nature of the object and came to the conclusion that, in fact, it is a planetary nebula (PN). The object was renamed \\png, following the nomenclature for Galactic PNe from the Strasbourg-ESO catalogue of Galactic Planetary Nebulae \\citep{1992secg.book.....A}. For the sake of brevity, we will refer to it as \\TS\\ in the rest of the paper. This PN is special in at least three important aspects. First of all, its oxygen abundance is very low, significantly lower than in any other PN known up to now \\citep{2001A&A...370..456T, 2002A&A...395..929R, 2002AJ....124.3340J, 2005A&A...430..187P}. Second, its nucleus is a spectroscopic binary, with a period of only a few hours \\citep{2004ApJ...616..485T}. Third, it appears, from estimates of the nature and masses of the two stellar components, that \\TS\\ could turn into a double degenerate Type Ia Supernova \\citep{2004ApJ...616..485T}. Each of these aspects, even taken alone, makes \\TS\\ an exceptional object. In this paper, we reexamine the chemical composition of \\TS. Briefly, the story of the determination of the chemical composition of this object is the following. Tovmassian et al. (2001) had optical spectra of \\TS\\ in the range 3900--7000\\,\\AA\\ obtained with 2\\,m class telescopes which showed no lines from heavy elements except a very weak \\Oiii, with an intensity a few percent of H$\\beta$. A coarse photoionization analysis suggested an oxygen abundance smaller than 1/100 of Solar. Note that standard empirical methods for abundance determinations in PNe cannot be used for \\TS, since the electron temperature cannot be determined directly from observations. To go further in the abundance determination of \\TS\\ required an estimate of the effective temperature of the central star. One way is to obtain a good blue spectrum of the PN, and use the \\Nev/ \\Neiii\\ ratio (or a limit on it) as a constraint. \\cite{2002A&A...395..929R} at the Canada-France-Hawaii Telescope (CFHT) and \\cite{2002AJ....124.3340J} at the Multiple Mirror Telescope (MMT) secured deep blue spectra in order to detect these lines. \\cite{2002AJ....124.3340J} detected the \\Nev\\ line at a level of ~0.8 H$\\beta$. \\cite{2002A&A...395..929R} found only an upper limit of 0.1 H$\\beta$! Concerning the \\Neiii\\ line, {\\cite{2002AJ....124.3340J} measured an intensity about 10 times larger than \\cite{2002A&A...395..929R}. The two papers appeared within a few days of each other on astro-ph, revealing this big conflict in the observations. The two groups conducted independent photoionization analyses, and both concluded that the O/H ratio is less than 1/100 of Solar (the main reason for their similar result for the oxygen abundance was the similar \\Nev/\\Neiii\\ ratio used by both studies). \\cite{2005A&A...430..187P} merged and discussed the two observational data sets and conducted their own photoionization analysis. They concluded that the O/H ratio of \\TS\\ lies between 1/30 - 1/15 of Solar (still holding the record for the most oxygen poor planetary nebula but much higher than previously published). However, \\cite{2005A&A...430..187P} neglected to consider observations of \\TS\\ made with the \\textit{Hubble Space Telescope} (HST) and the \\textit{Far Ultraviolet Spectroscopic Explorer} (FUSE). As a result, some of their ``predicted'' line intensities are in conflict with what is actually observed in the UV. HST observations were obtained in 2003, and presented in a short, preliminary version by {\\cite{2006IAUS..234..431J}. Those authors quoted an oxygen abundance of 1/30 - 1/40 of Solar, and carbon and nitrogen abundances roughly 1/10 of Solar. Before embarking on a new determination of abundances, we have chosen to gather the best possible observations at all wavelength ranges. These data provide many more constraints than were available in any previous study. In order to make the best use of the large amount of data obtained with different telescopes, we use a pseudo-3D photoionization code, Cloudy\\_3D, which is able to account for the nebular geometry as we see it now, and with which we can properly take into account the aperture effects. This code is based on CLOUDY \\citep{1998PASP..110..761F} and was written by \\cite{2006IAUS..234..467M}. The paper is organized as follows. Section 2 presents the new observational material: several optical spectra, HST imaging and spectroscopy, infrared spectroscopy with the \\emph{Spitzer} Telescope, and mentions our X-ray observations with XMM. Section 3 summarizes other data that we used as constraints for the photoionization modelling. Section 4 describes our modelling strategy, and presents our ``reference model''. Section 5 evaluates the error bars on the derived elemental abundances, taking into account observational uncertainties in emission-line fluxes, uncertainties in model input parameters as well as uncertainties arising from an imperfect description of the physical processes included in the models. In Section 6, we compare the chemical composition of \\TS\\ with that of other PNe in the Galactic halo and discuss it in terms of stellar nucleosynthesis in the Asymptotic Giant Branch (AGB) phase. Finally, Section 7 summarizes our main findings. %__________________________________________________________________ ", "conclusions": "" }, "0911/0911.3689_arXiv.txt": { "abstract": " ", "introduction": "Much progress has been achieved in modeling the central stars of planetary nebulae, seen through the progressive refinement by many authors (eg. Harm and Schwartzschild 1975; Sch\\\"{o}nberner 1979, 1981, 1983; Iben 1984; Wood and Faulkner 1986). Together with dynamical models for the evolution of the ejected nebula (Mathews 1966; Sofia and Hunter 1968; Kwok, Purton and Fitzgerald 1978; Kwok 1982), most authors are now modeling PN evolution with a more holistic approach (Okorokv et al. 1985; Schmidt-Voigt and K\\\"{o}ppen 1987a; Stasinska 1989; 2008, Dopita and Meatheringham 1990, 1992; Guenther et al. 2003; Schwarz \\& Monteiro 2006; Holovatyy et al. 2008). This increasingly places more dependence on our understanding of the photoionisation of the nebula and may affect how we interpret diagnostic tools such as excitation class which has been established to give some indication of the central star's effective temperature. Most meaningful physical parameters that can be derived from a PN's observed emission lines, including ionized and total nebular mass and the brightness and evolutionary state of the central star, depend on accurate distances (Ciardullo et al. 1999). This is difficult in our own Galaxy due to inherent problems with variable extinction and lack of central star homoegeneity (Terzian, 1997). The well determined 50.6~Kpc LMC distance (e.g. Keller \\& Wood 2006), modest, 35~degree inclination angle and disk thickness (only $\\sim500$~pc, van der Marel \\& Cioni 2001), mean that LMC PNe are effectively co-located. Since dimming of their light by intervening gas and dust is low and uniform (e.g. Kaler \\& Jacoby, 1990), we can better estimate absolute nebula luminosity and size. This allows us to test the various methods for deriving excitation class and place this diagnostic on a better footing. We have constructed the most complete, least biased and homogeneous census of a PNe population ever compiled for a single Galaxy via discoveries from our deep AAO/UKST H$\\alpha$ multi-exposure stack of the LMC's central 25~deg$^{2}$ (Reid \\& Parker 2006a,b; Parker et al. 2005). PNe were confirmed by 2dF spectroscopy which gave 460 new PNe and independently recovered all 169 previously known PNe in the central area. % These data have led to significant advances in our understanding of the PN luminosity function (PNLF) (Reid \\& Parker, 2010) and physical parameters such as temperatures, densities, nebula masses and abundances (Reid 2007). Perfect for excitation studies, the LMC sample provides a large range of intrinsic flux intensities for individual emission lines (eg. 8.8E-12 to 3.7E-16 ergs cm$^{2}$ s$^{-1}$ for \\OIII\\,5007) and interesting differences in line ratios. Despite all the advantages of using PNe in the Magellanic Clouds for the modeling of dynamical evolution, without deep HST imaging the central stars cannot be directly observed. In order to understand the evolution of these PNe, properties of the central star such as mass and luminosity must be modeled using properties of the nebula. Fortunately, much work has been conducted in this area and it has been convincingly argued that the strength of emission lines in the nebula are uniquely determined by the properties of the central star (Dopita et al. 1987, 1988, 1990, 1992). Much of this central star modeling depends on some evaluation of the so-called excitation class parameter which is determined from ratios of specific emission lines, measured from the ionised nebula. In this paper, we use the large, new population of PNe in the LMC to compare the three main methods currently used for determining excitation class. Each of these different methods is somewhat grounded in the physics of low to high ionisation states but produce differing results. At last, however, with a large, more representative population of PNe in one system, covering a large evolutionary range, we can observe the different effects produced by these differing excitation schemes. Having plotted, assessed and compared the existing excitation schemes, each is found to be reflecting different aspects of the nebula-central star relationship, none of which is fully adequate. We suggest a new method for the determination of excitation class. This method provides a smooth transition between low and high excitation regimes and allows both the helium-to-hydrogen and the oxygen-to-hydrogen ratios to contribute a weighted proportion toward the excitation estimate. Each method is then assessed and compared by constructing diagrams, based on the standard H-R diagram. In this case, excitation class substitutes for central star temperature and the H$\\beta$ luminosity of the nebula represents a rough estimate of stellar luminosity. The diagram constructed using the most correct method should provide an evolutionary map, over which previously determined evolutionary tracks should show some degree of unity if excitation class has any true intrinsic value. ", "conclusions": "The utility of the commonly used excitation class estimator for PNe is investigated and appraised using a large, new sample of PNe in the LMC. The importance of determining the best form for deriving the excitation class is going to be an important part of our on-going study of LMC PNe. The large number of PNe now available over a wide luminosity and evolutionary range has permitted the excitation class parameter to be re-investigated and re-worked. The 3 main methods current in the literature for determining excitation class have been carefully plotted, compared and evaluated. We advise against using the Ex$_\\mathrm{neb}$ scheme in its current form due to the distortion it creates at the medium-to-high excitation region. The Ex$_{\\ast}$ scheme also appears to create some anomalies at the medium excitation levels where the input equation is re-worked. Using (\\OIII\\,$\\lambda4959+\\lambda5007)/\\mathrm{H\\beta}$ for all the PNe produces a smooth transition but suffers for not applying the \\HeII\\,$\\lambda$4686~ratio. The new Ex$_\\rho$ method offered in this paper appears to improve excitation and stellar temperature estimates by allowing both \\HeII\\,$\\lambda$4686 and \\OIII\\,$\\lambda$5007\\ to become constant variables against the H$\\beta$ intensity. It also minimises any sudden increase in PNe at the point where the \\HeII\\ line is introduced into the derived equation for high-excitation PNe. The method also correlates well with the best published Zanstra temperatures, placing this new excitation class parameter on a far more solid footing. This presentation has been prepared in order to lay the groundwork for further research using central star temperatures, electron densities, nebula mass and expansion velocities and new photoionisation codes. The disparity between the current methods for assigning excitation classes has been highlighted in this paper. Our suggested Ex$_{\\rho}$ method appears to produce encouraging results as it allows excitation to increase using both the \\OIII/H$\\beta$ and \\HeII\\,$\\lambda$4686/H$\\beta$ CS temperature indicators. It is clear that excitation is a quality that remains key to the ionised nebulae well into the ageing process and must be correctly modeled. Guided by these results and comparisons, we will continue to model the excitation class based on the ratios of the same ionic species, for example the ratios of O$^{+}$ to O$^{++}$ and He$^{+}$ to He$^{++}$. This method, used already by some authors (e.g. Martin \\& Viegas 2002) is expected to be a good CS temperature indicator. Our preliminary results using the LMC PNe are encouraging. With new photoionisation models, we intend refining the input parameters and anticipate that new evolutionary tracks and diagrams describing the AGB to WD stages will be produced." }, "0911/0911.1339_arXiv.txt": { "abstract": "We report the detection of a 115 day periodicity in {\\it SWIFT}/XRT monitoring data from the ultraluminous X-ray source (ULX) NGC 5408 X-1. Our ongoing campaign samples its X-ray flux approximately twice weekly and has now achieved a temporal baseline of $\\approx 485$ days. Periodogram analysis reveals a significant periodicity with a period of $115.5 \\pm 4$ days. The modulation is detected with a significance of $3.2 \\times 10^{-4}$. The fractional modulation amplitude decreases with increasing energy, ranging from $0.13 \\pm 0.02$ above 1 keV to $0.24 \\pm 0.02$ below 1 keV. The shape of the profile evolves as well, becoming less sharply peaked at higher energies. The periodogram analysis is consistent with a periodic process, however, continued monitoring is required to confirm the coherent nature of the modulation. Spectral analysis indicates that NGC 5408 X-1 can reach 0.3 - 10 keV luminosities of $\\approx 2 \\times 10^{40}$ ergs s$^{-1}$. We suggest that, like the 62 day period of the ULX in M82 (X41.4+60), the periodicity detected in NGC 5408 X-1 represents the orbital period of the black hole binary containing the ULX. If this is true then the secondary can only be a giant or supergiant star. ", "introduction": "The existence of black holes in the mass range from $10^{2} - 10^{4}$ $M_{\\odot}$--intermediate mass black holes (IMBH)--is still not widely accepted. It has been argued based on their extreme luminosities that some of the bright X-ray sources found in nearby galaxies, the ultraluminous X-ray sources (ULXs), may be IMBHs (Colbert \\& Mushotzky 1999), \u0153but it has also been suggested that these objects may appear luminous due to beaming of their X-ray radiation (King et al. 2001). As yet there has been no direct measurement of the mass of a ULX. At present the best IMBH candidates include X41.4+60, the brightest ULX in the starburst galaxy M82 (also referred to as M82 X-1, Kaaret, Feng \\& Gorski 2009; Strohmayer \\& Mushotzky 2003), and the ULX NGC 5408 X-1 (Strohmayer et al. 2007; Kaaret \\& Corbel 2009). Very recently, Farrell et al. (2009) have reported the identification of an X-ray source (2XMM J011028.1-460421) very near the absorption line galaxay ESO 243-49. They argue for a physical association with this galaxy which at a redshift of $z = 0.0224$ implies a luminosity in excess of $10^{42}$ erg s$^{-1}$, and suggest it is an IMBH in excess of 500 $M_{\\odot}$. If the association with ESO 243-49 is correct then this object is the most luminous ULX currently known. X41.4+60 is also amongst the most luminous ULXs, on occasion having a luminosity upwards of $10^{41}$ ergs s$^{-1}$ (Kaaret et al. 2009). This object also shows a 62 day periodic modulation in its X-ray flux, which has been proposed to be the orbital period of the binary system containing the black hole (Kaaret, Simet \\& Lang 2006; Kaaret \\& Feng 2007). Both NGC 5408 X-1 and X41.4+60 show quasiperiodic oscillations (QPOs) in their X-ray fluxes, and these ULX QPOs appear at systematically lower frequencies than the analogous QPOs observed in Galactic black hole binary systems (Strohmayer \\& Mushotzky 2003; Strohmayer et al. 2007). Indeed, Strohmayer \\& Mushotzky (2009) have recently shown that the QPO properties in NGC 5408 X-1 correlate with the energy spectrum and X-ray flux in a manner fully consistent with the behavior observed for so-called Type C QPOs in Galactic systems. The Type C QPOs in stellar mass black holes are strong (fractional rms amplitude $\\sim 15\\%$), relatively coherent ($\\nu_{qpo} / \\Delta\\nu > 10$) oscillations with characteristic frequencies of $\\sim 1 - 10$ Hz that vary in frequency in correlation with source flux and spectral index (see Sobczak et al 2000; Vignarca et al. 2003; Casella, Belloni \\& Stella 2005). They are associated with an energy spectral state in which approximately half or more of the flux is carried by a power-law component with slope $\\Gamma \\sim 2 - 2.5$, commonly referred to as the Steep Power Law state (SPL, McClintock \\& Remillard 2006), or the Hard Intermediate State (HIMS, Belloni 2006). The precise origin of these QPOs is still uncertain, but their phenomenology has been used to estimate the masses of black holes (see Shaposhnikov \\& Titarchuk 2009). The Type C QPOs in NGC 5408 X-1 identified by Strohmayer \\& Mushotzky (2009) have frequencies of a few tens of mHz, a factor of 100 lower than the Galactic black holes. Scaling the observed QPO frequencies in NGC 5408 X-1 to those observed in Galactic systems of known mass suggests that NGC 5408 X-1 is an IMBH with a mass greater than 1000 $M_{\\odot}$ (Strohmayer \\& Mushotzky 2009). Moreover, recent optical spectroscopy reported by Kaaret \\& Corbel (2009) indicates that the optical counterpart to NGC 5408 X-1 is associated with reprocessed emission from an accretion disk as well as a strongly photoionized optical nebula. This, combined with the presence of a powerful radio nebula also appears consistent with the presence of an IMBH (Lang et al. 2007; Soria et al. 2006; Kaaret et al. 2003). While it is likely that many of the ULXs are accreting black hole binary systems, very little is known about the nature of these putative binaries. If some of the brighter systems are indeed IMBHs accreting via Roche lobe overflow, then their orbital periods could be as long as a few hundred days (Portegies Zwart et al. 2004). With the exception of X41.4+60 noted above, very few objects have been monitored with a cadence and duration that would enable sensitive searches for X-ray periods in the range of 10s to 100s of days. The {\\it Swift} observatory uniquely provides both the X-ray imaging sensitivity and scheduling flexibility to enable such searches. Starting with Observing Cycle 4 {\\it Swift} has been monitoring NGC 5408 X-1 several times per week as part of an approved program. These observations have continued into Cycle 5 and are presently ongoing. Here we present results from this campaign which provide evidence for a 115 day periodicity in NGC 5408 X-1 which could very likely be the orbital period of a Roche lobe overflow binary containing an IMBH. ", "conclusions": "A number of accreting binaries show X-ray modulations at their orbital periods. Among black hole binaries Cyg X-1 and Cyg X-3 are well known examples (Wen et al. 1999; Elsner 1980), and more recent detections include LMC X-3 (Boyd et al. 2001), 1E 1740.7-2942 and GRS 1758-258 (Smith et al. 2002). Many neutron star binaries also show X-ray modulations at the binary orbital period. The orbital modulations in accreting binaries have typically been attributed to periodic obscuration produced by a vertically and azimuthally structured accretion disk or scattering in an extending accretion disk corona or stellar wind from the donor (Parmar \\& White 1988; Wen et al. 1999). A recent example of an X-ray modulation at the putative orbital period in a neutron star system is that of GX 13+1 (Corbet 2003). Thus, if ULXs are indeed accreting black hole binaries, then it is not unexpected to detect X-ray modulations at their orbital periods. While their have been several claims of detection of orbital modulations in ULXs (see, Kaaret, Simet \\& Lang 2006 for a brief summary), the most compelling detection to date is that of the 62 day modulation from the M82 ULX X41.4+60 (Kaaret, Simet \\& Lang 2006; Kaaret \\& Feng 2007). The average modulation in X41.4+60 as observed with the RXTE/PCA is roughly sinusoidal with an amplitude of $\\approx 20\\%$ (Kaaret \\& Feng 2007). This is qualitatively consistent with the $0.2 < E < 8$ keV modulation amplitude and profile that we report here for NGC 5408 X-1, and suggests that similar physical processes may produce the observed modulation in both systems. Superorbital periods are known in both neutron star and black hole binaries. Well known examples include the 35 day modulation in the neutron star system Her X-1, and the 164 day period in the putative black hole binary SS 433 (Wijers \\& Pringle 1999). These variations have generally been ascribed to accretion disk precession (Ogilvie \\& Dubus 2001). Kaaret \\& Feng (2007) argued that the 62 day period in X41.4+60 was unlikely to be a superorbital modulation because the observed period is only consistent with the observed superorbital periods of neutron star systems, but that the extreme luminosity of X41.4+60 comfortably rules out an accreting neutron star (see their Figure 4 for a distribution of observed superorbital periods). Similar arguments would appear valid for the 115 day modulation in NGC 5408 X-1. The 115 day period is shorter than all known superorbital periods for black hole candidate binaries, and its peak luminosity is well in excess of the Eddington limit for a neutron star ($\\approx 10^{38}$ ergs s$^{-1}$). Using X-ray timing measurements with {\\it XMM-Newton}, Strohmayer et al. (2009) have recently shown that the strong QPO observed in NGC 5408 X-1 varies in frequency and amplitude with changes in the X-ray flux and energy spectrum in a manner that closely mimics the correlated temporal and spectral variations observed in stellar mass black holes in the so-called Intermediate State (also known as the Steep Power-law State). Moreover, recent optical spectroscopy reported by Kaaret \\& Corbel (2009) indicates that its optical counterpart has a significant contribution from reprocessing of the X-ray luminosity in the outer parts of the disk. The so-called ``slim disks'' that may exist at super-Eddington accretion rates preferentially radiate more flux out along the disk axis, and therefore have relatively less available for reprocessing (Abramowicz et al 1988). These models thus have difficulty accounting for the observed optical to X-ray flux ratio in NGC 5408 X-1. Further, modeling of the observed nebular optical emission lines supports the conclusion that NGC 5408 X-1 radiates $\\sim 10^{40}$ ergs s$^{-1}$ in an approximately isotropic manner (Kaaret \\& Corbel 2009). These observations support the notion that NGC 5408 X-1 harbors a ``standard'' thin, optically thick accretion disk similar to those inferred to exist in Galactic black hole binaries. The 115 day modulation we have found in NGC 5408 X-1 shows interesting energy dependent effects that would appear consistent with orbital modulation. A system which shows qualitatively similar effects to those seen in NGC 5408 X-1 is Cyg X-1. Specifically, the orbital modulation in Cyg X-1 has been modeled in the context of absorption and scattering of X-rays in the partially ionized wind from the companion star (Wen et al. 1999). This model predicts a decrease in modulation amplitude with increasing energy, and a minimum in the hardness ratio at the maximum of the modulation profile (see Wen et al. 1999, Figure 3), both of which are evident in the NGC 5408 X-1 data. While the shapes of the modulation profiles in Cyg X-1 do not exactly match the results we find for NGC 5408 X-1, the behavior of the $E > 1$ profile (see Figure 3) appears qualitatively similar. An important distinction would seem to be that in the case of NGC 5408 X-1 we can directly observe the thermal disk flux whereas the orbital modulation measurements for Cyg X-1 concern the low-hard state, when presumably its thermal flux is weak or absent. Moreover, any disk flux appears largely shortward of the RXTE/PCA energy band. Cyg X-1 is also a wind accreting system, whereas wind accretion would likely be unable to account for the high luminosity of NGC 5408 X-1. If the 115 day period is indeed the orbital period of a Roche lobe filling binary, then the mean density of the secondary is constrained to be $\\rho_{mean} \\approx 0.2 (P_{\\rm days})^{-2}$ g cm$^{-3}$ (Frank et al. 2002). For a period of 115.5 days this gives $\\rho_{mean} = 1.50 \\times 10^{-5}$ g cm$^{-3}$, and the companion would have to be a giant or supergiant star. Recent observations suggest that a substantial fraction of the optical counterpart is due to reprocessing of the X-ray flux in an accretion disk, and while the companion star still has not been observed directly, the recent optical measurements suggest that it is probably a giant in the 3 - 5 $M_{\\odot}$ range, and with a spectral type of B or later (Kaaret \\& Corbel 2009). Binary evolution calculations for IMBHs and massive companions appear consistent with the notion that NGC 5408 X-1 contains a $\\sim 1000 M_{\\odot}$ black hole with a 3 - 5 $M_{\\odot}$ companion (Li 2004). Continued monitoring of NGC 5408 X-1 with {\\it Swift} is important in order to confirm the orbital nature of the modulation." }, "0911/0911.4036_arXiv.txt": { "abstract": "{} {Our goal is to understand the nature of blazars and the mechanisms for the generation of high-energy $\\gamma$-rays, through the investigation of the blazar 3C 66A.} {We model the high energy spectrum of 3C 66A, which has been observed recently with the Fermi-LAT and VERITAS telescope. The spectrum has a hard change from the energy range of 0.2-100 GeV to 200-500 GeV in recent almost contemporaneous observations of two telescopes. } { The de-absorbed VERITAS spectrum greatly depends on the redshift, which is highly uncertain. If z=0.444 is adopted, we are able to use the SSC model to produce the Fermi-LAT component and the EC model to the VERITAS component. However, if z=0.1, the intrinsic VERITAS spectrum will be softer, there will be a smooth link between the Fermi-LAT and VERITAS spectra which can be explained using a SSC model. } {} ", "introduction": "Blazars are a peculiar class of active galactic nuclei (AGN) and their jets point at small angles with respect to our line of sight. Many of them have been observed at all wavelengths, from radio to very high energy (VHE) $\\gamma$-rays. Their spectral energy distribution (SED) consists of two bumps which are attributed to the synchrotron and the inverse Compton (IC) emission of ultrarelativistic particles. The different soft photon sources deduce synchrotron self-Compton (SSC) and external Compton (EC) models to produce high energy emission. In the SSC model, the soft photons are provided by the synchrotron emission of the same electrons (\\cite{Marscher80}; \\cite{Ghisellini89}; \\cite{Marscher96}). However in the EC model, the soft photons mostly come from the outside of the jet, such as outer disk, broad-line region (BLR) clouds (etc. \\cite{Dermer93}; \\cite{Sikora94}). In 3C 66A, \\cite{Miller78} have given the redshift z=0.444 by a weak Mg II emission line detection, but it is very uncertain (\\cite{Bramel05}). When 3C 66A locates at z=0.444, its TeV photons will suffer the strong pair production absorption of the Extragalactic Background light (EBL). After corrected by the EBL absorption, TeV emission presents an inverted intrinsic spectrum (see the Acciari et al. (2009) Fig. 2, the de-absorbed photon spectral index is calculated to be $1.1 \\pm 0.4$). In this paper, we take z = 0.444 to analyze TeV emissive mechanism, and will discuss the behavior of the de-absorbed VERITAS spectrum in different redshifts. Generally, 3C 66A is classified as a low-frequency peaked BL Lac object (LBL). The peak of the low-frequency component of LBLs usually lies in the IR or optical regime, whereas the peak of high-energy component locates at several GeV. The luminosity of $\\gamma$-ray is typically comparable to or slightly higher than the luminosity of the synchrotron component. Such as, \\cite{Joshi07} have argued that for 3C 66A the peak of low-frequency component locates at the optical regime, the peak of high-frequency component reaches multi MeV-GeV range. However, \\cite{Perri03} have revealed that the synchrotron peak locates in between $10^{15}$ and $10^{16}$ Hz, then 3C 66A is classified as an intermediate-frequency peaked BL Lac (IBL). From the X-ray spectrum with the photon spectral index $\\Gamma \\sim 2.5$ (Bottcher et al. 2005; Donato et al. 2005; Foschini et al. 2006), which might be the tail of the synchrotron emission, 3C 66A is considered as an IBL in this paper. 3C 66A is observed in radio, IR, optical, X-rays and $\\gamma$-rays, and shows large luminosity variations. As described in \\cite{Bottcher05}, the object exhibits several outbursts in the optical band and the variations of $\\Delta$m $\\sim$ 0.3-0.5 over a timescale of several days. Until now, the majority of BL Lacs detected at VHE (very high energy: E $>$ 100 GeV) are HBLs (high-frequency peaked BL Lacs). Only IBL W Comae (\\cite{Acciari08}), LBL BL Lacertae (\\cite{Albert07}) and 3C 279 (\\cite{Albert08}) display the potential to enlarge the extragalactic TeV source. For 3C 66A, the Crimean Astrophysical Observatory have reported a 5.1$\\sigma$ detection above 900 GeV(Stepanyan et al. 2002). Recently, VERITAS have carried out 14 hours' observations for 3C 66A from September 2007 through January 2008 (hereafter, the 2007-2008 season), and from September through November 2008 (hereafter, the 2008-2009 season) a further 46 hours' data have been taken (\\cite{Acciari09}). Due to the limited spatial resolution of Cherenkov telescopes it is difficult to accurately identify the emission region. The radio galaxy 3C 66B lies in the same view field of 3C 66A at a separation of 0.12$^{\\circ}$ and is also a plausible source of VHE radiation (\\cite{Tavecchio08}). The recent detection by MAGIC favored 3C 66B as VHE source and excluded 3C 66A at an 85\\% confidence level (\\cite{Aliu09}). However , VERITAS have found that 3C 66A lies 0.01$^{\\circ}$ from the fit position while 3C 66B lies 0.13$^{\\circ}$ away, and 3C 66A is VHE source. If 3C 66A has a redshift of z = 0.444, its de-absorbed spectral index is 1.1$\\pm $0.4 showing very hard intrinsic spectrum (\\cite{Acciari09}). In first three months, the Fermi-LAT Gamma-ray Space Telescope have observed 3C 66A (\\cite{Abdo09}), almost at contemporaneous observation with VERITAS in the 2008-2009 season. However, very soft spectrum with the spectral index of 1.97$\\pm $0.04 appears in the Fermi-LAT observing energy range. The $\\gamma$-ray spectrum suddenly hardens from 0.2-100 GeV to 200-500 GeV and challenges the one-zone homogeneous SSC model. In Section 2 we present the jet models for application to 3C 66A. We use the observed data to constrain the model parameters in Section 3. We finish with discussions and conclusions in Section 4. Throughout this paper, we use a soft cosmology with a deceleration factor $q_0 = 0.5$ and a Hubble constant $H_0 = 75 km s^{-1} Mpc^{-1}$. ", "conclusions": "The redshift of 3C 66A has an uncertain value (\\cite{Bramel05}), and is usually adopted as z = 0.444. However the redshift is crucial in constructing intrinsic high energy spectrum because of the EBL absorption (\\cite{Hauser01}). This absorption decreases the observed flux and softens the observed spectrum. If the redshift is less than 0.444, such as just $\\geq$ 0.096 suggested by \\cite{Finke08}, the intrinsic spectrum in the VERITAS energy range will be softer. There will be a smooth link spectrum between the Fermi-LAT and the VERITAS energy ranges. In the Fig.~\\ref{fig.2}, we generate the intrinsic spectra of VERITAS observations corrected by EBL absorption according to \\cite{Franceschini08} model, assuming the source to be at the different redshifts z=0.03, 0.1, 0.3, and 0.5. It is shown that the de-absorbed spectrum strongly depends on the redshift. When z=0.3, the de-absorbed spectrum has a little inverted, but it becomes an inverted spectrum in z=0.5. If the redshift is less than 0.1, the de-absorbed spectrum will present the usual SED of an LBL such as W Comae (\\cite{Acciari08}). In the Fig.~\\ref{fig.3}, we show the de-absorbed SEDs under z=0.1 and the modeling. A smooth spectrum can link the Fermi-LAT and VERITAS data and be explained with a SSC model, in which $U_B=9.95 \\times 10^{-5} erg \\cdot cm^{-3}$ and $U_e=2.31 \\times 10^{-3} erg \\cdot cm^{-3}$. In fact the bulk motion of the emitting blob affects the observed SED (e.g., \\cite{Dermer95}; Georganopoulos et al. (2001)). The peak frequencies are given by $\\nu_s\\approx 3.7 \\times 10^6 \\gamma _{peak}^2 \\delta B /(1+z)$ , $\\nu_{SSC}\\approx \\frac{4}{3}\\gamma _{peak}^2 \\nu_{s}$, and $\\nu_{EC}\\approx \\frac{4}{3} \\gamma _{peak}^2 \\nu_{ext} \\Gamma \\delta /(1+z)\\label{nu_ec}$. We can see that $\\nu_{EC}$ would be larger than $\\nu_{SSC}$ provided the blob has larger $\\delta$ or $\\Gamma$ (Usually the viewing angle $\\theta\\sim 1/\\Gamma$ is assumed, $\\delta\\sim \\Gamma$.) From the ratio of peak luminosity, $\\frac{L_{EC}}{L_{SSC}} \\approx \\frac{L_{EC}^{'}}{L_{SSC}^{'}} = \\frac{\\dot\\gamma_{peak,EC}}{\\dot\\gamma_{peak,SSC}} \\approx \\frac{U^{'}_{ext}}{U^{'}_{syn}} \\approx \\delta^4 \\Gamma^2 \\frac{\\xi L_{ext}}{L_s} \\frac{R^2}{R^2_{ext}}$ ($U_{ext}$ is amplified by a factor of $\\Gamma^2$, see the Sikora et al. (1994) and Dermer (1995)), where $\\xi$ is the reproduce fraction of the $L_{ext}$, we show that $\\frac{L_{EC}}{L_{SSC}}$ is strongly affected by the bulk motion of the blob. % In the Fig. 1, using $U_B^{'}=4.6 \\times 10^{-5} erg \\cdot cm^{-3}$, $U_e^{'}=1.67 \\times 10^{-3} erg \\cdot cm^{-3}$, and $U_{ext}^{'} =3.5 \\times 10^{-6} erg \\cdot cm^{-3}$ we can model the SED. $U_{ext}^{'}$ is obviously lower than $U_B^{'}$, however, the EC luminosity is comparable with the synchrotron ones (see the Fig. 1). In the observer frame, the beaming factor is different for EC ($\\delta ^{4+2\\alpha}$ (\\cite{Dermer95}; Georganopoulos et al. 2001)), synchrotron and SSC emission ($\\delta ^{3+ \\alpha}$). The difference of EC and synchrotron luminosity is reasonable by considering their beaming factor. The EC emission is less clear for the BL Lac objects. The lack of strong emission lines and UV excesses suggests that the external photon density $\\xi L_{ext}$ is very low, while the Lorentz factor of BL Lac objects is typically smaller than that of quasars (\\cite{piner08}). Their high energy emission strongly favors the SSC mechanism over the EC mechanism. But, 3C 66A might be an exception and reveal an existence of larger bulk velocity in the high energy emissive region. Therefore, the high energy emission caused by EC mechanism is likely observed in the IBL. This possibility needs future Fermi-LAT and VERTAS observations for 3C 66A and a precise redshift determination." }, "0911/0911.3774.txt": { "abstract": "The velocity of the inner ejecta of stripped-envelope core-collapse supernovae (CC-SNe) is studied by means of an analysis of their nebular spectra. Stripped-envelope CC-SNe are the result of the explosion of bare cores of massive stars ($\\geq 8$ M$_{\\odot}$), and their late-time spectra are typically dominated by a strong [O {\\sc i}] $\\lambda\\lambda$6300, 6363 emission line produced by the innermost, slow-moving ejecta which are not visible at earlier times as they are located below the photosphere. A characteristic velocity of the inner ejecta is obtained for a sample of 56 stripped-envelope CC-SNe of different spectral types (IIb, Ib, Ic) using direct measurements of the line width as well as spectral fitting. For most SNe, this value shows a small scatter around 4500 km s$^{-1}$. Observations ($< 100$ days) of stripped-envelope CC-SNe have revealed a subclass of very energetic SNe, termed broad-lined SNe (BL-SNe) or hypernovae, which are characterised by broad absorption lines in the early-time spectra, indicative of outer ejecta moving at very high velocity ($v \\geq 0.1 c$). SNe identified as BL in the early phase show large variations of core velocities at late phases, with some having much higher and some having similar velocities with respect to regular CC-SNe. This might indicate asphericity of the inner ejecta of BL-SNe, a possibility we investigate using synthetic three-dimensional nebular spectra. ", "introduction": "\\label{intro} Massive stars ($>8$ M$_\\odot$) collapse when the nuclear fuel in their central regions is consumed, producing a core-collapse supernova (CC-SN) and forming a black hole or a neutron star. CC-SNe with a H-rich spectrum are classified as Type II. If the envelope was stripped to some degree prior to the explosion, the SNe are classified as Type IIb (strong He lines, and weak but clear H), Type Ib (strong He lines but no H), and Type Ic (no He or H lines) \\citep{1997ARA&A..35..309F}. Some CC-SNe, called broad-lined SNe (BL-SNe), exhibit very broad absorption lines in their early phase, resulting from the presence of sufficiently massive ejecta expanding at high velocities. BL-SNe seem to be preferentially of Type Ic [two exceptions are the Type IIb SN 2003bg \\citep{2009arXiv0908.1783H} and the Type Ib SN 2008D \\citep{2008Sci...321.1185M,2009ApJ...702..226M}]; see also SN 1987K \\citep{1988AJ.....96.1941F}, which is mentioned by \\citep{2009arXiv0908.1783H}. Some BL-SNe can reach kinetic energies of $\\ge$ 10$^{52}$ ergs. They are sometimes called hypernovae, and can be associated with long-duration gamma-ray bursts (GRBs) \\citep[see][and references therein]{2006ARA&A..44..507W}. However not all BL-SNe are associated with GRBs. An important question in the context of CC-SNe is how the gravitational energy is converted into outward motion of the ejecta during the collapse; see \\citep{2007PhR...442...38J} for a recent review. In GRB scenarios a relativistic outflow is launched by the central engine and deposits some fraction of its energy into the SN ejecta, probably preferentially along the polar axis, which might cause strong asymmetries \\citep[e.g.,][]{2002ApJ...565..405M}. The nearest, best-studied GRB-SNe are SN 1998bw / GRB 980425 \\citep{1999A&AS..138..465G}, SN 2003dh / GRB 030329 \\citep{2004cetd.conf..351M}, SN 2003lw / GRB 031203 \\citep{2004ApJ...609L...5M}, and SN 2006aj / GRB/XRF 060218 \\citep{2006Natur.442.1011P}, although it is not fully established that the GRBs (or X-ray flashes) accompanying nearby CC-SNe can be compared directly to high-redshift GRBs. CC-SNe may be characterised by asphericities although a jet does not necessarily form \\citep{2003ApJ...584..971B,2007ApJ...655..416B,2004ApJ...608..391K,2006MNRAS.370..501M,2009ApJ...691.1360T}. Extremely massive stars ($> 100 $ M$_\\odot$) are thought to end their lives as pair-instability SNe (PI SNe). A star with enough mass to form a He core with more than 40 M$_\\odot$ will suffer electron-positron pair instability, leading to rapid collapse. This triggers explosive oxygen burning, leading to the complete disruption of the star \\citep{1967PhRvL..18..379B,2005IAUS..228..297H}. This process can produce large amounts of $^{56}$Ni. The ejecta mass and the explosion kinetic energy are high, but the ejecta velocities are moderate \\citep{2005ApJ...633.1031S}. Stripped CC-SNe, which lost part of their envelope before collapse, offer a clearer view of their inner ejecta than SNe which have retained it. Thus, in this paper we exclusively address stripped CC-SNe (Types IIb, Ib, Ic); we do not include Type II SNe, despite using the term ``CC-SN.'' Asphericities in their inner and outer ejecta \\citep[e.g.,][]{2001ApJ...559.1047M} are evident in at least some CC-SNe. Two main indicators are velocity differences of Fe and lighter-element lines, and polarisation measurements \\citep[e.g.,][]{1991A&A...246..481H}. Independent of their type, with time SNe become increasingly transparent to optical light, as the ejecta thin out. At late times ($> 200$ days after the explosion), the innermost layers of the SN can be observed. This epoch is called the nebular phase, because the spectrum turns from being absorption dominated to emission, mostly in forbidden lines. In this phase the radiated energy of a SN is provided by the decay of radioactive $^{56}$Co (which is produced by the earlier decay of $^{56}$Ni). Decaying $^{56}$Co emits $\\gamma$-rays and positrons which are absorbed by the SN ejecta. As the deposition rate of $\\gamma$-rays and positrons depends on the density and $^{56}$Ni distribution, the inner parts of the SN dominate the nebular spectra. Recently, several authors \\citep{2008Sci...319.1220M,2008ApJ...687L...9M,2009arXiv0904.4632T} have studied nebular spectra of CC-SNe. They concentrated on the shape of the [O {\\sc i}] $\\lambda\\lambda$6300, 6364 doublet (which is produced by much of the mass) and concluded that torus-shaped oxygen distributions might cause the double-peaked [O {\\sc i}] profile observed in many CC-SNe nebular spectra. However, there is ongoing discussion regarding whether geometry is the dominant reason for this type of line profile \\citep[][also see Appendix A]{2009arXiv0904.4256M}. Several authors have modelled nebular-phase spectra of SNe to derive quantities such as the $^{56}$Ni mass \\citep[e.g.,][]{2004ApJ...614..858M,2006A&A...460..793S,2006MNRAS.369.1939S,2007ApJ...658L...5M}, ejecta velocities \\citep[e.g.,][]{2007Sci...315..825M}, asphericities \\citep[e.g.,][]{2005Sci...308.1284M,2007ApJ...670..592M,2008Sci...319.1220M}, and elemental abundances \\citep[e.g.,][]{2007ApJ...658L...5M, 2007ApJ...666.1069M}. The nebular phase is especially suitable for studying the core of SNe. If different explosion scenarios are involved for different types of CC-SNe, one might expect the largest, most revealing differences to be in the central region of the explosion. Therefore, in contrast to the standard classification of BL-SNe, which is based primarily on early-time spectroscopy and describes the velocity field of the outer SN layers, here we focus on the centre of the explosion. We have modelled the nebular spectra of over 50 SNe, the largest sample of CC-SNe so far, and obtained a statistically significant representation of their core velocities. We describe our data set in Section 2 and the modelling procedure in Section 3. In Section 4 we test the reliability of the modelling approach. Results are discussed in Section 5. ", "conclusions": "\\label{conc} We estimated $^{56}$Ni masses for 29 SNe for which $^{56}$Ni masses were not previously known. Their average agrees with the average of 18 CC-SN $^{56}$Ni masses estimated before rather well, however there might be a small systematic over-estimate of the $^{56}$Ni mass in this work. Individual estimates might be quite erroneous given the poor quality of the data set. We then measured characteristic velocities for 56 CC-SN cores, which range from 3000 km s$^{-1}$ to 7000 km s$^{-1}$ ($v_{\\alpha}$). Several BL-SNe with high-velocity outer ejecta have high-velocity cores as well, but some BL-SNe do not. We have shown that this might be due to ejecta asphericities, which are expected theoretically and might have been detected by different methods before. We found that the average core velocities of SNe~IIb are slightly lower than core velocities of SNe~Ib and SNe~Ic. SNe~Ib and Ic have very similar average core velocities. SNe~IIb show much less variance ($\\sigma_v \\approx 400$ km s$^{-1}$) of their core velocities than SNe~Ib and Ic ($\\sigma_v \\approx 850$ km s$^{-1}$). There seems to be no strong dependence of core velocity on $^{56}$Ni mass. We also checked for correlations between total ejecta mass or kinetic energy and the core velocity and found none. There is only a weak correlation between the ratio of total kinetic energy to total mass and the core velocity (BL-SNe have the highest core velocities on average, but the scatter is so large that there is almost no predictive power). Therefore, although the total mass of a SN might be estimated (if the spectrum is flux calibrated), it is not possible to estimate the SN total kinetic energy from a nebular spectrum with good accuracy, since the core velocity seems to correlate with the outer ejecta velocities only weakly. We will further study the relation between outer and inner ejecta velocities in future work. The uncertainties in our estimates are rather large. They are caused by the general properties of our method (background subtraction, reddening, time evolution of line width, uncertainty on epoch, $^{56}$Ni distribution, and $^{56}$Ni mass) and are difficult to improve upon when only limited high-quality data are available that cover different CC-SNe at several different epochs. Therefore more and better data is needed." }, "0911/0911.4493_arXiv.txt": { "abstract": "We present recent results of 3D magnetohydrodynamic simulations of neutron stars with small misalignment angles, as regards the features in lightcurves produced by regular movements of the hot spots during accretion onto the star. In particular, we show that the variation of position of the hot spot created by the infalling matter, as observed in 3D simulations, can produce high frequency Quasi Periodic Oscillations with frequencies associated with the inner zone of the disk. Previously reported simulations showed that the usual assumption of a fixed hot spot near the polar region is valid only for misalignment angles $\\mis$ relatively large. Otherwise, two phenomena challenge the assumption: one is the presence of Rayleigh-Taylor instabilities at the disk-magnetospheric boundary, which produce tongues of accreting matter that can reach the star almost anywhere between the equator and the polar region; the other one is the motion of the hot spot around the magnetic pole during stable accretion. In this paper we start by showing that both phenomena are capable of producing short-term oscillations in the lightcurves. We then use Monte Carlo techniques to produce model lightcurves based on the features of the movements observed, and we show that the main features of kHz QPOs can be reproduced. Finally, we show the behavior of the frequencies of the moving spots as the mass accretion rate changes, and propose a mechanism for the production of double QPO peaks. ", "introduction": "\\label{sec:intro} Quasi Periodic Oscillations (QPO) have been observed in many X-ray sources for the last 24 years \\citep{vanderKlis:1985p2318,Lamb:1985p2327}. They appear as broad peaks in the power spectra, with a behavior dependent on the source state, the mass accretion rate, and other physical properties of the sources. They are present in both neutron star (hereafter NS) and black hole (BH) sources, sometimes with peculiar similarities between the two classes of objects \\citep[see][for a review]{2006csxs.book...39V}. Of particular interest are the kHz QPOs: they were first discovered in the Low Mass X-ray Binary (LMXB) and Z source Sco-X1 \\citep{1996ApJ...469L...1V}, to be then observed in the spectra of nearly all Z and atoll sources, as well as several weak LMXBs \\citep[see for example][]{Strohmayer:1996p2317,Wijnands:1997}. They have frequencies between 300Hz and 1.2kHz \\citep[see][]{2006csxs.book...39V}. The peaks move with timescales of hours/days and their quality factor $Q=\\nu/\\Delta\\nu$ (ratio between the frequency and the width of the peak, a measure of coherence) is not constant, but can be related with the QPO frequency \\citep{Mendez:2001p4640, diSalvo:2003p5128,Barret:2006p7645} and with the mass accretion rate \\citep[see for example][]{Mendez:2006p5133}. The kHz QPOs usually appear as a pair of peaks (the {\\em twin peaks}), labeled ``upper'' and ``lower'' (and their frequencies, accordingly, $\\nu_u$ and $\\nu_l$). While on timescales of hours the two frequencies change, their difference $\\Delta\\nu=\\nu_u-\\nu_l$ tends to be almost costant. As with the discovery of burst oscillations \\citep{Strohmayer:1996p2317} and Accreting Millisecond X-ray Pulsars (hereafter AMXP) \\citep {Wijnands:1998p1343} the spin frequencies $\\spinf$ of more and more accreting neutron stars were measured, an interesting behavior was observed: $\\Delta\\nu$ was usually found to be close to either $\\spinf$ or $0.5\\spinf$. Recent works \\citep[see for example][]{Mendez:2007p5756} call this assumption into question. In particular, it can be shown that $\\Delta \\nu$ could in principle be independent from $\\spinf$. It has been shown that the kHz QPOs frequency is correlated, at least on a given system and for short timescales, with the X-ray count rate \\citep{1999ApJ...511L..49M} and thus also, probably, with the mass accretion rate. Many models have been developed to explain the kHz QPO phenomenon. They involve a number of different mechanisms, like for example relativistic precession frequencies \\citep{Stella:1998p4748}, relativistic resonances \\citep{Kluzniak:2001p4749}, beats between the keplerian frequency at a particular radius in the disk and the frequency of the star \\citep{Alpar:1985p2315,Lamb:1985p2327,Miller:1998p4107,Lamb:2004p6829}, oscillations in the comptonizing medium \\citep{Lee:1998p4811}. Recently, \\citet{Lovelace:2007p4213} and \\citet{Lovelace:2009p4323} found and analyzed a radially localized Rossby wave instability which exists in the inner region of a disk around a magnetized star where the angular rotation rate of the matter decreases approaching the star. This instability may explain the twin kilo-Hertz QPOs observed in some LMXBs. The importance for QPOs of the disk angular rotation rate having a maximum was discussed earlier by \\citet{Alpar:2005,Alpar:2008} Most of the models of kHz QPOs, based on the fact that QPOs also appear in black hole systems and there seems to be a link between $\\nu_l$ with frequencies observed in black hole systems \\citep{Psaltis:1999p4878}, try to explain the behavior of kHz QPOs in both the neutron star and the black hole cases, excluding the possibility of a surface emission. \\citet{2005AN....326..812G} claim that, according to spectral emission properties of QPOs, in Atoll and Z sources periodic and quasi-periodic emission is more likely to be produced on the surface of the star. During the past years MHD simulations have been amongst the most fruitful methods to investigate accreting magnetized stars \\citep[see for example][]{Goodson:1997p5065, Romanova:2002p4187, Romanova:2003, Romanova:2004p4095, Kulkarni:2005p4093, Bessolaz:2008p4909, Long:2008p4237, Romanova:2008}. 3D simulations have given insight into the way matter accretes, showing in more detail trajectories, shapes of magnetic field lines, shapes and movements of hot spots on the surface of the star \\citep[see for example][]{Romanova:2003,Romanova:2004p4095}. Simulations show that matter may accrete in several different regimes: \\begin{itemize} \\item the {\\em stable accretion} regime, when matter is stopped by the magnetic field and conveyed to the magnetic poles via the so called {\\em funnel flow} \\citep{Romanova:2002p4187,Romanova:2003,Kulkarni:2005p4093}; \\item the {\\em unstable} regime, when Rayleigh-Taylor instabilities form at the boundary between the disk and the magnetosphere \\citep{Romanova:2006p2479,Romanova:2008,Kulkarni:2008}, and accretion takes place through tongues of matter from the inner edge of the disk to random places between the equator and the magnetic poles; \\item the {\\em magnetic boundary layer} regime, during which the disk is very close to the surface of the star and the accretion takes place via two big antipodal equatorial tongues which rotate with the frequency of the inner disk \\citep{Romanova:2009p6603}. \\end{itemize} One of the interesting phenomena observed in 3D simulations is the movement of spots around the magnetic pole of the accreting star, both in stable \\citep[][]{Romanova:2003,Romanova:2004p4095,Romanova:2006p3862} and in unstable regime \\citep{Kulkarni:2009}. It was noticed that in some cases, such as in the magnetic boundary layer regime, the moving spots produce QPO features \\citep{Romanova:2009p6603}. Usually, models of NS surface emission (like in the case of Accreting Millisecond Pulsars) consider funneling of matter onto the polar caps as a static process, with a fixed polar cap and thus a fixed emission spot \\\\citep[e.g.][]{Beloborodov:2002p4173,Poutanen:2003p4169}. 3D simulations show that this may not be the case in particular for small misalignment angles. Recent works \\citep{Lamb:2009A,Lamb:2009B} investigate the wandering of hot spots at which matter accretes in sources with a small misalignment angle, finding that it might affect the coherence of the pulsed signal from neutron stars and even destroy it, being the origin of the transient emission in AMXPs. The movements hypothesized in those works, though, are different from the movements observed in our 3D simulations, where the motion of the hotspots is a rotational motion around the magnetic pole or in the equatorial region, on timescales of milliseconds. Moreover, the dimension of the hotspot in those two works (a radius of $\\sim 25^{\\circ}$) does not find confirmation in our simulations. It corresponds to the dimension of the whole ``hot ring'' where our hotspots (much smaller) move (see \\fref{fig:sf-stab-join}). In this paper we show that the observed movements carry with them the information of disk frequencies near the inner disk, and their effect is the production of quasi-periodic oscillations at the relative frequencies. We perform a series of 3D MHD simulations (mostly for the cases of very small $\\mis$) and analyze the rotation of the spots and the related frequencies in detail. MHD simulations can only follow the motion of the disk on very short timescales compared to the disk evolution timescale and the observational time. For this reason, we use the features of the moving spots produced in simulations to generate long lightcurves by means of Monte Carlo techniques, showing that the motion observed produces QPOs. Then, we show that the phenomenon follows a predictable behavior as the mass accretion rate changes. We start in \\sref{sec:model} with a general description of the 3D MHD model used for this work. Then, we show the way spots move in our simulations in \\sref{sec:3d}. We then show with a Monte Carlo-generated lightcurve our model of surface QPO production in \\sref{sec:mc}. Finally, in \\sref{sec:mdot} we show how the angular velocities of hot spots correlate with the mass accretion rate. ", "conclusions": "We developed a model of quasi-periodic oscillations based on the spot movements on the surface of the accreting neutron star with a slightly misaligned dipole magnetic field. 3D MHD simulations have shown that spots may rotate faster/slower than the star in cases of stable or unstable accretion \\citep[see also][]{Romanova:2003,Romanova:2004p4095,Kulkarni:2008,Kulkarni:2009} and in the magnetic boundary layer regime \\citep{Romanova:2009p6603}. We have shown that in all above cases, the spot movements develop ordered oscillations in the light-curves. Modeling these oscillations as short-lived sinusoidal trains, we showed that if they are powerful enough to be seen, their fingerprint is a QPO. Given the range of values of the frequencies involved, we hypothesize that kHz QPOs could be produced in this way. There are several features of QPOs that can give confirmation to this model. First of all, the general trend with the mass accretion rate, as shown in \\sref{sec:mdot}, is as it is expected. Though it is not generally true that luminosity and mass accretion rate are proportional to each other, we can assume it to be true on a given source and for a short timescale. This is the case of the simulations in this paper, where only one parameter (the mass accretion rate) changed between the various simulations. The rms amplitude of real life QPOs has a strong dependence on energy \\citep {Berger:1996p4473}, suggesting that the mechanism of emission is different from that of the background. In our opinion, this is because the QPOs are produced on the surface, or perhaps just above the surface by means of shocks in the accretion column similarly to what happens in AMXPs \\citep[see for example][]{Gierlinski:2002p5829}, while most of the background is produced by the disk. The rms amplitude is also generally higher in atoll sources (up to 20\\%) than in Z sources (2-5\\%) \\citep{Jonker:2001p4470}. The reason for this difference is probably related to the higher accretion rate of Z sources. In some cases 3D simulations show that two frequencies appear at the same time. This numerical discovery is important for possible interpretation of two QPO peaks observed in a number of LMXBs. In fact, this mechanism for the production of multiple QPOs is able to explain more than the mere appearance of two peaks in the Power Spectrum. When double QPOs appear in observations, they normally present different behavior as regards their $Q$ factor \\citep{diSalvo:2003p5128,Barret:2006p7645}, and hence (according to our assumption) their coherence time with the upper peak generally less coherent; in our model, this is due to the fact that the lower one comes from a smooth motion around the magnetic pole, while the upper one is produced by instability tongues, shorter-lasting and thus with a lower coherence. The term ``unstable'' should not mislead the reader though: the unstable regime is very common in simulations, expecially for fast rotators, also at small values of the mass accretion rate, as shown in \\fref{fig:mdot18}. A higher coherence of the upper kHz QPO, in our model, could in principle be produced by a very high accretion rate and the onset of the magnetic boundary layer regime, in which instabilities produce long-lasting opposite tongues in the equatorial plane. During this regime though, it is very probable that the zone around the surface of the neutron star is optically thick. Just as observations show, multiple QPOs can appear simultaneously or at different times. As regards the models predicting the upper limit of $\\nu_u$ as the Keplerian frequency at the ISCO \\citep{Miller:1998p4107}, it would be interesting to investigate the matter further. For the values of mass, magnetic field and radius used in the present job, the ISCO is never reached by the inner disk during QPO production with the two channels described because of the onset of the magnetic boundary layer regime. Further investigation should include the use of more extreme values of the mass in order to study the behavior of the inner disk when the ISCO is considerably far from the surface. Another interesting point is the $\\nu-L_x$ correlation observed in many papers \\citep[for example][]{Yu:1997p4754,Mendez:1998p6564}. The correlation only holds for a given source and on time scales shorter than a few hours. The main reason why it does not hold across sources can easily be found in the magnetic field: for a given value of the mass accretion rate, different values of the magnetic field correspond to a different size of the magnetosphere and, as a consequence, to different values of the frequency of the moving spots. Again, this is evident from our simulations: the exact same frequency in figure \\ref{fig:mdot15} is obtained for $\\mdot\\approx4\\cdot 10^{-11}\\msun y^{-1}$ in a system with $B=10^8$G and for $\\mdot=10^{-9}\\msun y^{-1}$ in a system with $B=5\\cdot10^8$G. But the simulations presented here show only one of the many possible configurations of the disk around the star. For a given source, it is possible that changes in the structure (density distribution for example) of the disk produce a different behavior of the moving hot spots, even with the same mass accretion rate. Besides, the relationship between luminosity and mass accretion rate is not clear. This point will be investigated more deeply in future papers, by simulating systems similar to the one shown here but with different initial disk conditions. The results presented here are based on simulations of neutron stars with very small misalignment angles. An equivalent discussion can be made at slightly larger misalignment angles, where the behavior of the spots does not change considerably (see \\fref{fig:spotvsmis}, but also \\citealt{Kulkarni:2008,Kulkarni:2009,Romanova:2008}), and obtain almost the same features in the power spectrum. In this case though, from simple geometrical considerations it is reasonable to expect some kind of modulation at the frequency of the star. The least we can expect is a broad Lorentzian-like feature at the frequency of the star \\citep{Burderi:1993p328,Lazzati:1997p8008,Menna:2002p949} just from the action of the rotating spot in the polar region, due to the rotation of the star. Its quality factor $Q$ would be comparable to that of the moving hotspots ($Q_*/Q_{spot} \\approx {\\spinf/\\nu_{spot}}$, according to the considerations of \\sref{sec:q}). The observation of such a feature, even very dim, in observations where the lower kHz QPO (the one from the funnel flow in this model) is present, would give interesting evidence in favor of this model. As a final remark, we notice that the difference between the frequencies of the two QPOs is $200-300$Hz, close to the frequency of the star. It is generally known that for relatively slow rotators the frequency difference of QPOs is similar to the frequency of the star. Until about two years ago, it seemed that the $\\Delta\\nu$ was approximately equal to $\\spinf$ for slow rotators and $0.5\\spinf$ for fast rotators. \\citet{Mendez:2007p5756}, analyzing the matter in detail, find that this distinction is not so clear, and that data can be interpreted and fitted equally well by values of $\\Delta\\nu$ around $300$Hz for all sources, with the caveat \\citep{vanStraaten:2005p4594} of a few AMXPs showing all variability components (and thus also $\\Delta\\nu$) shifted down by a factor of $\\sim 1.5$. The range of values of $\\Delta\\nu$ found in this work are compatible with both hypotheses, because the simulated system is a slow rotator and $\\Delta\\nu$ is both near the frequency of the star and $300$Hz. We are planning to investigate the matter with simulations of faster rotators. Nevertheless, this mechanism for the production of double QPOs does not need the two frequencies to be origined from one another, giving a fixed value of $\\delta\\nu$, which was the main caveat of a somewhat similar model, the sonic point beat frequency model by \\citet{Miller:1998p4107}. From what we observe in this work, the two QPOs are two separate effects of the interaction between the rotating magnetosphere and the disk. The funnel flow comes from a zone just behind the inner radius where the magnetic field is strong enough to capture matter and make it drift towards the pole. The instabilities form right at the inner radius, where magnetic field energy density and ram pressure are equal (the green line in \\fref{fig:sf-stab-join}). The results of this paper can give insight into the matter: the frequencies found in observations are a probe of the interaction between the magnetic field and the disk." }, "0911/0911.1878_arXiv.txt": { "abstract": "The virtual observatory (VO) is a collection of interoperable data archives, tools and applications that together form an environment in which original astronomical research can be carried out. The VO is opening up new ways of exploiting the huge amount of data provided by the ever-growing number of ground-based and space facilities, as well as by computer simulations. This presentation summarises a variety of scientific results spanning various fields of astronomy, obtained thanks to the VO, after highlighting its structure, infrastructure and various capabilities. ", "introduction": "\\label{sec:intro} A Virtual Observatory (VO) is a collection of inter-operating data archives and software tools which utilize the Internet to form an environment in which original research can be conducted. The VO is an international astronomical community-based initiative, whose main goal is to allow transparent and distributed access to data available worldwide. This is achieved by developing and applying common standards and by ensuring the interoperability between the various data collections, tools and services. This allows scientists to discover, access, analyze, and combine observational and model data from heterogeneous data collections in a coherent and user-friendly manner. ", "conclusions": "\\label{conclude} The scientific usage of the VO tools and infrastructure is now taking up, and this is reflected by the increasing number of refereed and other publications making use of the VO capabilities. The EURO-VO and other national VO initiative are here to help interested groups to carry out VO-enabled research and provide support for projects that want to publish their data in the VO. Dedicated initiatives such a workshops and Schools are organised world-wide in order to make astronomers aware of the grand potential of the VO and the scientific possibilities it offers." }, "0911/0911.3006.txt": { "abstract": "We present Integral Field Spectroscopy (IFS) of NGC~595, one of the most luminous \\hii\\ regions in M33. This type of observations allows us to study the variation of the principal emission-line ratios across the surface of the nebula. At each position of the field of view, we fit the main emission-line features of the spectrum within the spectral range 3650-6990~\\AA, and create maps of the principal emission-line ratios for the total surface of the region. The extinction map derived from the Balmer decrement and the {\\it absorbed} \\ha\\ luminosity show good spatial correlation with the 24~\\mi\\ emission from {\\it Spitzer}. We also show here the capability of the IFS to study the existence of Wolf-Rayet (WR) stars, identifying the previously catalogued WR stars and detecting a new candidate towards the north of the region. The ionization structure of the region nicely follows the \\ha\\ shell morphology and is clearly related to the location of the central ionizing stars. The electron density distribution does not show strong variations within the \\hii\\ region nor any trend with the \\ha\\ emission distribution. We study the behaviour within the \\hii\\ region of several classical emission-line ratios proposed as metallicity calibrators: while [\\nii]/\\ha\\ and [\\nii]/[\\oiii] show important variations, the \\rdostres\\ index is substantially constant across the surface of the nebula, despite the strong variation of the ionization parameter as a function of the radial distance from the ionizing stars. These results show the reliability in using the R$_{\\rm 23}$ index to characterize the metallicity of \\hii\\ regions even when only a fraction of the total area is covered by the observations. ", "introduction": "Emission-line spectra of Galactic and extragalactic HII regions are normally obtained to characterize their physical properties, to extract information of the stellar populations that ionize them \\citep{Osterbrock:2006p498} as well as to study the variation of the physical properties of the star-forming regions across the galaxy disc (e.g. \\citealt*{McCall:1985p496}; \\citealt{Vilchez:1988p505}; \\citealt{VilaCostas:1992p506}). The \\hii\\ regions are far from the idealized spherical ionized gas clouds with homogenous physical properties, as has clearly been shown in detailed studies of the nearest Galactic and extragalactic \\hii\\ regions: e. g. Orion Nebula (\\citealt{Baldwin:1991p116}; \\citealt{Kennicutt:2000p110}; \\citealt{Sanchez:2007p501}), 30~Doradus (\\citealt{Mathis:1985p115}; \\citealt{Kennicutt:2000p110}) and NGC~604 (\\citealt{GonzalezDelgado:2000p486}; \\citealt{MaizApellaniz:2004p493}), among others. The derived physical properties, generally obtained from long-slit spectroscopic measurements centred on the most intense knots, are not necessarily representative of the conditions in all the locations within the regions (\\citealt{Oey:2000p124}; \\citealt{Oey:2000p121}). The variations of the physical properties are especially seen in star-forming regions where the action of the massive stars strongly influence the surrounding medium (e.g. \\citealt*{Kobulnicky:1996p489}, 1997). Moreover, in some cases the gas within the \\hii\\ regions is ionized by multiple star clusters which are not always centrally located. This will produce different temperature and ionization structures than the normally assumed configurations from spherical models (\\citealt*{Ercolano:2007p482}; \\citealt{Jamet:2008p570}). In order to study the variations of the physical properties within the \\hii\\ regions, the usual technique carried out up to now consists of a set of spectroscopic observations with slits located at different positions in the region and covering as much surface as possible. This technique has limitations however: it requires a large amount of observation time and does not cover, in general, the complete face of the region; thus interpolations of the observational parameters in the gap between slits need to be made. Integral Field Spectroscopy (IFS) overcomes these limitations and offers the opportunity of extracting spectroscopic information at different positions across a continuous field of view. Therefore, this kind of surveys allows the study of the variation of the physical properties within star-forming regions. We present here IFS observations of NGC~595, the second most luminous \\hii\\ region in M33. NGC~595 presents an angular size of $\\sim$1 arcmin, which at the distance of M33 (840~kpc; \\citealt*{Freedman:1991p485}) translates into a linear physical size of $\\sim$200~pc. This makes the region particularly suitable for being studied with the IFS technique. Although the current IFS facilities usually have a relatively small field of view, the proximity of this region (which implies a relatively high surface brightness) makes possible to map it in a reasonable amount of observing time. NGC~595 has an \\ha\\ shell morphology that shows the action of the stellar winds of the massive stars located in its interior. The stellar content studied by \\citet*{Malumuth:1996p494} using optical photometry reveals the existence of $\\sim$250 OB-type stars, $\\sim$13 supergiants and 10 candidate Wolf-Rayet (WR) stars. These authors derived an age of 4.5 $\\pm$ 1.0~Myr for the stellar cluster, which is consistent with the age predicted from the synthesis of integrated spectra in the far-ultraviolet (FUV) wavelength range \\citep{Pellerin:2006p499}. Recently, \\citet{Drissen:2008p481} spectroscopically confirmed nine of the WR candidates previously identified by \\citet*{Drissen:1993p105} using photometric observations and discarded two possible WR candidates from their sample. The physical properties for NGC~595 have been derived with long-slit spectroscopic observations located at the most intense knots within the region \\citep{Vilchez:1988p505} and recently with echelle spectroscopy \\citep{Esteban:2009p563}. \\citet{Vilchez:1988p505} obtained a temperature of T$_{\\rm e} \\sim$ 8000~K, an electron density consistent with the low density limit and a metallicity of 12 + log[O/H] = 8.44 $\\pm$ 0.09 (Z = 0.6\\zsun\\footnote{We assume solar abundance from \\citet*{Asplund:2005p404}, 12 + log[O/H] = 8.66}), which was recently confirmed by \\citet{Esteban:2009p563}. The extinction within the region has been studied previously by other authors: \\citet*{Bosch:2002p478} obtained the extinction using the \\ha/\\hb\\ emission-line ratio and \\citet*{Viallefond:1983p504} derived the extinction with the \\ha\\ and thermal radio emission ratio. \\citet{Relano:2009p558} revised the extinctions using radio-to-\\ha\\ emission and derived new values using the 24~\\mi-to-\\ha\\ integrated emission ratio, finding consistent results from both methods. They also found differences in the emission distributions at several wavelength ranges: while the infrared emission at 24~\\mi\\ and \\ha\\ are spatially correlated with each other, the FUV emission is located within the observed \\ha\\ shell structure of the region. They suggest that the dust emitting at 24~\\mi\\ will probably be mixed with the ionized gas of the region and will be heated by the central ionizing stars. In this paper, we analyse IFS observations covering the whole face of NGC~595. We are able to perform a detailed analysis of the principal emission-line ratios which are widely used to describe the main properties of \\hii\\ regions such as extinction and density structure, ionization parameter, metallicity, existence of shocks and evolutionary state. With these observations, we are in a position to test the influence of the geometry and the stellar distribution within the region on the integrated fluxes of the emission lines, as has been predicted by the models. The comparison of our results with those obtained from long-slit spectroscopy makes it possible to estimate the bias of the observations when only one part of the region, generally the most intense, is covered. Finally, the power of the IFS and the quality of the observations presented here allow us to analyse the WR stellar content of the total surface of the region and to make a complete census of this stellar population in NGC~595. ", "conclusions": "We present IFS of NGC~595, one of the most luminous \\hii\\ regions in M33, covering an unprecedented area in the disc of the spiral galaxies of the Local Group, a $\\sim$174$\\times$340~parsec$^2$ field of view representing the complete surface of the region. Taking advantage of the power of these observations, we are able to identify and catalogue the WR population of the region and to make the best census of WR stars in NGC 595 up till now. We have analysed the variations within the region of the main emission-line ratios that describe the physical properties of the region. The analysis of the observations yields the following results. \\begin{itemize} \\item We present the properties of NGC~595 derived from the integrated spectrum of the region. The electron density and the metallicity are consistent with previously reported values in the literature using long-slit spectroscopy, covering just the most intense knots in the region. \\item The extinction map obtained from the \\ha/\\hb\\ emission-line ratio at each \\emph{spaxel} in the field of view presents a concentrated distribution with the maximum located at the centre of the \\ha\\ shell structure. We have compared this map with the {\\it Spitzer} 24 and 8~\\mi\\ bands using elliptical radial profiles. The 24~\\mi\\ emission and the \\ha/\\hb\\ radial profiles are similar, with their maxima located at the same distance from the ionizing stars. The maximum of the 8~\\mi\\ emission radial profile shows a displacement of $\\sim$4-5\\arcsec\\ (15-18~pc) with respect to the maximum of the \\ha/\\hb\\ radial profile. This shows that the extinction suffered by the gas is produced by dust emitting at 24~\\mi, which is probably mixed with the ionized gas within the region, while the dust emitting at 8~\\mi\\ is more related to the outer parts of the region where the molecular cloud is located. \\item The [\\sii]$\\lambda$6717/[\\sii]$\\lambda$6731 emission-line ratio map does not show any structure, implying that the electron density is quite constant within the region, despite the pronounced \\ha\\ shell morphology of NGC~595. The values of [\\sii]$\\lambda$6717/[\\sii]$\\lambda$6731 are consistent with the low electron density limit. \\item We have produced BPT diagrams using the emission-line ratios at each \\emph{spaxel} in the field of view. We find that the \\emph{spaxels} with a low \\ha\\ flux cover the total excitation range of the region. The location of the \\emph{spaxels} with a high \\ha\\ flux in the [\\oiii]/\\hb-[\\sii]/\\ha\\ diagram is displaced with respect to the position given in the diagram for the integrated values. This shows the limitations of the long-slit observations, which usually cover the most intense \\ha\\ knot in the regions, to obtain a reliable value for the [\\sii]/\\ha\\ emission-line ratio. \\item The study of the WR population for NGC~\u00ca595 has been completed with these observations. We have covered the total surface of the region and found a new WR candidate whose spectrum is characteristic of a WN star. The WR candidate, situated far away from the other WR stars in the region, is close to the \\ha\\ shell but corresponds to a zone of low \\ha\\ emission. \\item We have analysed the behavior of the most used metallicity calibrators: \\rdostres\\ shows small variations of 0.13~dex despite the spatial distribution of the stars and gas in the region and the strong trend of [\\oiii]/[\\oii] to decrease to the outer parts of the region. This shows the robustness of \\rdostres\\ to estimate the metallicity of the region; in spite of the \\hii\\ region morphology, far away from the classical Str\\\"omgren sphere picture, the \\rdostres\\ parameter varies slightly within the region and thus can be used to obtain a representative value of the \\hii\\ region metallicity. In contrast, other parameters such as [\\nii]/\\ha\\ and [\\nii]/[\\oiii] show strong variations within the region (up to one order of magnitude in the case of [\\nii]/[\\oiii]), which show their strong dependence on the ionization parameter. \\end{itemize} We show in this paper the power of the IFS, which overcomes the limitations of long-slit observations in \\hii\\ regions. The results presented here show that for the emission lines dominated by the contribution of the most intense knots, long-slit observations are representative of the integrated emission of \\hii\\ regions. The situation is different in the case of lower excitation emission lines, as it happens in the [\\sii] emission lines, for which there is a significant bias for long-slit observations. This bias has only been possible to quantify using IFS observations with a complete \\hii\\ region coverage. The physical properties can vary within \\hii\\ regions and the emission-line ratio maps derived from IFS observations allow us to make a detailed study of these variations. For NGC~595 we find that the ionization parameter shows strong variations with the radial distance from the stars; the ionization structure is clearly depicted with the corresponding emission-line diagnostics and the reddening map presents a non-uniform distribution with a maximum located at the center of the observed \\ha\\ shell structure. The electron density map, however, does not present significant variations within the region. A change of one order of magnitude in [\\oii]/[\\oiii] within NGC~595 allows us to study the dependence of the ionization parameter on the most widely used metallicity calibrators, since other properties such as density or the ionizing spectrum do not change within this \\hii\\ region. The main result is that \\rdostres\\ does not depend significantly on the ionization parameter, but other emission-line ratios such as [\\nii]/\\ha\\ and [\\nii]/[\\oiii] show important variations within the region. Finally, the capability of the IFS data to make complete censuses of the WR population in stellar clusters is very nicely shown here: while the hunting of WR stars are based on identification in broad-band images with follow-up spectroscopic observations, we are able to perform such an analysis in only one night of observation, covering the whole surface of the region and allowing us to identify WR stars that are not located close to the centre of the cluster." }, "0911/0911.1080_arXiv.txt": { "abstract": "From observations collected with the ESPaDOnS spectropolarimeter at the Canada-France-Hawaii Telescope (CFHT), we report the detection of Zeeman signatures on the low-mass classical T~Tauri star (cTTS) V2247~Oph. Profile distortions and circular polarisation signatures detected in photospheric lines can be interpreted as caused by cool spots and magnetic regions at the surface of the star. The large-scale field is of moderate strength and highly complex; moreover, both the spot distribution and the magnetic field show significant variability on a timescale of only one week, as a likely result of strong differential rotation. Both properties make V2247~Oph very different from the (more massive) prototypical cTTS BP~Tau; we speculate that this difference reflects the lower mass of V2247~Oph. During our observations, V2247~Oph was in a low-accretion state, with emission lines showing only weak levels of circular polarisation; we nevertheless find that excess emission apparently concentrates in a mid-latitude region of strong radial field, suggesting that it is the footpoint of an accretion funnel. The weaker and more complex field that we report on V2247~Oph may share similarities with those of very-low-mass late-M dwarfs and potentially explain why low-mass cTTSs rotate on average faster than intermediate mass ones. These surprising results need confirmation from new independent data sets on V2247~Oph and other similar low-mass cTTSs. ", "introduction": "\\label{sec:int} Whereas our understanding of most phases of stellar evolution made considerable progress throughout the twentieth century, stellar formation remained rather enigmatic and poorly constrained by observations until about two decades ago with the advent of the most powerful ground-based and space telescopes. One of the major discoveries obtained then is that protostellar accretion discs are often associated with extremely powerful and highly collimated jets escaping the systems along their rotation axis \\citep[e.g.,][]{Snell80}. This finding has revolutionized the field of stellar formation; in particular, it provided us with strong evidence that magnetic fields are playing an important role throughout stellar formation. Magnetic fields are expected to modify significantly the contraction of molecular clouds into protostellar cores and discs \\citep[e.g.,][for a review]{Andre08}. They are also expected to influence the next formation stage, at least for stars with masses lower than about 2.5~\\msun; these protostars (called classical T~Tauri stars or cTTSs) apparently host magnetic fields strong enough to disrupt the central regions of their accretion discs, connect the protostars to the inner disc rim through discrete accretion funnels and possibly even slow down the rotation of protostars through the resulting star/disc magnetic torque \\citep[e.g.,][for a review]{Bouvier07}. While magnetic fields have been repeatedly reported over the last decade at the surface of a number of prototypical TTSs \\citep[e.g.,][]{Johns99b, Johns07}, it is only recently that their large-scale topologies can be investigated into some details through phase-resolved spectropolarimetric observations \\citep{Donati07, Donati08b, Hussain09}. This new option offers a direct opportunity to pin down the role of magnetic fields in this crucial phase of the formation process. In particular, it straightforwardly allows studying how magnetic topologies of cTTSs depend on mass and rotation rate, giving hints on the potential origin of such fields; moreover, it provides material for quantitatively studying how protostars magnetically connect and interact with their accretion discs. Up to now, magnetic topologies of 4 cTTSs have been investigated with this method. Among these 4, only BP~Tau ($\\mstar\\simeq0.7$~\\msun) is fully convective and hosts a rather simple, dominantly dipolar, large-scale magnetic field, the 3 others being partly convective (as a result of their relatively higher mass, in excess of 1.3~\\msun) and hosting a more complex magnetic topology. In particular, this transition from simple to complex fields for stars on either sides of the full convection limit is reminiscent of the magnetic properties of low-mass main-sequence dwarfs \\citep{Morin08b, Donati08d}; in addition to suggesting that magnetic fields of cTTSs are likely of dynamo origin, it also gives potential hints on why magnetospheric accretion and rotation properties of cTTSs are discrepant on both sides of the full convection limit (e.g., with higher mass cTTSs rotating more quickly in average). Extending this magnetic survey to a larger sample of cTTSs is needed to go further; this is the exact purpose of the MaPP (Magnetic Protostars and Planets) program, in the framework of the international MagIcS research initiative (aimed at qualifying the magnetic properties of stars throughout the HR diagram). This survey has been allocated 690~hr of observing time on the 3.6~m Canada-France-Hawaii Telescope (CFHT) over 9 successive semesters (from 2008b to 2012b). The present paper concentrates on the low-mass\\footnote{Throughout this paper, cTTSs with masses lower than 0.5~\\msun, ranging from 0.5 to 1~\\msun\\ and larger than 1~\\msun\\ are respectively called low-mass, intermediate-mass and high-mass cTTSs. } cTTS V2247~Oph (spectral type M1) in the $\\rho$~Oph star forming region (Lynds~1688 dark cloud); with a mass of only about 0.35~\\msun\\ (see Sec.~\\ref{sec:v22}), it is less massive than all cTTSs yet magnetically imaged and thus appears as an ideal candidate to expand our sample towards cooler cTTSs. ", "conclusions": "\\label{sec:dis} We presented in this paper the first magnetic map for the low-mass cTTS V2247~Oph. This is the lowest mass cTTS to be magnetically imaged yet; despite it being in a low-state of accretion (with $\\log \\Mdot \\simeq -9.8$ with \\Mdot\\ expressed in \\mspy), this study brings a number of new and unexpected results. The brightness distribution we derive at the surface of V2247~Oph (from the rotational modulation of Stokes $I$ photospheric LSD profiles) is relatively simple, with dark spots covering about 3.5\\% of the total surface; in particular, it does not feature a highly-contrasted dark polar spot as seen, e.g., on V2129~Oph \\citep{Donati07}, BP~Tau \\citep{Donati08b} or other cTTSs \\citep{Hussain09}. It is however roughly similar to what is found on low-mass fully-convective dwarfs where large-scale spots are usually low-contrast and cover only a small fraction of the stellar surface \\citep[e.g.,][]{Morin08a}. We also detect clear Zeeman signatures from V2247~Oph, tracing a rather complex and moderately strong multipolar large-scale magnetic topology with an average strength of 300~G and a mixed amount of poloidal and toroidal field; the poloidal field is mostly non-axisymmetric and features a weak (tilted) dipole component of about 80--100~G. Again, this is fairly different from what is seen on intermediate-mass cTTSs like BP~Tau, where the field is far simpler, predominantly poloidal, axisymmetric with a dipole component exceeding 1~kG \\citep{Donati08b}. This is also in contrast to low-mass fully-convective mid-M dwarfs \\citep[e.g.,][]{Morin08b} that almost always exhibit very intense, mainly potential large-scale fields with a simple topology roughly aligned with the rotation axis. New results however indicate that a significant fraction of very-low-mass dwarfs (late-M dwarfs with $\\mstar<0.2$~\\msun, i.e., below about half the mass threshold for full convection on the main sequence) are found to host moderate and complex non-axisymmetric large-scale magnetic fields (Morin et al 2009, in preparation), very different from the strong and simple fields of mid-M dwarfs. Although V2247~Oph is significantly more massive (at 0.35~\\msun) than late-M dwarfs, the difference in evolutionary stage may easily compensate for the discrepancy; the mass of V2247~Oph is indeed also lower than half the mass threshold for full convection at an age of about 1~Myr. Further observations are of course needed to confirm whether this analogy truly holds; if so, it would bring additional evidence that magnetic fields of intermediate- and low-mass cTTSs are similar in nature to those of mid-M and late-M dwarfs and are thus produced through dynamo processes (rather than being fossil leftovers from an earlier formation stage). In addition, we find that the spot distribution and the magnetic field of V2247~Oph evolve on a very short timescale, of order one week; in both cases, this evolution is compatible with surface differential rotation shearing the photosphere 5 to 6 times faster than on the Sun, with the equator lapping the pole by one complete rotation every $19\\pm2$~d. This is again fairly unexpected since all fully-convective stars have been reported to show little to no differential rotation up to now \\citep[e.g.,][]{Donati06a, Morin08a, Morin08b}. Only late F stars with very shallow convective zones are yet known to exhibit a similar degree of photospheric shear \\citep[e.g.,][]{Donati08c}. Further confirmation from new data sets on V2247~Oph (and other similar low-mass cTTSs) are thus needed to validate this surprising result. The amount of differential rotation we estimate is compatible with the range of photometric periods measured on V2247~Oph over the last 2 decades \\citep[from 3.4 to 3.6~d, implying $\\dom>0.1$~\\rpd,][]{Grankin08}. Given the differential rotation law we derive, these periods suggest that spots preferentially cluster at low to intermediate latitudes (15--40\\degr) on V2247~Oph; this is at least qualitatively compatible with the fact that no high-contrast polar spot is present at the surface. This is the first time that differential rotation is detected on a cTTS; the recent claim \\citep{Herbst06} that the high-mass cTTS HBC~338 (spectral type G9) is differentially rotating (given the observed variations of the photometric period) needs confirmation, cTTSs with massive accretion discs being prone to photometric perturbations (including sudden changes of the light-curve period) likely caused by accretion \\citep[e.g.,][]{Simon90, Donati08b} rather than resulting from differential rotation. Finally, we find that V2247~Oph features a region of excess IRT (and \\hal\\ and \\hbe) emission at phase 0.45 and intermediate latitude. There is no obvious correlation between the location of this emission region and the cool spots mapped from the distorted LSD profiles of photospheric lines; however, we find that it roughly coincides with one of the strongest field region detected on V2247~Oph, the positive radial field spot whose magnetic flux reaches about 0.5~kG. We speculate that this spot may be the footpoint of an accretion funnel linking the star to its accretion disc at the time of our observations. If confirmed, we expect such accretion funnels to be rather short lived, with differential rotation continuously distorting the magnetic connections between the star and the disc as well as the stellar field itself. In particular, this effect may be partly responsible for the sporadic accretion episodes observed on V2247~Oph, with more intense accretion episodes occurring when the magnetic topology linking the star to the disc is favouring accretion. In addition to the differences in their magnetic topologies, V2247~Oph and BP~Tau show quite distinct rotational properties; as emphasised in Sec.~\\ref{sec:v22}, each of them belongs to a distinct population of young stars \\citep{Lamm05}, V2247~Oph to the low-mass cTTSs with short rotation period (of about half a week in average) while BP~Tau to the intermediate-mass cTTSs with long rotation periods (of about one week in average). These two issues may relate together, i.e., the magnetic topology may have an impact on the rotation period through a coupling mechanism such as, e.g., the well-known disc-locking picture \\citep{Camenzind90, Konigl91}. In this model, the rotation period at the surface of the star settles close to the Keplerian period of the point in the disc (called the Alfven radius and noted \\rmag) where the ram pressure of the average accretion flow roughly equals the magnetic pressure of the largest-scale (i.e., the dipole component) poloidal field \\citep[e.g.,][]{Bessolaz08}. Using Eq.~1 of \\citet{Bessolaz08} (with $B_*$ set to the average dipole field strength at the rotation equator, i.e., about 40~G for V2247~Oph), we estimate $\\rmag\\simeq3.9$~\\rstar; we can also work out that the radius at which the Keplerian period matches the average rotation period (called the corotation radius and noted \\rcor) is equal to $\\rcor\\simeq3.4$~\\rstar\\ for V2247~Oph. Both estimates roughly agree with each other, suggesting a reasonable agreement with the disc-locking model predictions. Comparing directly with BP~Tau however \\citep[and assuming a logarithmic mass accretion rate of --7.5 taken from the literature as in][]{Donati08b}, the model predicts that both stars should have similar \\rmag\\ and thus behave similarly with respect to the disc-locking mechanism, the 15-fold weaker dipole field of V2247~Oph being almost compensated by the 200-fold weaker mass accretion rate; re-estimating $\\log \\Mdot$ for BP~Tau using the same procedure as for V2247~Oph, we actually find a much smaller value, equal to $-8.6\\pm0.4$. While it is not clear whether this new estimate is more accurate than the one used previously, it is at least more consistent with the one we derived for V2247~Oph when it comes to comparing both stars; in this case, we obtain that $\\rmag\\simeq7.1$~\\rstar\\ for BP~Tau (taking $B_*\\simeq600$~G) in reasonable agreement with $\\rcor\\simeq7.4$~\\rstar. Obviously, more stars of different masses and rotation periods, and with consistent estimates of $B_*$ and \\Mdot, are needed to assess the validity of the disc-locking picture; if confirmed, it may ultimately explain why low-mass cTTSs rotate in average faster than their intermediate mass equivalents. The MaPP program should bring new results along this line in the near future." }, "0911/0911.0423_arXiv.txt": { "abstract": "We present a sample of \\noptall\\ optically selected \\bl\\ candidates from the SDSS DR7 spectroscopic database encompassing 8250~deg$^2$ of sky; our sample constitutes one of the largest uniform BL Lac samples yet derived. Each \\bl\\ candidate has a high-quality SDSS spectrum from which we determine spectroscopic redshifts for $\\sim$60\\% of the objects. Redshift lower limits are estimated for the remaining objects utilizing the lack of host galaxy flux contamination in their optical spectra; we find that objects lacking spectroscopic redshifts are likely at systematically higher redshifts. Approximately 80\\% of our \\bl\\ candidates match to a radio source in FIRST/NVSS, and $\\sim$40\\% match to a ROSAT X-ray source. The homogeneous multiwavelength coverage allows subdivision of the sample into \\noptrl\\ radio-loud \\bl\\ candidates and \\noptrq\\ weak-featured radio-quiet objects. The radio-loud objects broadly support the standard paradigm unifying \\bl\\ objects with beamed radio galaxies. We propose that the majority of the radio-quiet objects may be lower-redshift ($z<2.2$) analogs to high-redshift weak line quasars (i.e., AGN with unusually anemic broad emission line regions). These would constitute the largest sample of such objects, being of similar size and complementary in redshift to the samples of high-redshift weak line quasars previously discovered by the SDSS. However, some fraction of the weak-featured radio-quiet objects may instead populate a rare and extreme radio-weak tail of the much larger radio-loud \\bl\\ population. Serendipitous discoveries of unusual white dwarfs, high-redshift weak line quasars, and broad absorption line quasars with extreme continuum dropoffs blueward of rest-frame 2800~\\AA\\ are also briefly described. ", "introduction": "BL~Lacertae objects compose an especially rare subclass of Active Galactic Nuclei (AGNs) traditionally interpreted as low-luminosity radio galaxies with a relativistic jet pointed toward the observer \\citep[e.g., see][]{blandford78, urry95}. Emission from the relativistically boosted jet dominates the observed flux, yielding featureless (or nearly featureless) optical spectra, strong radio and X-ray emission, strong polarization, flat radio spectra, and rapid variability across the entire electromagnetic spectrum \\citep[e.g., see][]{kollgaard94,perlman01}. Given their odd spectral characteristics, and the need to obtain optical spectra to confirm a candidate as a \\bl\\ object, the most efficient \\bl\\ selection algorithms tend to search for optical counterparts to strong radio and/or X-ray sources \\citep[e.g., see][]{stickel91,stocke91,perlman98,laurent99,landt01,anderson07,padovani07,piranomonte07,turriziani07,plotkin08}. Optically selected \\bl\\ samples have historically been extraordinarily difficult to assemble, but they are highly desired. Systematic searches based only on optical characteristics would allow the exploration of \\bl\\ properties with minimal concern for biases introduced by radio and X-ray surveys (which generally tend to have shallower flux sensitivities than optical surveys.) Also, optically selected samples might reveal new populations of \\bl\\ objects, and they might even lead to serendipitous discoveries of other, perhaps new, classes of featureless objects. Unfortunately, the odd spectral characteristics of \\bl\\ objects conspire to make selection based solely on optical colors ineffective. For example, only 4 \\bl\\ objects were identified in the Palomar-Green Survey for ultraviolet-excess objects \\citep[PG,][]{green86}; another 3 PG sources initially misclassified as DC (i.e., featureless) white dwarfs were identified as \\bl\\ objects only after the {\\it ROSAT} All Sky Survey \\citep[RASS,][]{voges99,voges00} came online \\citep{fleming93}. Attempts have been made to exploit other distinctive \\bl\\ properties, like polarization and flux variability, to derive optically selected samples. Such techniques, however, require observing large areas of the sky at multiple epochs just to assemble manageable lists of candidates suitable for spectroscopic follow-up. For example, it would seem that the strong polarization of \\bl\\ objects ($P\\gtrsim2\\%$) should allow for efficient identification of \\bl\\ candidates. However, most attempts along these lines have so far been unsuccessful \\citep{impey82_optsel,borra84,jannuzi93}, probably because the polarized flux is variable with a low duty cycle: for example, X-ray selected \\bl\\ objects from the {\\it Einstein} Observatory Extended Medium-Sensitivity Survey \\citep[EMSS,][]{stocke91} spend over half their time with $P<4\\%$ \\citep[][]{jannuzi94}. Not until we entered the age of large-scale spectroscopic digital sky surveys has optical selection of \\bl\\ objects been a realistic endeavor. Traditionally, the primary limitation of efficient \\bl\\ recovery algorithms is the large number of false positives returned, even when radio and/or X-ray information is consulted. Optical spectroscopy is then required to make firm \\bl\\ identifications. However, with a digital survey such as the Sloan Digital Sky Survey \\citep[SDSS,][]{york00}, this task is less daunting because the requisite spectra already exist. Furthermore, the most relevant spectra can be searched within databases, allowing for automated removal of many of the contaminants that previously had to be culled manually (after a significant investment in telescope time.) The first successful attempt at assembling an optically selected sample is the 2QZ \\bl\\ survey \\citep{londish02,londish07}, derived from the 2-degree field (2dF) and the 6-degree field (6dF) quasi-stellar object Redshift Surveys \\citep{croom04}. They searched over $\\sim$10$^3$~deg$^2$ and recovered 7 confident \\bl\\ objects. All 7 objects were additionally identified with radio and X-ray sources (post-selection), and they also showed optical flux variations in a photometric monitoring campaign over 2002--2004 \\citep{nesci05}. An additional object ($z=0.494$) is identified as a radio-quiet/weak \\bl\\ candidate \\citep{londish04}. This source lacks both a radio and an X-ray detection (with a radio to optical flux ratio limit placing it firmly in the radio-quiet regime\\footnote{Radio-quiet quasars are commonly defined as having radio to optical flux ratios $R<10$ \\citep[e.g., see][]{kellermann89, stocke92}. Throughout this paper, we adopt the approximately similar $\\alpha_{ro}<0.2$, where $\\alpha_{ro}$ is the radio-to-optical broad-band spectral index: $\\alpha_{ro}=-\\log(L_o/L_r)/5.08$, where $L_o$ and $L_r$ are optical and radio specific luminosities at rest-frames 5000~\\AA\\ and 5~GHz, respectively \\citep{stocke85}.}), % but its optical properties are similar to (radio-loud) \\bl\\ objects. The recovery of only a single extragalactic object with a featureless spectrum lacking radio emission by the 2QZ survey confirms the \\citet{stocke90} result that radio-quiet \\bl\\ objects must be extremely rare if they even exist at all \\citep[also see][]{jannuzi93}. Searching 2860~deg$^2$, \\citet{collinge05}, hereafter \\citetalias{collinge05}, assembled a larger optically selected sample from the SDSS. Their sample contains 386 nearly featureless spectra, including 240 {\\it probable} \\bl\\ candidates and 146 {\\it possible} \\bl\\ candidates (the latter are most likely DC white dwarfs, proper motions are consulted for their classification.) The vast majority of the {\\it probable} candidates either have firm extragalactic redshifts, or they match to a radio source in the Faint Images of the Radio Sky at Twenty-cm survey \\citep[FIRST,][]{becker95} and/or the NRAO VLA Sky Survey \\citep[NVSS,][]{condon98}, or to a RASS X-ray source (correlations to multiwavelength catalogs are performed post-selection.) \\citetalias{collinge05} also present a list of 27 intriguing radio-quiet/weak \\bl\\ candidates. These objects have nearly featureless optical spectra, and their radio fluxes (or limits in the cases of radio non-detections) place their radio to optical flux ratios within or tantalizingly close to the radio-quiet regime. All but one of these also lack X-ray detections in RASS (but their extant X-ray limits are not sensitive enough to declare them X-ray weak.) Some of these objects are at high-redshift ($z>2.2$), and those high-redshift objects may be alternately described as members of a population of high-redshift weak line quasars (WLQs) discovered by the SDSS \\citep[e.g., see][]{fan99,anderson01,shemmer06,shemmer09,diamond09}. Three of these high-redshift objects were detected in follow-on {\\it Chandra} X-ray observations by \\citet{shemmer09}, and they are X-ray weaker than typical radio-loud \\bl\\ objects. Other radio-quiet \\bl\\ candidates in \\citetalias{collinge05} have lower redshifts, and it is unclear if they are more closely related to \\bl\\ objects or to WLQs. Regardless of their true nature, \\citetalias{collinge05} demonstrate that optical selection of \\bl\\ objects from the SDSS is efficient, producing a large enough sample to reveal especially rare sub-populations of objects with featureless spectra in numbers comparable to entire venerable radio and X-ray selected \\bl\\ samples. Here, we expand on \\citetalias{collinge05} and continue the search for \\bl\\ objects in the SDSS (which now covers almost three times the sky area as the \\citetalias{collinge05} sample). We recover \\noptall\\ objects. Our automated selection algorithm is described in \\S \\ref{sec:ch4_sampsel}. We visually inspect $\\sim$23,000 spectra, and we discuss the removal of various classes of contaminants (including stars, post-starburst galaxies, quasars, etc.)\\ in \\S \\ref{sec:ch4_visinspec}; we also present lists of serendipitously discovered unusual white dwarfs, WLQs, and unusual broad absorption line quasars. In \\S \\ref{sec:ch4_postvisinspec} we describe the removal of contaminants that survive visual inspection (including featureless white dwarfs and large red galaxies.) The final sample is presented in \\S \\ref{sec:ch4_finalsamp}, and we describe a spectral decomposition of the \\bl\\ host galaxy from the AGN. We use this decomposition to measure optical spectral indices for the AGN, to estimate lower redshift limits by assuming \\bl\\ host galaxies are standard candles, and to calculate decomposed AGN component optical fluxes and luminosities. We also correlate the \\noptall\\ \\bl\\ candidates to the FIRST and NVSS radio surveys, and to the RASS X-ray survey. The sample is discussed in \\S \\ref{sec:ch4_discussion}, which focuses on a particularly interesting subset of \\noptrq\\ radio-quiet objects with weak spectral features. Finally, our main conclusions are summarized in \\S \\ref{sec:ch4_summary}. Throughout, we use {\\it quasar} to refer to any AGN (regardless of luminosity or radio loudness and including \\bl\\ objects), and we adopt $H_0= 71$~km s$^{-1}$~Mpc$^{-1}$, $\\Omega_m=0.27$, and $\\Omega_{\\Lambda}=0.73$. ", "conclusions": " 1.\\ Among the tens of thousands of contaminants removed during visual inspection, we note potentially interesting serendipitous discoveries, including 17 unusual white dwarfs, 9 unusual BALQSOs with extreme continuum dropoffs blueward of \\ion{Mg}{2}, and \\tabnwlq\\ high-redshift WLQs (with reliable $z>2.2$ and Ly$\\alpha$+\\ion{N}{5} $REW<10$~\\AA). Six of these WLQs are additionally classified as \\bl\\ candidates because their emission features have $REW<5$~\\AA. 2.\\ Approximately 60\\% of our \\bl\\ candidates have spectroscopic redshifts (or limits from \\ion{Mg}{2} absorption). For objects lacking spectroscopic redshifts, we assume \\bl\\ host galaxies are standard candles, and we place redshift limits to those objects utilizing the fact that we do not detect host galaxy contamination to their SDSS spectra. We estimate their ``host galaxy'' redshift limits to be accurate to $\\sigma_z = \\pm0.064$. We thus have redshift information (either from weak spectral features or limits placed from the lack of a host galaxy detection) for {\\it every} single object in our sample. We infer the objects lacking spectroscopically-derived redshifts are at systematically higher redshift. We expect this result, since the objects with the weakest spectral features are arguably among the most highly beamed, which requires a low probability geometry. So, we are more likely to find the most highly beamed objects at higher redshift, where more volume is probed. However, there may also be some examples of highly beamed nearby objects in this sample as well. 3.\\ Each object emerges with homogeneous multiwavelength data coverage, and we include radio/X-ray fluxes (or limits) for all objects. Around 80\\% of the \\bl\\ candidates presented here match to a radio source in FIRST or NVSS, and we place upper limits on radio flux densities for the rest (except for \\noptallNoRadioInfo\\ objects outside the FIRST footprint and not detected by NVSS.) Approximately 40\\% match to an X-ray source in RASS, and we place upper X-ray flux limits for the rest. 4.\\ We use each object's high-quality SDSS spectrum to perform spectral decompositions of each object's host galaxy and nucleus, and we report AGN component optical fluxes and luminosities. 5.\\ Based on radio fluxes and limits from the FIRST/NVSS radio surveys, we subdivide our sample into \\noptrl\\ radio-loud \\bl\\ candidates and \\noptrq\\ radio-quiet weak-featured spectra. There is significant overlap between the optical properties of the radio-loud and radio-quiet objects. However, the radio-quiet objects tend to be at higher redshift, and they also have a smaller X-ray detection rate. We argue that the $\\alpha_{ro}<0.2$ objects are a mix of different populations: some of these objects (especially ones lacking spectroscopic redshifts or with small redshifts) probably populate the radio-weak tail of the much larger radio-loud \\bl\\ population. Other sources (especially the ones at higher redshifts) are likely WLQs. The very high-redshift nature of published SDSS WLQs is a selection effect (because they are identified by the presence of the Ly$\\alpha$ forest), and this study provides a systematic search of the SDSS database capable of revealing lower-redshift analogs in large numbers. Even just the $z<2.2$ radio-quiet objects constitute, to our knowledge, the largest sample of such weak-featured radio-quiet objects. The sample size is larger than many entire venerable \\bl\\ samples, and this subset is of similar size and complementary to the number of known high-redshift SDSS WLQs. 6.\\ The radio-loud objects in general support the standard unification paradigm, which maintains that radio and X-ray selected \\bl\\ objects populate extreme ends of a single, larger, and continuous \\bl\\ population. We recover a mix of LBLs, IBLs, and XBLs, although the SDSS appears biased toward finding HBLs. However, we note that the large observed scatter in \\bl\\ properties (e.g., the distribution of their AGN optical spectral indices) is probably due to innate differences as well. Even just the \\noptrl\\ radio-loud objects constitute one of the largest samples of \\bl\\ objects derived from a single selection technique. 7.\\ It will be interesting to assemble light curves for these \\bl\\ candidates once large-scale time-domain surveys, like the Large Synoptic Survey Telescope \\citep[LSST,][]{ivezic08_ph} and the Panoramic Survey Telescope \\& Rapid Response System (Pan-STARRS\\footnote{\\url{http://pan-starrs.ifa.hawaii.edu}}), come online. It may even be possible to use flux variability from those surveys as a \\bl\\ selection criterion. For example, \\citet{bauer09} recently found $\\sim$3000 objects in the Palomar-QUEST survey identified by their optical flux variability; many of these objects may be blazars. Time-domain \\bl\\ searches over huge areas of the sky requires telescope access that is unrealistic with standard non-survey telescopes, but well within the capabilities of LSST and Pan-STARRS." }, "0911/0911.1205_arXiv.txt": { "abstract": "{Is the Doppler interpretation of galaxy redshifts in a Friedmann-Lema\\^{\\i}tre-Robertson-Walker (FLRW) model valid in the context of the approach to comoving spatial sections pioneered by de~Sitter, Friedmann, Lema\\^{\\i}tre and Robertson, i.e. according to which the 3-manifold of comoving space is characterised by both its curvature and topology?} {Holonomy transformations for flat, spherical and hyperbolic FLRW spatial sections are \\postrefereechanges{proposed}.} {\\postrefereechanges{By quotienting a simply-connected FLRW spatial section by an appropriate group of holonomy transformations, the} Doppler interpretation in a non-expanding Minkowski space-time, obtained via four-velocity parallel transport along a photon path, is found to imply that an inertial observer is receding from herself at a speed greater than zero, implying contradictory world-lines. The contradiction \\postrefereechanges{in the multiply-connected case} occurs for arbitrary redshifts in the flat and spherical cases, and for certain large redshifts in the hyperbolic case.} {The link between the Doppler interpretation of redshifts and cosmic topology can be understood physically as the link between parallel transport along a photon path and the fact that the comoving spatial geodesic corresponding to a photon's path \\postrefereechanges{can be} a closed loop in an FLRW model of any curvature. Closed comoving spatial loops are fundamental to cosmic topology.} ", "introduction": "Much debate has recently taken place regarding the interpretation of the redshifts of comoving galaxies in cosmological models as a special-relativistic Doppler effect in the absence of the concept of expanding space \\citep{Chod07a,Barnes06,Chod07b,Abram07,Francis07,Lewis08,BunnHogg09,Peacock08,Abram09,Chod08,Faraoni09}. Here, it is shown that in the flat and spherical Friedmann-Lema\\^{\\i}tre-Robertson-Walker (FLRW) models, where galaxies are comoving massless test objects in the standard comoving-coordinate system, the Doppler interpretation in non-expanding Minkowski space-time leads to a contradiction for galaxies at distances from the observer that are arbitrarily small, provided that these galaxies are considered to be comoving. The corresponding (weaker) argument in hyperbolic FLRW models is also presented. In \\SSS\\ref{s-cos-top}, some mathematical properties of the comoving spatial sections in FLRW models are recalled. The methods of using these properties in flat, positively curved, and negatively curved FLRW models are presented in \\SSS\\ref{s-meth-zero}, \\SSS\\ref{s-meth-pos}, and \\SSS\\ref{s-meth-neg}, respectively. The consequences are described in \\SSS\\ref{s-results}. The way that density perturbations modify these consequences is discussed in \\SSS\\ref{s-almostflrw}. The failure of the Doppler interpretation to arbitrarily small separations for non-negatively curved exact-FLRW models may seem to be inconsistent with the Minkowski nature of FLRW models in the limit towards a point in space-time. This paradox is explained in \\SSS\\ref{s-Mink-limit}. A definition of ``expansion of space'' and related statements from the literature are briefly discussed in \\SSS\\ref{s-expanding}, and the notion of an expanding Minkowski space-time is mentioned in \\SSS\\ref{s-expanding-Mink}. Conclusions are presented in \\SSS\\ref{s-conclu}. For clarity, the terms ``hyperbolic'' and ``spherical'' are used for negatively and positively curved FLRW models, respectively, rather than the ambiguous terms ``open'' and ``closed''. The term ``galaxy'' is used for an external massless galaxy located at some non-zero distance from the observer in the covering space. The observer should be considered to be located in our (massless) Galaxy. The ``galaxy'' and the ``Galaxy'' are massless in order not to violate homogeneity. \\postrefereechanges{A ``comoving spatial geodesic'' is distinct from a space-time geodesic. The former refers to a geodesic (a curve that minimises the metric distance between any close pair of points on that curve) of a comoving spatial section (where the metric only has a spatial component). It can also be thought of as a space-time geodesic (e.g. the path of a photon) projected to a comoving spatial section by ignoring the cosmological time coordinate. For example, in a positively curved FLRW model, a comoving spatial geodesic is an arc that is part of a great circle of the hypersphere $S^3$, since great circles are straight lines.} \\postrefereechanges{For a short introduction to the topology of FLRW models and observational approaches to measuring it, see \\cite{Rouk00BASI}. For reviews, see \\cite{LaLu95,Lum98,Stark98,LR99,BR99,RG04}. Analyses of WMAP data presently include analyses suggesting that sub-matter-horizon cosmic topology has been detected \\citep[][and references therein]{Gundermann2005,Caillerie07,RBG08,Aurich09a}. and analyses disfavouring it \\citep[][and references therein]{KeyCSS06,NJ07}. Key terminology regarding the comoving spatial section, which is a 3-manifold, includes the following, using the 3-torus as an example. The manifold $M = T^3$ can be thought of as a cube, the ``fundamental domain'', with identified faces. It can also be thought of as a tiling of Euclidean 3-space, $\\mathbb{R}^3$, called the ``covering space'', $\\widetilde{M}$, which can also be informally called the ``apparent space'', since it contains many copies of any individual physical object. A mapping $f$ from one copy of the cube in $\\widetilde{M}$ to another copy is a (non-arbitrary) isometry of $\\widetilde{M}$, i.e. it preserves metric distances, and is called a ``holonomy transformation''. The set of holonomy transformations for a given manifold $M$ is the group $\\Gamma$. This group has the same structure as the group of all possible closed comoving spatial paths in $M$ that cannot be continuously transformed into one another. The latter group is called the first homotopy group, $\\pi_1(M)$. Switching from $\\widetilde{M}$ to $M = \\widetilde{M}/\\Gamma$ can be described as ``quotienting'' $\\widetilde{M}$ by $\\Gamma$.} ", "conclusions": "\\label{s-conclu} Cosmic topology is not just a luxury referred to by de~Sitter, Friedmann, Lema\\^{\\i}tre and Robertson \\citep[and ][in the pre-relativistic era]{Schw00}\\footnote{English translation: \\protect\\citet{Schw98}.}. Apart from providing a competitor to the infinite flat model for understanding WMAP observational data, cosmic topology can also help to improve physical insight into fundamentals of FLRW models that initially appear to be unrelated to global topology. \\postrefereechanges{ If the parallel-transported four-velocity Doppler interpretation of galaxy redshifts depends only on the metric (including $a(t)$), then it} leads to the physical contradiction of an inertial observer in a non-expanding Minkowski space-time receding from herself. This contradiction can be resolved, but only at the cost of introducing the notion of expanding space, in which case the motivation for a Doppler interpretation is weakened, or by explicitly stating that the comoving distance to the galaxy must be less than twice the injectivity radius of comovinig space. The link between the Doppler interpretation of redshifts and cosmic topology can be summarised physically as follows. \\begin{list}{(\\roman{enumi})}{\\usecounter{enumi}} \\item The relativistic Doppler interpretation of a cosmological redshift is obtained by parallel transporting the emitting galaxy's four-velocity along a photon path to the observer. \\item \\postrefereechanges{For an FLRW model of any of the three curvatures (with the restrictions as detailed above), the comoving spatial geodesic of the photon's path can be considered to be a closed path by changing from a simply-connected FLRW model to a multiply-connected FLRW model, without any modification of the metric. This should not affect the Doppler interpretation if that interpretation depends only on the metric.} \\item Closed comoving spatial paths constitute the mathematical foundation of cosmic \\postrefereechanges{topology \\citep[e.g., \\SSS{}3.4,][]{LaLu95}}. \\end{list} Hence, it is unsurprising that insistence on the possibility of interpreting galaxy redshifts as a relativistic Doppler effect leads to the properties of cosmic topology, and in turn requires the physical concept of expanding space, i.e. epochs \\postrefereechanges{when $\\dot{a}>0$} (Defn~\\ref{d-eos})." }, "0911/0911.3346_arXiv.txt": { "abstract": "We construct a new Hartree-Fock-Bogoliubov (HFB) mass model, labeled HFB-18, with a generalized Skyrme force. The additional terms that we have introduced into the force are density-dependent generalizations of the usual $t_1$ and $t_2$ terms, and are chosen in such a way as to avoid the high-density ferromagnetic instability of neutron stars that is a general feature of conventional Skyrme forces, and in particular of the Skyrme forces underlying all the HFB mass models that we have developed in the past. The remaining parameters of the model are then fitted to essentially all the available mass data, an rms deviation $\\sigma$ of 0.585 MeV being obtained. The new model thus gives almost as good a mass fit as our best-fit model HFB-17 ($\\sigma$ = 0.581 MeV), and has the advantage of avoiding the ferromagnetic collapse of neutron stars. ", "introduction": "Astrophysical considerations require that one have available nuclear-mass models as rigorously based as possible. In this way one might hope to be able to extrapolate from the mass data, which cluster fairly closely to the stability line, out towards the neutron drip line, and make reliable estimates of the masses of nuclei so neutron rich that there is no hope of measuring them in the foreseeable future; such nuclei are found in the outer crusts of neutron stars, and also play a vital role in the r-process of nucleosynthesis. To this end we have developed a series of nuclear-mass models based on the Hartree-Fock-Bogoliubov (HFB) method with Skyrme and contact-pairing forces, together with phenomenological Wigner terms and correction terms for the spurious collective energy; all the model parameters are fitted to essentially all the experimental mass data (see Ref.~\\cite{gcp09} and references quoted therein). To make the extrapolations to neutron-rich nuclei as reliable as possible, the underlying Skyrme force in model HFB-9 \\cite{sg05} and all later models was constrained to fit the zero-temperature equation of state (EoS) of neutron matter (NeuM), as calculated by Friedman and Pandharipande~\\cite{fp81} (FP) for realistic two- and three-nucleon forces. (We have so far been unable to obtain mass fits as good as our published ones when constraining to the more complete, and slightly stiffer, realistic neutron-matter EoS labeled A18 + $\\delta\\,v$ + UIX$^*$~\\cite{apr98}. Our findings are consistent with a recent analysis of the $\\pi^-/\\pi^+$ ratio in central heavy-ion collisions indicating that this EoS is too stiff~\\cite{xiao09}.) Because of the neutron-matter constraint our models can be used to extrapolate beyond the neutron drip line and calculate with some confidence the EoS of the inner crust of neutron stars~\\cite{onsi08} (throughout this paper we assume zero temperature). Since the good agreement of our forces with the FP calculation \\cite{fp81} of neutron matter extends to the highest density encountered in neutron stars it might be thought that our inner-crust EoS is continuous with that of the homogeneous core, the transition between the two regions taking place at around $0.5\\rho_0$, where $\\rho_0$ ($\\simeq$ 0.16 nucleons.fm$^{-3}$) is the equilibrium density of symmetric nuclear matter (SNM). However, in fitting our forces to the neutron matter of FP \\cite{fp81}, we {\\it assume} that our ground state is spin unpolarized, but in fact the underlying Skyrme forces of all our models, like all conventional Skyrme forces of the form (\\ref{1}), predict that beyond a certain supernuclear density the ground state of NeuM becomes ferromagnetic, i.e., at least partially polarized~\\cite{vida84,kw94,mar02}. In the case of the Skyrme force BSk17, the force underlying our best-fit model, HFB-17~\\cite{gcp09}, complete polarization sets in at a density of $\\rho_{frmg} = 1.24\\,\\rho_0$. On the other hand, microscopic calculations using different realistic forces and different methods~\\cite{fan01, bomb06, sam07, bord08,vid02} all predict no such polarization, at least up to about $5\\rho_0$. Moreover, in the case of all our own previously published HFB forces the predicted ferromagnetic state is unstable against collapse, the energy becoming more and more negative as the density increases (see, for example, the lowest curve in Fig.~\\ref{fig1}, constructed for force BSk17). This predicted ferromagnetic collapse sets in at densities that are certainly encountered in the cores of all neutron stars, and is contradicted by the very existence of neutron stars. Actually, the core of neutron stars does not consist of pure NeuM but rather of so-called neutron-star matter (N*M), which just below the crust is composed of neutrons with a weak admixture of proton-electron pairs in beta equilibrium (for simplicity we will assume that these are the only particles present also at higher densities). But if NeuM were indeed unstable against collapse then N*M would be likewise, since, at the very least, it would always be energetically advantageous for N*M to spontaneously transform into NeuM through electron capture beyond some critical density. Thus the stability of NeuM against collapse is a necessary condition for the existence of neutron stars. But even if no such instability is implied, there is a general tendency for Skyrme forces of the conventional form (\\ref{1}) to lead to the ground state of N*M being polarized. However, calculations based on realistic forces show the ground state of N*M, like that of NeuM, to be unpolarized at all densities~\\cite{bord08b}. Our purpose here is to show that by adding suitable terms to the Skyrme force it is possible to eliminate the anomalous prediction of a ferromagnetic transition in neutron stars, with essentially no impact on the high-quality fits to the mass data that we have previously obtained with conventional Skyrme forces. (An alternative approach to this problem has been followed by Margueron {\\it et al.} \\cite{marg09a, marg09}.) In Section II we discuss in more detail the nature of the spurious transition to a ferromagnetic state in NeuM and N*M associated with conventional Skyrme forces of the form (\\ref{1}), considering not only our own BSk17 \\cite{gcp09}, but also the widely used SLy4 \\cite{cha98}, which was specifically constructed for neutron-star calculations. Section III shows how extra terms in the Skyrme force can stop this spurious transition, while in Section IV we describe the new mass fit. Our conclusions are summarized in Section V, and in the Appendix we present the full formalism for the generalized form of Skyrme force used here. ", "conclusions": "We have extended our earlier Skyrme-HFB mass models by the inclusion of terms that are density-dependent generalizations of the usual $t_1$ and $t_2$ terms. We have shown that these new terms can be chosen in such a way as to prevent the high-density ferromagnetic collapse of neutron stars that was a general feature of our previous HFB mass models, without compromising the excellent fit to the mass data that we obtained in the past, and without relaxing any of the previously imposed constraints of conformity to reality. The mass predictions made by the new model, HFB-18, are, in fact, very similar to those made by the preceding model, HFB-17~\\cite{gcp09}. These two models not only give better fits to the mass data than does any other published model except that of Duflo and Zucker~\\cite{dz95}, but they are also by far the most microscopically founded models, and in particular their underlying interactions (BSk17 and BSk18, respectively) have been fitted to realistic calculations of both the EoS and the $^1S_0$ gap of neutron matter. They can thus be expected to make more reliable predictions of the highly neutron-rich nuclei that appear in the outer crust of neutron stars and that are involved in the r-process. Moreover, these mass models can be used to extrapolate beyond the drip line to the inner crust of neutron stars, using the respective interactions to calculate the EoS in this region. Our confidence in this extrapolation derives not only from the fit of the interactions to neutron matter but also from the precision fit to masses, which means that the presence of protons and the existence of inhomogeneities in the inner crust are well represented. Finally, pursuing our aim of a unified treatment of the different regions of a neutron star, we can use these effective interactions to calculate the EoS of the homogeneous core (at least in its outer parts where no complications from the possible appearance of hyperons and other particles arise). Of course, for pure NeuM nothing new can be obtained in this way beyond what has already been given by the realistic calculations to which our effective interactions have been fitted, but these interactions can then be reliably used to calculate N*M, which is not treated in the realistic calculations of FP~\\cite{fp81}. Another important application of these effective interactions, which would hardly be practical with realistic forces, is to make a detailed study of the transition between the inner crust and the fluid core. We have seen that for N*M, as for masses, the two interactions BSk17 and BSk18 give very similar results, provided we assume that the ground state for the former is unpolarized, which in fact is not the case. Here lies the basic difference between the two interactions: with BSk18 the ground state of NeuM and N*M is indeed unpolarized at all densities prevailing in neutron stars. This elimination of the spurious polarization is the essential contribution of this paper, made possible by the introduction of the terms in $t_4$ and $t_5$. However, we have so far made only a partial search in the new, extended parameter space, and there remains the possibility not only of improving the fit to the mass data still further but also of imposing further realistic constraints on the Skyrme force. \\appendix" }, "0911/0911.4085_arXiv.txt": { "abstract": "Coronal emission line intensities are commonly used to measure electron temperatures using emission measure and/or line ratio methods. In the presence of systematic errors in atomic excitation calculations and data noise, the information on underlying temperature distributions is fundamentally limited. Increasing the number of emission lines used does not necessarily improve the ability to discriminate between different kinds of temperature distributions. ", "introduction": "Spectroscopic measurements of the temperature of coronal plasma have been made for decades \\citep[e.g.][]{Seaton1962c,Noci2003}. \\citet{Feldman+others1999,Feldman+Landi2008} and \\citet{Landi+Feldman2008} presented intriguing evidence that electron temperatures of solar coronal plasma, measured from emission line spectra obtained high in the coroma, are clumped into several peaks, and are not broadly distributed. Narrow distributions of plasma temperature have important implications for the energy balance of the corona. In the methods used, frequency integrated line intensities $I$ are assumed to be simple functions $G(T)$ of the logarithm of the electron temperature, $T$, because of the dominance of two body collisional processes which lead to the well-known ``coronal approximation'' \\citep[e.g.][]{Woolley+Allen1948,Seaton1964a}. Two approaches were used. In one they solved for the differential emission measure $\\xi(T)$ which is an optimal solution to the inverse problem: \\begin{equation} I_i = \\int G_i(T) \\xi(T) dT,\\ \\ \\ \\ i=1\\ldots n \\end{equation} where there is a set of $n$ different emission lines. The other method sought the single logarithmic temperature $T_0$ such that $\\xi(T) \\propto \\delta(T-T_0)$, by plotting $I_i/G_i(T)$ and identifying intersection points for the observed lines. These approaches are, in fact, formally equivalent \\citep{McIntosh+Brown+Judge1997}. In both cases, information must be added to obtain both $\\xi(T)$ and $T_0$: in the first case the problem has to be ``regularized'' as it is ill-posed \\citep[e.g.][]{Craig+Brown1986}, one searches for example for the ``least structured'' solution $\\xi(T)$ that is compatible with the data. In the second case it is assumed that the plasma is indeed approximately isothermal. In the analysis of \\citet{Feldman+Landi2008}, the individual peaks in the $\\xi(T)$ functions have widths of 0.1-0.2 in $T$, and the curves of $I_i/G_i(T)$ intersect one another within similar margins. Some general questions arise. Given typical uncertainties in observable and model parameters, are the observed data compatible with different $\\xi(T)$ distributions? What, then, is the accuracy of the derived temperatures? In this paper these questions are addressed by asking, how broad can $\\xi(T)$ functions be to be incompatible with observed data. By how much can one change the temperature of an isothermal plasma before the differences in line intensities become significant? In the problem at hand there are unavoidably large and systematic uncertainties in the $G(T)$ functions. These uncertainties limit the information which can be extracted from emission line spectra. It is found that acceptable widths $\\width$ of the $\\xi(T)$ functions exceed the precision by which the lines can in principle determine that two different plasmas have slightly different temperatures. Thus, these widths set the lower limit to the ability of emission line techniques to diagnose electron temperatures. ", "conclusions": "The largest values of $\\width$ compatible with observations yield the accuracy with which emission lines can measure electron temperatures. The widths of the distributions compatible with known sources of uncertainties set a natural limit on the sharpness of a detectable peak in the underlying emission measure distributions. Typically the full width, $2\\width$, is 0.2 to 0.3 in the logarithmic electron temperature. The precision is an order of magnitude better, being estimated using the sensitivity of the spectra to two isothermal plasmas differing in logarithmic temperature by $\\tau$. Although this implies that two strictly isothermal plasmas {\\em could} be differentiated by using line ratios to a precision of order $\\epsilon_i \\widi/2 \\sim 0.015$ in the logarithmic electron temperature, we will perhaps never know if such isothermal plasmas exist with a width narrower than $\\width$, based upon these data. These properties prompt the following comments: \\begin{enumerate} \\item Error bars quoted below $\\sim0.1$ in logarithmic temperatures are not credible using these techniques. Landi and Feldman quote error bars of 0.04 and 0.05 in their work which seeks to provide spectroscopic evidence that the coronal temperature is ``quantized''. \\item The widths (FWHM) of peaks in the $\\xi(T)$ functions found by \\citet[][their figure 5]{Landi+Feldman2008} are between 0.1 and 0.2. These widths are characteristic of the limit with which emission lines can determine the isothermality of the emitting plasma, and are therefore determined more by regularization than by the data themselves. Their analysis is consistent with isothermal plasma, but is limited by the uncertainties discussed here. \\item Coronal temperatures are controlled by (unknown) heating mechanisms and cooling by heat conduction, flows, radiation. The scalings of \\citet{Rosner+Tucker+Vaiana1978}, based on simple energy balance considerations, show that a given coronal electron temperature $T_e$ in a loop of length $L$ requires an energy flux density $$ {\\cal F} \\propto (T_eL)^{7/2} $$ to sustain it agains conductive and radiation losses. Therefore, for a given $L$, if $T_e$ is uncertain to $\\pm0.15$ in its logarithm, the absolute value of log$_{10}{\\cal F}$ can be determined from a given length loop to an accuracy of $\\sim \\pm0.5$. The energy flux is constrained only to somewhere within a range spanning a factor of ten using emission lines. Conversely, variations of the energy flux ${\\cal F}$ dissipated in coronal loops by $\\lta 0.5$ in the logarithm cannot be distinguished through temperature dependent emission line methods. \\item {\\em Relative} changes in ${\\cal F}$ could, however, be detected to within about $\\pm 12\\%$, but again this precision depends on the plasma being nearly isothermal (i.e. actual $\\width \\sim \\delt$), which is neither known from the data nor expected from first principles. \\end{enumerate} The analytical results offer some hope that carefully selected sets of emission lines can be used to reduce the widths of distributions compatible with the data. Specifically, using lines within the same element removes errors arising from incorrect abundances, and can include lines with large values of $|x_i|$. In this case just a few lines with different values of $x_i$ can reduce $\\width$ via equation~(\\pref{eqn:bigx})). Using lines from within the same ion stage yields significantly smaller values of $\\epsilon_i \\sim 0.1$, but while the precision improves linearly with $\\epsilon_i$ the accuracy improves only quadratically." }, "0911/0911.3170_arXiv.txt": { "abstract": "\\begin{list}{ } {\\rightmargin 1in} \\baselineskip = 11pt \\parindent=1pc {\\small We survey the basic principles of atmospheric dynamics relevant to explaining existing and future observations of exoplanets, both gas giant and terrestrial. Given the paucity of data on exoplanet atmospheres, our approach is to emphasize fundamental principles and insights gained from Solar-System studies that are likely to be generalizable to exoplanets. We begin by presenting the hierarchy of basic equations used in atmospheric dynamics, including the Navier-Stokes, primitive, shallow-water, and two-dimensional nondivergent models. We then survey key concepts in atmospheric dynamics, including the importance of planetary rotation, the concept of balance, and simple scaling arguments to show how turbulent interactions generally produce large-scale east-west banding on rotating planets. We next turn to issues specific to giant planets, including their expected interior and atmospheric thermal structures, the implications for their wind patterns, and mechanisms to pump their east-west jets. Hot Jupiter atmospheric dynamics are given particular attention, as these close-in planets have been the subject of most of the concrete developments in the study of exoplanetary atmospheres. We then turn to the basic elements of circulation on terrestrial planets as inferred from Solar-System studies, including Hadley cells, jet streams, processes that govern the large-scale horizontal temperature contrasts, and climate, and we discuss how these insights may apply to terrestrial exoplanets. Although exoplanets surely possess a greater diversity of circulation regimes than seen on the planets in our Solar System, our guiding philosophy is that the multi-decade study of Solar-System planets reviewed here provides a foundation upon which our understanding of more exotic exoplanetary meteorology must build. \\\\~\\\\~\\\\~}% \\end{list} ", "introduction": "\\label{Intro} The study of atmospheric circulation and climate began hundreds of years ago with attempts to understand the processes that determine the distribution of surface winds on the Earth \\citep[e.g.,][]{hadley-1735}. As theories of Earth's general circulation became more sophisticated \\citep[e.g.,][]{lorenz-1967}, the characterization of Mars, Venus, Jupiter, and other Solar-System planets by spacecraft starting in the 1960s demonstrated that the climate and circulation of other atmospheres differ, sometimes radically, from that of Earth. Exoplanets, occupying a far greater range of physical and orbital characteristics than planets in our Solar System, likewise plausibly span an even greater diversity of circulation and climate regimes. This diversity provides a motivation for extending the theory of atmospheric circulation beyond our terrestrial experience. Despite continuing questions, our understanding of the circulation of the modern Earth atmosphere is now well developed \\citep[see, e.g.,][]{held-2000, schneider-2006, vallis-2006}, but attempts to unravel the atmospheric dynamics of Venus, Jupiter, and other Solar-System planets remain ongoing, and the study of atmospheric circulation of exoplanets is in its infancy. For exoplanets, driving questions fall into several overlapping categories. First, we wish to understand and explain new observations constraining atmospheric structure, such as light curves, photometry, and spectra obtained with the {\\it Spitzer, } {\\it Hubble}, or {\\it James Webb Space Telescopes (JWST)}, thus helping to characterize specific exoplanets as remote worlds. Second, we wish to extend the theory of atmospheric circulation to the wide range of planetary parameters encompassed by exoplanets. Existing theory was primarily developed for conditions relevant to Earth, and our understanding of how atmospheric circulation depends on atmospheric mass, composition, stellar flux, planetary rotation rate, orbital eccentricity, and other parameters remains rudimentary. Significant progress is possible with theoretical, numerical, and laboratory investigations that span a wider range of planetary parameters. Third, we wish to understand the conditions under which planets are habitable, and answering this question requires addressing the intertwined issues of atmospheric circulation and climate. \\begin{figure*} \\epsscale{1.0} \\plotone{coupled-dynamics-radiation-crop2.jpg} \\caption{\\small Atmospheric circulation results from a coupled interaction between radiation and hydrodynamics: horizontal temperature and pressure contrasts generate winds, which drive the atmosphere away from local radiative equilibrium. This in turn allows the spatially variable thermodynamic (radiative) heating and cooling that maintains the horizontal temperature and pressure contrasts.} \\label{flow-chart} \\end{figure*} What drives atmospheric circulation? Horizontal temperature contrasts imply the existence of horizontal pressure contrasts, which drive winds. The winds in turn push the atmosphere away from radiative equilibrium by transporting heat from hot regions to cold regions (e.g., from the equator to the poles on Earth). This deviation from radiative equilibrium allows net radiative heating and cooling to occur, thus helping to maintain the horizontal temperature and pressure contrasts that drive the winds (see Fig.~\\ref{flow-chart}). Spatial contrasts in thermodynamic heating/cooling thus fundamentally drive the circulation, yet it is the existence of the circulation that allows these heating/cooling patterns to exist. (In the absence of a circulation, the atmosphere would relax into a radiative-equilibrium state with a net heating rate of zero.) The atmospheric circulation is thus a coupled radiation-hydrodynamics problem. On the Earth, for example (see Fig.~\\ref{earth-radiation-balance}), the equator and poles are not in radiative equilibrium. The equator is subject to net heating, the poles to net cooling, and it is the mean latitudinal heat transport that is both responsible for and driven by these net imbalances. \\begin{figure*} \\epsscale{1.0} \\plotone{earth-radiation-balance.jpg} \\caption{\\small Earth's energy balance. The Earth absorbs more sunlight at the equator than the poles (blue curve, denoted ``shortwave''). The Earth also radiates more infrared energy to space at the equator than the poles (red curve, denoted ``longwave''). However, because the atmospheric/oceanic circulation act to mute the latitudinal temperature contrasts (relative to radiative equilibrium), the longwave radiation exhibits less latitudinal variation than the shortwave absorption. Thus, the circulation leads to net heating at the equator and net cooling at the poles, which in turn drives the circulation. Data are an annual-average for 1987 obtained from the NASA Earth Radiation Budget Experiment (ERBE) project. Copyright M. Pidwirny, www.physicalgeography.net, used with permission.} \\label{earth-radiation-balance} \\end{figure*} The mean climate (e.g., the global-mean surface temperature of a planet) depends foremost on the absorbed stellar flux and the atmosphere's need to reradiate that energy to space. Yet even the global-mean climate is strongly affected by the atmospheric mass, composition, and circulation. On a terrestrial planet, for example, the circulation helps to control the distribution of clouds and surface ice, which in turn determine the planetary albedo and the mean surface temperature. In some cases, a planetary climate can have multiple equilibria (e.g., a warm, ice-free state or a cold, ice-covered ``snowball Earth'' state), and in such cases the circulation plays an important role in determining the relative stability of these equilibria. Understanding the atmosphere/climate system is challenging because of its nonlinearity, which involves multiple positive and negative feedbacks between radiation, clouds, dynamics, surface processes, planetary interior, and life (if any). The inherent nonlinearity of fluid motion further implies that even atmospheric-circulation models neglecting the radiative, cloud, and surface/interior components can exhibit a large variety of behaviors. From the perspective of studying the atmospheric circulation, transiting exoplanets are particularly intriguing because they allow constraints on key planetary attributes that are a prerequisite to characterizing an atmosphere's circulation regime. When combined with Doppler velocity data, transit observations permit a direct measurement of the exoplanet's radius, mass and thus surface gravity\\footnote{Combining Doppler velocity and transit measurements lifts the mass-inclination degeneracy.}. With the additional expectation that close-in exoplanets are tidally locked if on a circular orbit, or pseudo-synchronized\\footnote{Pseudo-synchronization refers to a state of tidal synchronization achieved only at periastron passage (=closest approach), as expected from the strong dependence of tides with orbital separation.} if on an eccentric orbit, the planetary rotation rate is thus indirectly known as well. Knowledge of the radius, surface gravity, rotation rate and external irradiation conditions for several exoplanets, together with the availability of direct observational constraints on their emission, absorption and reflection properties, opens the way for the development of comparative atmospheric science beyond the reach of our own Solar System. The need to interpret these astronomical data reliably, by accounting for the effects of atmospheric circulation and understanding its consequences for the resulting planetary emission, absorption and reflection properties, is the central theme of this chapter. Tidally locked close-in exoplanets, for example, are subject to an unusual situation of permanent day/night radiative forcing, which does not exist in our Solar System\\footnote{Venus may provide a partial analogy, which has not been fully exploited yet.}. To address the new regimes of forcings and responses of these exoplanetary atmospheres, a discussion of fundamental principles of atmospheric fluid dynamics and how they are implemented in multi-dimensional, coupled radiation-hydrodynamics numerical models of the GCM (General Circulation Model) type is required. Contemplating the wide diversity of exoplanets raises a number of fundamental questions. What determines the mean wind speeds, direction, and 3D flow geometry in atmospheres? What controls the equator-to-pole and day-night temperature differences? What controls the frequencies and spatial scales of temporal variability? What role does the circulation play in controlling the mean climate (e.g., global-mean surface temperature, composition) of an atmosphere? How do these answers depend on parameters such as the planetary rotation rate, gravity, atmospheric mass and composition, and stellar flux? And, finally, what are the implications for observations and habitability of exoplanets? At present, only partial answers to these questions exist \\citep[see reviews by][]{showman-etal-2008b, cho-2008}. With upcoming observations of exoplanets, constraints from Solar-System atmospheres, and careful theoretical work, significant progress is possible over the next decade. While a rich variety of atmospheric flow behaviors is realized in the Solar System alone---and an even wider diversity is possible on exoplanets---the fundamental physical principles obeyed by all planetary atmospheres are nonetheless universal. With this unifying notion in mind, this chapter provides a basic description of atmospheric circulation principles developed on the basis of extensive Solar-System studies and discusses the prospects for using these principles to better understand physical conditions in the atmospheres of remote worlds. The plan of this chapter is as follows. In Section~\\ref{equations}, we introduce several of the equation sets that are used to investigate atmospheric circulation at varying levels of complexity. This is followed (Section~\\ref{basic-concepts}) by a tutorial on basic ideas in atmospheric dynamics, including atmospheric energetics, timescale arguments, force balances relevant to the large-scale circulation, the important role of rotation in generating east-west banding, and the role of waves and eddies in shaping the circulation. In Section~\\ref{giants}, we survey the atmospheric dynamics of giant planets, beginning with generic arguments to constrain the thermal and dynamical structure and proceeding to specific models for understanding the circulation of our ``local'' giant planets (Jupiter, Saturn, Uranus, Neptune) as well as hot Jupiters and hot Neptunes.\\footnote{The terms hot Jupiter and hot Neptune refer to giant exoplanets with masses comparable to those of Jupiter and Neptune, respectively, with orbital semi-major axes less than $\\sim0.1\\,$AU, leading to high temperatures.} In Section~\\ref{terrestrial}, we turn to the climate and circulation of terrestrial exoplanets. Observational constraints in this area do not yet exist, and so our goal is simply to summarize basic concepts that we expect to become relevant as this field expands over the next decade. This includes a description of climate feedbacks (Section~\\ref{climate}), global circulation regimes (Section~\\ref{regimes}), Hadley-cell dynamics (Section~\\ref{hadley}), the dynamics of the so-called midlatitude ``baroclinic'' zones where baroclinic instabilities dominate (Section~\\ref{baroclinic-zone}), the slowly rotating regime relevant to Venus and Titan (Section~\\ref{slowly-rotating}), and finally a survey of how the circulation responds to the unusual forcing associated with synchronous rotation, extreme obliquities, or extreme orbital eccentricities (Section~\\ref{unusual-forcing}). The latter topics, while perhaps the most relevant, are the least understood theoretically. In Section~\\ref{highlights} we summarize recent highlights, both observational and theoretical, and in Section~\\ref{future} we finish with a survey of future prospects. \\bigskip ", "conclusions": "" }, "0911/0911.2495_arXiv.txt": { "abstract": "We give the evolution and constraint equations on an asymmetrically embedded brane in the form of average and difference equations. ", "introduction": "In a recent paper \\cite{3+1+1} we have developed the covariant gravitational dynamics in a 3+1+1 dimensional space-time in the spirit of the general relativistic 3+1 covariant cosmology \\cite{3+1}, generalizing both previous approaches to 5-dimensional (5d) gravitational dynamics \\cite{brane-dynamics}% , brane 3+1 covariant cosmology \\cite{Maartens-eqs} and 2+1+1 covariant dynamics \\cite{2+1+1}. Such a formalism may turn useful in discussing perturbations on the brane \\cite{branepert}. The singled-out directions are the normal $n^{a}$ to the hypersurface representing the time-evolving 3-dimensional space (the brane with metric $h_{ab}$ and tension $\\lambda $) and a temporal direction $u^{a}$ tangent to this hypersurface. All employed gravitational variables are projected to the brane. They consist of kinematical variables $(\\Theta ,\\sigma _{ab},\\omega _{ab},A_{a},K_{a},L_{a})$ related to the vector $u^{a}$; analogous quantities, carrying an overhat, related to $n^{a}$; two kinematical scalars $(K,\\hat{K})$ related to both; finally gravito-electro-magnetic quantities $(\\mathcal{E},\\mathcal{H}_{k},% \\mathcal{F}_{kl},\\mathcal{E}_{k},\\widehat{\\mathcal{E}}_{k},\\mathcal{E}_{kl},% \\widehat{\\mathcal{E}}_{kl},\\mathcal{H}_{kl},\\widehat{\\mathcal{H}}_{kl})$. For details on the definitions of these quantities see Ref. \\cite{3+1+1}. The matter on the brane has been decomposed as $T_{ab}=\\rho u_{a}u_{b}\\!+\\!q_{(a}u_{b)}\\!+\\!ph_{ab}\\!+\\!\\pi _{ab},$ while possible non-standard model fields nesting in the 5d space-\\negthinspace time as $% \\widetilde{T}_{ab}=\\widetilde{\\rho }u_{a}u_{b}\\!+\\!2\\widetilde{q}% _{(a}u_{b)}\\!+\\!2\\widetilde{q}u_{(a}n_{b)}\\!+\\!\\widetilde{p}h_{ab}\\!+\\!% \\widetilde{\\pi }n_{a}n_{b}\\!+\\!2\\widetilde{\\pi }_{(a}n_{b)}\\!+\\!\\widetilde{% \\pi }_{ab}.$ The gravitational coupling constants $\\widetilde\\kappa^2$ and $% \\kappa^2=\\widetilde\\kappa^4\\lambda/6$ act in 5d and on the brane, respectively. The generic form of the evolution and constraint equations in the 5-dimensional spacetime were given as Appendix C of Ref. \\cite{3+1+1} and they were specified on a $Z_{2}$-symmetrically embedded brane in Subsection IV.D. The embedding however is not necessarely symmetric: an asymmetric embedding is known to generate late time acceleration in a cosmological setup \\cite{Induced}. Therefore here we generalize the formalism to an asymmetrically embedded brane. Following the recipe presented in Subsection IV.D of Ref. \\cite{3+1+1}, we give here the evolution and constraint equations obtained both as averages and differences across the brane. We denote the average over the two sides of the brane of any quantity by angle brackets and the jump by $\\Delta $. For the extrinsic curvature components $% \\mathcal{K}\\equiv (\\widehat{\\Theta },\\widehat{\\sigma }_{ab},\\widehat{K},% \\widehat{K}_{a})$ the latter is directly related to the brane matter variables and brane tension cf. the Israel-Lanczos condition, Eqs. (76)-(79) of Ref. \\cite{3+1+1}. Angle brackets on indices indicate projection to the 3-space, symmetrization and trace-free character. ", "conclusions": "Both the average and the difference equations reduce to the corresponding equations given in Subsection IV.D of Ref \\cite{3+1+1}, by taking into account that in the particular case of a symmetric embedding for quantities defined with an odd (even) number of $n^{a}$, the conditions $\\Delta f=2f,\\langle f\\rangle =0$ ($\\Delta f=0,\\langle f\\rangle =f$) hold. In particular, the extrinsic curvature components $\\mathcal{K}$ belong to the first group. \\textit{Acknowledgements}: This work was supported by the Pol\\'{a}nyi and Sun Programs of the Hungarian National Office for Research and Technology (NKTH), and by the Hungarian Scientific Research Fund (OTKA) grant 69036. We acknowledge financial support from the organizers of the Grassmannian Conference in Fundamental Cosmology (Grasscosmofun'09)." }, "0911/0911.0035_arXiv.txt": { "abstract": "We use {\\it Spitzer} data to infer that the small infrared excess of V819 Tau, a weak-lined T Tauri star in Taurus, is real and not attributable to a ``companion'' 10\\arcsec\\ to the south. We do not confirm the mid-infrared excess in HBC 427 and V410 X-ray 3, which are also non-accreting T Tauri stars in the same region; instead, for the former object, the excess arises from a red companion 9\\arcsec\\ to the east. A single-temperature blackbody fit to the continuum excess of V819 Tau implies a dust temperature of 143 K; however, a better fit is achieved when the weak 10 and 20 $\\mu$m silicate emission features are also included. We infer a disk of sub-$\\mu$m silicate grains between about 1 AU and several 100 AU with a constant surface density distribution. The mid-infrared excess of V819 Tau can be successfully modeled with dust composed mostly of small amorphous olivine grains at a temperature of 85 K, and most of the excess emission is optically thin. The disk could still be primordial, but gas-poor and therefore short-lived, or already at the debris disk stage, which would make it one of the youngest debris disk systems known. ", "introduction": "The evolution and dissipation of protoplanetary disks has been an active area of research for the last few decades; especially the advent of sensitive near- and mid-infrared observations has allowed us to explore inner disk regions (out to a few AU), encompassing the radii where planets are thought to form. The role of planet formation in disk dissipation has been all but proven; while disks likely evolve from a flared, optically thick configuration to a flat, settled disk via grain growth and settling, it is not clear to what extent the eventual dissipation of the remaining dust and gas and the formation of planets are linked. A key evolutionary stage in disk dissipation is the transition from the classical T Tauri stage, when a pre-main-sequence star is surrounded by an accreting disk, to the weak-lined T Tauri phase, when accretion ends and the disk disappears. This transitional stage is believed to be short ($\\lesssim$ 10$^5$ years), since only few stars with vanishing infrared excesses have been observed \\citep[e.g.,][]{skrutskie90,simon95}. The Infrared Spectrograph\\footnote{The IRS was a collaborative venture between Cornell University and Ball Aerospace Corporation funded by NASA through the Jet Propulsion Laboratory and the Ames Research Center.} \\citep[IRS;][]{houck04} on board the {\\it Spitzer Space Telescope} \\citep{werner04} has provided new insights into the study of transitional disks. For example, in the nearby and well-studied Taurus star-forming region \\citep{kenyon95}, out of a total of 111 T Tauri stars, five objects were identified whose inner disk regions are cleared to different degrees, but outer, optically thick disks remain, delimited by a well-defined inner disk rim \\citep{dalessio05,calvet05,furlan06,espaillat07}. Planet formation, photoevaporation, or inner disk draining induced by the magneto-rotational instability may play a role in removing the inner disk \\citep{clarke01,quillen04,alexander07,chiang07}, but also close binary companions can clear inner regions due to orbital resonances \\citep{artymowicz94,ireland08}. V819 Tau is a weak-lined T Tauri star with a spectral type of K7 in the Taurus star-forming region \\citep{herbig88}. We presented its {\\it Spitzer} IRS spectrum in \\citet{furlan06} as part of the sample of Class III objects in Taurus, noting that it showed an infrared excess beyond about 12 $\\mu$m (see Figure \\ref{V819Tau_IRS}). However, we could not confirm whether this excess was real due to the presence of a ``companion'' 10\\arcsec\\ to the south that is detected in 2MASS images \\citep{skrutskie06}. Two other weak-lined T Tauri stars in Taurus, HBC 427 and V410 X-ray 3, with spectral types of K5 and M6, respectively \\citep{steffen01,strom94}, were also introduced in \\citet{furlan06} as objects with uncertain infrared excesses. While HBC 427 has a ``companion'' star about 15\\arcsec\\ to the southeast that is seen in 2MASS images, V410 X-ray 3 appears to be single. In the meantime, we obtained IRS peak-up images of V819 Tau and HBC 427, and we re-reduced the IRS spectra of all three objects. We do not confirm the mid-infrared excess in HBC 427 and V410 X-ray 3, but, together with Multiband Imaging Photometer for {\\it Spitzer} \\citep[MIPS;][]{rieke04} 24 $\\mu$m images from the {\\it Spitzer} archive, we establish that the infrared excess is intrinsic to V819 Tau, and we apply simple models to derive the distribution of dust around this T Tauri star. \\begin{figure} \\centering \\includegraphics[angle=90, scale=0.35]{f1.eps} \\caption{The spectral energy distribution of V819 Tau; the optical photometry is from \\citet{kenyon95}, JHK$_s$ photometry from 2MASS \\citep{skrutskie06}, IRAC fluxes from \\citet{luhman06}, and the 450 $\\mu$m upper limit from \\citet{andrews05}. The IRS spectrum (also shown with error bars in the figure inset), IRS peak-up fluxes, and MIPS fluxes are from this work. The data were dereddened using Mathis's reddening law for $R_V$=3.1 \\citep{mathis90} and $A_V$=1.7 \\citep{furlan06}. The stellar photosphere is a scaled NextGen model with $T_{eff}$=4000 K and $log(g)$=3.5 (see text for details). \\label{V819Tau_IRS}} \\end{figure} \\newpage ", "conclusions": "Of the three Class III objects in Taurus that showed tentative mid-infrared excess emission in \\citet{furlan06}, we could confirm an infrared excess only in V819 Tau. While V819 Tau is a single star, HBC 427 and V410 X-ray 3 are close binaries: HBC 427 is a spectroscopic binary with a semi-major axis of about 0.03\\arcsec, comprised of a K5 and an M2 star \\citep{steffen01}; V410 X-ray 3 consists of an M6 and an M7.7 star separated by $\\sim$ 0.05\\arcsec\\ \\citep{kraus06}. Neither object shows signs of accretion \\citep{strom94, kenyon98, mohanty05}, and therefore both dust and gas have already dissipated in their inner disks. HBC 427 is also not detected at sub-millimeter wavelengths, implying an upper limit for the disk mass of 0.0007 \\Msun \\citep{andrews05}. It is likely that interaction between the binary and the disk in these systems is the cause for the absence of disk material \\citep{jensen94}. V819 Tau is also not accreting any more; its H${\\alpha}$ 10\\% width amounts to 166 km s$^{-1}$ \\citep{nguyen09}, which is considerably less than the lower limit of 270 km s$^{-1}$ that characterizes accretors \\citep{white03}. Also the low H${\\alpha}$ equivalent width of 1.7-3.2 {\\AA} \\citep{strom94, kenyon98} confirms its nature as a weak-lined T Tauri star. \\citet{white01} determined an upper limit of $1.4 \\times 10^{-9}$ {\\Msun} yr$^{-1}$ for the mass accretion rate based on the lack of $U$-band excess. No H$_2$ emission from warm inner disk regions was detected \\citep{bary03}, implying that both the gas and dust have been removed there. Our data indicate that there are no small, warm dust grains within approximately 1 AU from the star. One process that can generate cleared inner disk regions is photoevaporation \\citep{clarke01}; it sets in once the mass accretion rate has dropped to levels below a few 10$^{-10}$ {\\Msun} yr$^{-1}$, and it acts by eroding the disk in a photoevaporative flow beyond a certain radius, which, for V819 Tau, amounts to 7.1 AU. This value is larger than the inner radius of our best-fit optically thin disk models, but it is remarkably close to the value of 6.8 AU we estimated by assuming the excess emission to arise from an optically thick, blackbody surface. However, it seems that the dust emission is optically thin, given the presence of 10 and 20 $\\mu$m silicate emission features and the low infrared excess luminosity ($L_{IR}/L_{bol} \\sim $ 10$^{-3}$). Moreover, our models require small dust grains at distances of at least $\\sim$ 1 AU from the star. Thus, photoevaporation cannot explain the inner disk clearing of V819 Tau; a change in grain opacity caused by grain growth to sizes well beyond 1 $\\mu$m in this region, or the gravitational influence of another body, such as a planet, are possible explanations for the inner disk clearing. The question also arises whether the material we detect around V819 Tau is primordial or already second-generation dust. Detection of gas emission from the inner disk would support the primordial nature of the disk; so far, the absence of accretion signatures and H$_2$ emission from the inner disk suggest that the gas has already dissipated. Without the gas, dust particles would be subject to collisions, Poynting-Robertson (PR) drag, radiation pressure, and corpuscular stellar wind pressure \\citep[e.g.,][]{wyatt99, chen06}, which, in the region from $\\sim$ 1 to 100 AU from V819 Tau, would limit the lifetime of sub-$\\mu$m grains to $\\lesssim$ 10$^5$ years. Therefore, in the absence of gas, dust would have to be continuously replenished, since the age of V819 Tau is about 2 Myr. V819 Tau would be one of the youngest debris disk systems; its infrared excess would lie on the higher side of what is usually measured in debris disks \\citep[e.g.,][]{chen06}, but still within the observed range. Collisions among planetesimals, as are thought to occur in debris disks, typically produce larger grains ($\\gtrsim$ 10 $\\mu$m), since the mid-infrared spectra of the majority of debris disks are featureless \\citep{chen06,carpenter09}. The fact that small, amorphous silicate grains dominate the optically thin emission of V819 Tau, with just a minor possible presence of crystalline silicates, suggests that the dust we observe consists of mostly pristine material. Colliding or evaporating comet nuclei, which likely form in the outer disk regions and therefore incorporate more unprocessed silicates into ice, could be the source of this pristine dust. It is also possible that we are just observing ``true'' primordial, unprocessed dust from large disk radii; this would imply that we are witnessing a rare stage in the life of a protoplanetary disk, when the gas has already been mostly removed from the inner disk, and the remaining dust is dissipated on a timescale of 10$^5$ years or less. Then, not only could the disk structure of V819 Tau be described as transitional due to the lack of infrared excess below $\\sim$ 8 $\\mu$m and the clear presence of such an excess beyond 12 $\\mu$m, but also the evolutionary stage of this object could be considered transitional, i.e., having just transitioned from a primordial, optically thick disk to an optically thin one. Even though a few percent of the T Tauri population in Taurus can be considered as transitional (defined as disks with substantial inner disk clearings and thus low near-infrared excesses; \\citealt{furlan09}), V819 Tau would stand out by its very low infrared excess. It would be the only object in Taurus observed so far at an advanced transitional stage, when the disk material is already optically thin. Given the small amount of dust left in the disk, planet formation would have to be already completed in this disk; this conclusion would also apply if the dust were second-generation and therefore the result of dynamically perturbed, remnant planetary building blocks." }, "0911/0911.0729_arXiv.txt": { "abstract": "One of the main science goal of the future European Extremely Large Telescope will be to understand the mass assembly process in galaxies as a function of cosmic time. To this aim, a multi-object, AO-assisted integral field spectrograph will be required to map the physical and chemical properties of very distant galaxies. In this paper, we examine the ability of such an instrument to obtain spatially resolved spectroscopy of a large sample of massive ($0.1 \\leq M_{stellar} \\leq 5\\times10^{11}M_{\\odot}$) galaxies at $2\\leq z < 6$, selected from future large area optical-near IR surveys. We produced a set of about one thousand numerical simulations of 3D observations using reasonable assumptions about the site, telescope, and instrument, and about the physics of distant galaxies. These data-cubes were analysed as real data to produce realistic kinematic measurements of very distant galaxies. We then studied how sensible the scientific goals are to the observational (i.e., site-, telescope-, and instrument-related) and physical (i.e., galaxy-related) parameters. We specifically investigated the impact of AO performance on the science goal. We did not identify any breaking points with respect to the parameters (e.g., the telescope diameter), with the exception of the telescope thermal background, which strongly limits the performance in the highest (z$>$5) redshift bin. We find that a survey of $N_{gal}$ galaxies that fulfil the range of science goals can be achieved with a $\\sim$90 nights program on the E-ELT, provided a multiplex capability $M \\sim N_{gal}/8$. ", "introduction": "Over the last decade, the synergy of 8-10 meter class telescopes with HST has strongly re-invigorated the field of galaxy formation and evolution by unveiling very distant galaxies up to z$\\sim$6 (e.g., \\citealt{cuby03,rhoads03,bremer04,bouwens06}), by allowing the first determination of the global star formation history since redshift z$\\sim$6 (e.g., \\citealt{hopkins04}), and by providing the first insights on the stellar mass assembly history out to z$\\sim$5 (e.g., \\citealt{drory05,pozzetti07,marchesini07,perez07}). Despite these recent progresses, the outstanding question remains on how and when galaxies assembled their baryonic mass across cosmic time. The CDM standard model has provided a satisfactory scenario describing the hierarchical assembly of dark matter halos, in a bottom-up sequence which is now well-established over the whole mass structure spectrum. In contrast, little progress has been made in the physical understanding of the formation and evolution of the baryonic component because the conversion of baryons into stars is a complex, poorly understood process. As a result, all intellectual advances in galaxy formation and evolution over the last decade have been essentially empirical, often based on phenomenological (or semi-analytical) models, which heavily rely on observations to describe, with simplistic rules, such processes as star formation efficiency, energy feedback from star formation and AGN, chemical evolution, angular momentum transfer in merging, etc. Cornerstones observations in this empirical framework are the total and stellar mass of galaxies and their physical properties, including the age and metallicities of their underlying stellar populations, dust extinction, star formation rate, and structural/morphological parameters. The study of well-established scaling relations involving a number of these physical parameters (e.g. mass-metallicity, fundamental plane, colour-magnitude, morphology-density) are essential for understanding the physical processes driving galaxy evolution. However, with the current generation of 10m-class telescopes, we have been able to construct for example the fundamental plane of early-type galaxies, or to measure the Tully-Fisher relation of late-type galaxies over a wide range of masses only at low and intermediate redshifts (z$<$1), whereas only the brightest or most massive galaxies have been accessible at z $>$ 1, and a direct measurement of masses has been almost completely out of reach at z $>$ 2. Thus, our ability to explore the evolution and origin of the aforementioned scaling relations has rapidly reached the limit of 10m-class telescopes. Hence, most of the outstanding questions arisen from recent observational galaxy evolution studies which have pushed the 10m-class telescopes to their limits call for an Extremely Large Telescope (ELT), specifically to extend the spectroscopic limit by at least two magnitudes with near-diffraction limit angular resolution. IFU spectrographs on 8-10m class telescopes (e.g., VLT/SINFONI and Keck/Osiris) are now routinely deriving the spatially-resolved kinematics of massive (i.e., with stellar masses larger than 10$^{10}M_\\odot$), distant galaxies, from z$\\sim$0.4 to z$\\sim$3 \\citep{flores06,puech06,yang08,forster06,wright07,shapiro08,bournaud08,genzel08,vanstarkenburg08,law09,wright09,epinat09,forsterschreiber09,lemoine09}. Such studies have brought new and essential insights into galaxy evolution processes, such as the fraction of rotators as a function of redshift \\citep{yang08,forsterschreiber09}, a better understanding of the evolution of the Tully-Fisher relation \\citep{puech08,cresci09,puech09}, or the first glimpse into the evolution of the angular momentum \\citep{puech07,bouche07}. However, several uncertainties and limitations remain, especially at the highest redshifts (i.e., z$>$1). For instance, z$\\sim$2 surveys were drawn from different selection criteria (e.g., BX, BM, or BzK galaxies), and this might bias the resulting sample in a subtle and quite uncontrolled way. Besides, samples in the NIR are selected in optimal atmospheric windows, free of OH sky lines and maximising the atmospheric throughput, which intrinsically limits the redshift range and size of the resulting sample. The only way to overcome these issues is to move to telescopes with larger collecting areas. ESO is currently developing a Phase B study for a European 42-meter telescope \\citep{gilmozzi08,spyromilio08}. As part of the project development, the Design Reference Mission (DRM) aims at producing a set of observing proposals and corresponding simulated data which together provide the project with a tool to assess the extent to which the E-ELT addresses key scientific questions and assist in critical trade-off decisions\\footnote{see http://www.eso.org/sci/facilities/eelt/science/drm/}. In the frame of the DRM effort, simulations of 3D high-z galaxy observations were undertaken. These observations are expected to yield direct kinematics of stars and gas in the first generation of massive galaxies (in the range 0.1 $\\leq$ M $\\leq$ 5$\\times$10$^{11}$M$_\\odot$), as well as their stellar population properties. This will allow one to derive dynamical masses, ages, metallicities, star-formation rates, dust extinction maps, to investigate the presence of disk and spheroidal components and the importance of dynamical processes (e.g. merging, in/outflows) which govern galaxy evolution. These data will also allow one to study the onset of well known scaling relations at lower redshifts, and to witness the gradual shift of star formation from the most massive galaxies in the highest density regions to less massive galaxies in the field. To assess the science achievement of an NIR IFS on the E-ELT, as well as to better understand the interactions between the telescope, site, and instrument, we have produced an extensive set of $\\sim$1000 simulations of observations of very distant galaxies. It is important to realize that if one wants to study what kind of instrument is needed to understand the galaxy mass assembly process, then one needs to explore a huge parameter space, i.e., from relaxed smooth or clumpy rotating disks to major and minor mergers, taking care of spanning all possible mass ratios, viewing angles, merger geometry and so on, which would make such a direct approach very difficult, and probably even impossible, on the practical side. Moreover, one has to recognise that our knowledge and understanding of the physics of distant galaxies is still incomplete, which necessitates to extrapolate some of their characteristics. Therefore, the simulations presented in this paper do not intend to be exhaustive in any sense. Rather, we aim at exploring this very large parameter space using reasonable assumptions and guesses. Indeed, the DRM exercise consisted in a broad exploration of the parameter space in realistic observing conditions, in order to identify possible limitations and/or breaking points, as well as interactions between the telescope, site, and instrument, which could potentially impact the telescope design. This paper is the second of a series that explores the performances of a NIR IFS on the E-ELT. In the first paper, we presented our methodology to simulate realistic observations of distant galaxies \\citep{puech08b}. We also produced first simulations and explored performance using a few scientifically-motivated cases. They illustrated the concept of ``scale-coupling'', i.e., the relationship between the IFU pixel scale and the size of the kinematic features that need to be recovered by 3D spectroscopy in order to understand the nature of the galaxy and its substructure. In \\cite{puech08b}, we focused on the largest spatial scales, which are of particular interest because they carry most of the kinematic information useful to reveal the process underlying galaxy dynamics, i.e., whether a given galaxy is in a coherent and stable dynamical state (e.g., rotation), or, on the contrary, out of equilibrium (e.g., subsequently to a merger). In this paper, we present an updated version of the simulation pipeline. The main improvement is related to the inclusion of thermal emissions from both the IFS and the telescope. This allowed us to explore a wider parameter space, and especially areas where observations are limited by the thermal emission rather than by the sky background (e.g., relatively faint targets in the K band). We also significantly extended the range of morpho-kinematic templates and of physical parameters (e.g., radius, mass, and velocity as a function of redshift) considered in the simulations. This paper is organised as follows. In Sect. 2, we present our methodology and the pipeline used for simulations. In Sect. 3, we detail the scientific and observational inputs used in the simulations. In Sect. 4, we present the results of the simulations, which are discussed in Sect. 5. A conclusion is drawn in Sect 6. Throughout, we adopt $H_0=70$ km/s/Mpc, $\\Omega _M=0.3$, and $\\Omega _\\Lambda=0.7$, and the $AB$ magnitude system. ", "conclusions": "We have conducted simulations of the ``Physics and mass assembly of galaxies out of z$\\sim$6'' science case for the E-ELT, exploring a wide range of observational and physical parameters. We have defined figures of merit for this science case despite the inherent complexity of the science goals, derived empirical scaling relations between the signal-to-noise ratio of kinematic and intensity maps and the main telescope and instrument parameters, as well as a relation between the limit in stellar mass that can be reached for a given signal-to-noise ratio as a function of redshift. We specifically investigated the impact of AO performance on the science goal. We did not identify any breaking points with respect to all parameters (e.g., the telescope diameter), with the exception of the telescope thermal background, which strongly limits the performance in the highest (z$>$5) redshift bin. We find that the full range of science goals can be achieved with a $\\sim$100 nights program on the E-ELT, provided a high multiplex advantage $M\\sim N_{gal}/8$. We stress that several assumptions and guided guesses had to be made on both the observational conditions and physical characteristics of distant galaxies under study. This introduces an inherent uncertainty, which can be mitigated with future simulations as the telescope design will be consolidated and more details about the physics of high-z galaxies will become available." }, "0911/0911.2212_arXiv.txt": { "abstract": "We have used the Gemini Near-infrared Integral Field Spectrograph (NIFS) to map the gas kinematics of the inner $\\sim$\\, 200$\\times$500\\,pc of the Seyfert galaxy NGC\\,4151 in the Z, J, H and K bands at a resolving power $\\ge$\\,5000 and spatial resolution of $\\sim$\\,8\\,pc. The ionised gas emission is most extended along the known ionisation bi-cone at position angle PA=60--240$\\degr$, but is observed also along its equatorial plane. This indicates that the AGN ionizes gas beyond the borders of the bi-cone, within a sphere with $\\approx$\\,1\\arcsec\\ radius around the nucleus. The ionised gas has three kinematic components: (1) one observed at the systemic velocity and interpreted as originating in the galaxy disk; (2) one outflowing along the bi-cone, with line-of-sight velocities between $-600$ and 600\\,\\kms\\ and strongest emission at $\\pm\\,(100-300)$\\,\\kms; (3) and another component due to the interaction of the radio jet with ambient gas. The radio jet (at PA=75--255$\\degr$) is not aligned with the NLR, and produces flux enhancements mostly observed at the systemic velocity, suggesting that the jet is launched close to the plane of the galaxy ($\\sim$ plane of the sky). The mass outflow rate, estimated to be $\\approx$\\,1\\,M$_\\odot$\\,yr$^{-1}$ along each cone, exceeds the inferred black hole accretion rate by a factor of $\\sim$\\,100. This can be understood if the Narrow-Line-Region (NLR) is formed mostly by entrained gas from the circumnuclear interstellar medium by an outflow probably originating in the accretion disk. This flow represents feedback from the AGN, estimated to release a kinetic power of $\\dot{E}\\approx\\,2.4\\,\\times10^{41}$\\,erg\\,s$^{-1}$, which is only $\\sim$\\,0.3\\% of the bolometric luminosity of the AGN. There is no evidence in our data for the gradual acceleration followed by gradual deceleration proposed by previous modelling of the [OIII] emitting gas. Our data allow the possibility that the NLR clouds are accelerated close to the nucleus (within 0$\\farcs$1 -- $\\approx$\\,6\\,pc) after which the flow moves at essentially constant velocity ($\\approx\\,600$\\kms), being consistent with NIR emission arising predominantly from the interaction of the outflow with gas in the galactic disk. The molecular gas exhiibits distinct kinematics relative to the ionised gas. Its emission arises in extended regions approximately perpendicular to the axis of the bi-cone and along the axis of the galaxy's stellar bar, avoiding the innermost ionised regions. It does not show an outflowing component, being observed only at velocities very close to systemic, and is thus consistent with an origin in the galaxy plane. This hot molecular gas may only be the tracer of a larger reservoir of colder gas which represents the AGN feeding. ", "introduction": "This is a continuing study of the narrow-line region (hereafter NLR) of the Seyfert galaxy NGC\\,4151 using data obtained with the Gemini Near-infrared Integral Field Spectrograph (NIFS). The data comprise spectra of the inner$~\\approx\\,200\\,\\times\\,500\\,$pc$^2$, at a spatial resolution of$~\\approx\\,8\\,$pc at the galaxy, covering the wavelength range$~0.95$--$2.51\\,\\mu$m at a spectral resolving power R$\\,\\geq 5000$. In a previous paper \\citep[][hereafter Paper\\,I]{sb09} we have used these data to map the NLR intensity distributions of $14$ emission lines, as well as their ratios. The main results were the distinct flux distributions and physical properties observed for the ionised, molecular and coronal gas. The ionised gas is co-spatial with the ionisation bi-cone observed in the optical \\foiii\\ emission at position angle (hereafter PA)$~60\\degr$ \\citep{evans93,hutchings98}, and seems to trace the outflow along the bi-cone \\citep{das05,crenshaw00a,hutchings99}. In the inner region, the NIR ionised-gas emission extends beyond the borders of the cone, which does not have a sharp apex as noted also by \\citet{kra08}. The \\hmol\\ molecular gas intensity distribution, on the other hand, avoids the region of the bi-cone and seems to originate in the galaxy disc, while the coronal gas emission is barely resolved. In another recent work \\citep{riffel09b}, we have studied the unresolved nuclear continuum showing that its origin is emission by hot dust within $\\approx$\\,4\\,pc from the nucleus -- as expected from the dusty torus postulated by the unified model \\citep{antmi85}. The origin of this structure is probably a dusty wind that originates in the outer parts of the accretion disk. This wind is probably clumpy, as suggested by recent models \\citep{elitzur06}, and necessary in order to allow the escape of radiation along the equatorial plane of the bi-cone in order to ionize the gas and produce the observed flux maps of Paper\\,I. In the present paper we use the NIFS data to map the NLR kinematics of NGC\\,4151. Although many papers have been devoted to such a study, most of them are based on long-slit spectroscopy obtained in the optical with the Hubble Space Telescope \\citep[e.g.][]{winge97,hutchings98,crenshaw00a,das05}. The present study was performed in the near-IR, a waveband which is less affected by dust, and using integral field spectroscopy, which allows a full two-dimensional (hereafter 2D) coverage of the NLR kinematics. With our data and analysis we aim to examine previous claims of acceleration of the gas along the NLR, quantify the mass outflow rate and the corresponding feedback power, as well as investigate the origin of the NLR gas. As NGC\\,4151 harbors the closest bright AGN, its NLR is one of the best suited for this type of study. Our approach for the analysis of the NLR kinematics of NGC\\,4151 is as follows. First we map the centroid velocity and velocity dispersion of the ionised, molecular and coronal gas, obtained from fits to the emission-line profiles and compare our results with those of previous studies. Then we use an alternative approach to map the NLR kinematics, made possible by integral field spectrographs: a ``velocity tomography'' of the emitting gas, obtained by slicing the line profiles in velocity bins of$~60$\\kms, producing channel maps which provide a clearer view of the velocity distribution in the different gas phases. We have grouped the channel maps together in a sequence of velocity bins, generating movies which can be recovered at the authors' website. \\footnote{http://www.if.ufrgs.br/$\\sim$thaisa/ifu\\_movies/ngc4151} We adopt in this paper the same distance to NGC\\,4151 as in Paper\\,I:$~13.3\\,$Mpc, corresponding to a scale at the galaxy of$~65\\,$pc\\,arcsec$^{-1}$ \\citep{mundell03}. The paper is organised as follows. In \\S\\,\\ref{data} we provide a quick description of the data, in \\S\\,\\ref{results} we present the centroid velocity and velocity dispersion measurements, in \\S\\,\\ref{tomography} we present the emission-line ``tomography'', in \\S\\,\\ref{discussion} we discuss our results, in \\S\\,\\ref{feed} we estimate the mass outflow rate along the NLR and compare with the mass accretion rate and in \\S\\,\\ref{conclusion} we present our conclusions. ", "conclusions": "\\label{conclusion} We have measured the kinematics of the ionised and molecular gas surrounding the nucleus of NGC\\,4151 using integral-field spectroscopy obtained with the NIFS instrument at the Gemini North telescope. The main results of this paper are: \\begin{itemize} \\item The ionised-gas kinematics are dominated by three velocity components: (1) extended emission at systemic velocity observed in a circular region around the nucleus; (2) a component outflowing along the bi-cone (PA$=60\\deg$); (3) another component due to the interaction of the radio jet with the galactic disk. \\item Regarding the extended emission at systemic velocity: The origin of this component is gas from the galaxy plane ionised by the AGN up to at least 1 arcsec (65\\,pc) from the nucleus, revealing that there is escape of ionizing radiation even along the equatorial plane of the bi-cone. \\item Regarding the outflowing component: Channel maps extracted along the emission-line profiles show high-velocity gas all the way from the nucleus to the border of the emitting region. This seems to contradict previous studies showing acceleration along the NLR. Our data suggest that the NLR clouds are accelerated very close to the nucleus (within $\\approx$ 10\\,pc), after which the flow moves at essentially constant velocity. \\item We do not detect NIR counterparts to the highest-velocity \\foiii\\ emission seen by previous authors. It may originate in very low-density gas, which could be ablated material flowing off NLR clouds. This emission would not be characteristic of the bulk of the NLR mass. \\item The origin of the bulk of NLR mass seems to be entrained gas from the galaxy plane. This interpretation is supported by the large mass-outflow rate in ionised gas of $\\sim$\\,1\\,M$_\\odot$\\,yr$^{-1}$ obtained for the outflowing component in each cone, which is $\\sim$100 times the nuclear mass accretion rate. \\item The estimated kinetic power of the outflow is $\\approx2.4\\times10^{41}$\\,erg\\,s$^{-1}$, which is $\\sim$\\,0.3\\% of L$_{bol}$, and is $\\approx$\\,10 times larger than values obtained previously by our group for other nearby Seyfert galaxies. \\item Regarding the third kinematic component: Our data clearly show the interaction of the radio jet with ambient gas at systemic velocity. This has not been seen previously and indicates that the jet is launched close to the plane of the galaxy, pushing the gas encountered on its way. \\item Molecular hydrogen emission arises in extended regions along the axis of the galaxy's stellar bar, avoiding the region of the bi-cone. The \\hmol\\ velocities are close to systemic with a small rotation component, supporting an origin in the galaxy plane. The emitting molecular gas is probably just the ``hot skin'' of a larger reservoir that may have been built up by the previously observed HI inflow along the bar of the galaxy, and is probably the source of the AGN feeding." }, "0911/0911.5574_arXiv.txt": { "abstract": "{ We report an indication ($3.22 \\sigma$) of $\\approx 1860$ Hz quasi-periodic oscillations from a neutron star low-mass X-ray binary 4U 1636-536. If confirmed, this will be by far the highest frequency feature observed from an accreting neutron star system, and hence could be very useful to understand such systems. This plausible timing feature was observed simultaneously with lower ($\\approx 585$ Hz) and upper ($\\approx 904$ Hz) kilohertz quasi-periodic oscillations. The two kilohertz quasi-periodic oscillation frequencies had the ratio of $\\approx 1.5$, and the frequency of the alleged $\\approx 1860$ Hz feature was close to the triple and the double of these frequencies. This can be useful to constrain the models of all the three features. In particular, the $\\approx 1860$ Hz feature could be (1) from a new and heretofore unknown class of quasi-periodic oscillations, or (2) the first observed overtone of lower or upper kilohertz quasi-periodic oscillations. Finally we note that, although the relatively low significance of the $\\approx 1860$ Hz feature argues for caution, even a $3.22 \\sigma$ feature at such a uniquely high frequency should be interesting enough to spur a systematic search in the archival data, as well as to scientifically motivate sufficiently large timing instruments for the next generation X-ray missions. ", "introduction": "Quasi-periodic oscillations (QPOs) are observed from many neutron star low-mass X-ray binary (LMXB) systems \\citep{vanderKlis2006}. There are several kinds of QPOs, such as millihertz QPO, horizontal branch QPO, normal/flaring branch QPO, hectohertz QPO, lower kilohertz (kHz) QPO, upper kHz QPO, etc. \\citep{vanderKlis2006}. In this paper we will mostly concentrate on high frequency QPOs, that can currently be observed only with the proportional counter array (PCA) instrument of the {\\it Rossi X-ray Timing Explorer} ({\\it RXTE}) space mission. KHz QPOs are such high frequency QPOs, and often they appear as a pair. For a given source, their frequencies normally move up and down together in the $\\approx 200-1200$ Hz range in correlation with the source state. KHz QPOs are scientifically very important for the following reasons: (1) their high frequencies indicate that they originate from regions close to the neutron stars, and hence could be used as a tool to probe the strong gravity region around these stars, as well as the neutron star properties, and (2) they have been observed from many sources, and repeatedly from a given source, with high significance. Indeed, there are several models that involve relativistic orbital frequencies and the neutron star spin frequency, as well as beating and resonances among them \\citep{StellaVietri1998, Lamb+1985, Miller+1998, AbramowiczKluzniak2001, Torok+2008, Wijnands+2003, LambMiller2003, Zhang2004, Mukhopadhyay2009}. However, although there are many kHz QPO models available in the literature, none of them can explain all the major properties of this timing feature. Therefore, these QPOs cannot yet be used as a reliable tool to probe the strong gravity region, or to measure the neutron star parameters. An intriguing aspect of the kHz QPOs is that so far they have never been observed with a frequency greater than 1330 Hz \\citep{vanderKlis2006}, while PCA can easily detect a feature with a much higher frequency. In fact, even this 1330 Hz QPO was not confirmed by a later analysis \\citep{Boutelieretal2009b}. This implies that no timing feature has ever been observed with $> 1330$ Hz from an accreting neutron star system. In this paper, we report the indication of an $\\approx 1860$ Hz QPO from the persistent neutron star LMXB 4U 1636-536 (see \\citet{Altamirano+08} for the other timing features). This plausible QPO was observed simultaneously with the pair of kHz QPOs, which suggests that the $\\approx 1860$ Hz QPO could be from a new and previously unknown class of very high frequency QPOs. Note that kHz QPOs from 4U 1636-536 were first discovered by \\citet{vanderKlisetal1996, Zhangetal1996a, Zhangetal1996b}, and later were reported by many authors \\citep{Wijnandsetal1997, Vaughanetal1997, Zhangetal1997, Vaughanetal1998, Psaltisetal1998, Mendezetal1998, Mendez1998, MisraShanthi2004, Kaaret+1999, Psaltisetal1999, Markwardtetal1999, Jonkeretal2000, Fordetal2000, Mendez2002, Jonkeretal2002, DiSalvoetal2003, MisraShanthi2004, Jonkeretal2005, Barretetal2007, Bellonietal2007, Torok+2008}. ", "conclusions": "\\label{Discussion} In this paper, we report the detection of a pair of kHz QPOs from a neutron star LMXB 4U 1636-536. The ratio of these QPO frequencies is roughly 1.5, which is somewhat consistent with the resonance model described in \\citet{Abramowiczetal2003} (but see \\citet{Boutelieretal2009a}; see also \\citet{Zhangetal2006}). It is generally observed that the separation of the twin kHz QPO frequencies clusters around the neutron star spin frequency, or half of that. Models have been proposed in order to explain this aspect. However, for the kHz QPOs reported in this paper, the separation is $\\approx 319$ Hz, which is quite different from the half of the stellar spin frequency (see also \\citet{MendezBelloni2007, Yinetal2007}). Note that the neutron star of 4U 1636-536 spins with a rate of 582 Hz \\citep{StrohmayerMarkwardt2002}. Apart from the kHz QPOs, we have found an indication of a QPO at $\\approx 1860$ Hz. This is by far the highest frequency feature observed from an accreting neutron star system, and hence could be extremely useful to understand such systems. Moreover, this plausible QPO was observed simultaneously with the lower and the upper kHz QPOs, which means that the $\\approx 1860$ Hz QPO could be from a new class of QPOs. We will now briefly discuss the plausible origin of this QPO using general arguments, and without going into specifics. In some of the models, the lower and upper kHz QPO and some other QPO frequencies are thought to be one or more of the following frequencies (for equatorial circular orbits in Kerr spacetime; \\citet{vanderKlis2006}): (1) Keplerian orbital frequency $\\nu_{\\phi} = \\nu_{K}(1+j(r_g/r)^{3/2})^{-1} = (2\\pi)^{-1}(GM/r^3)^{1/2}(1+j(r_g/r)^{3/2})^{-1}$; (2) radial epicyclic frequency $\\nu_r = \\nu_{\\phi}(1-6(r_g/r)+8j(r_g/r)^{3/2}-3j^2(r_g/r)^2)^{1/2}$; (3) vertical epicyclic frequency $\\nu_{\\theta} = \\nu_{\\phi}(1-4j(r_g/r)^{3/2}+3j^2(r_g/r)^2)^{1/2}$; (4) periastron precession frequency $\\nu_{\\rm peri} = \\nu_{\\phi} - \\nu_r$; and (5) nodal precession frequency $\\nu_{\\rm nodal} = \\nu_{\\phi} - \\nu_{\\theta}$. Here the last four frequencies are for infinitesimally tilted and eccentric orbits, $M$ is the neutron star mass, $r_g = GM/c^2$, $j = Jc/GM^2$, $J$ is the total angular momentum of the neutron star, and $r$ is the distance of an orbit from the centre of the neutron star. For a sample of radio pulsars, $M$ was found to be distributed around $1.35 M_\\odot$ in a narrow Gaussian ($\\sigma = 0.04 M_\\odot$; \\citet{ThorsettChakrabarty1999}). Therefore, considering $M = 1.35 M_\\odot$, we find that only $\\nu_{\\phi}$ or $\\nu_{\\theta}$ can be the frequency of the plausible $\\approx 1860$ Hz QPO, and that too for relatively large angular momentum parameter values (as far as a neutron star is concerned) and for orbits very close to the innermost stable circular orbit (ISCO; see Fig.~\\ref{freq1}). As the neutron star mass increases due to accretion, $M$ is expected to be greater than $1.35 M_\\odot$ for accreting neutron stars (such as the one in 4U 1636-536). Moreover, since $\\nu_{K} = (c^3/2\\pi G)(r/r_g)^{-3/2}(1/M)$, $\\nu_{K}$ decreases as $M$ increases for a given $r/r_g$. Therefore, it is unlikely that any of the above mentioned five frequencies can be the frequency of the plausible $\\approx 1860$ Hz QPO. Besides, since the spin frequency of the neutron star in 4U 1636-536 is $\\nu_{\\rm spin} = 582$ Hz, the beating of $\\nu_{\\rm spin}$ with any of $\\nu_{\\phi}$, $\\nu_r$ or $\\nu_{\\theta}$ cannot explain the frequency of this QPO. However, the $\\approx 1860$ Hz QPO could be an overtone of any of $\\nu_{\\phi}$, $\\nu_r$, $\\nu_{\\theta}$ or $\\nu_{\\rm peri}$, or beating between an overtone and a fundamental of these frequencies. This plausible QPO could also be an overtone of one of the kHz QPOs (see \\S~\\ref{DataAnalysisandResults}). Although it is unlikely that the $\\approx 1860$ Hz QPO frequency is a Keplerian frequency $\\nu_{\\phi}$, it is still instructive to examine what constraints $\\nu_{\\phi} = 1860$ Hz can impose on the neutron star parameter values. This is because, the measurement of neutron star parameters provides the only way to constrain the theoretically proposed equation of state (EoS) models of neutron star cores, and hence to understand the nature of supranuclear core matter \\citep{LattimerPrakash2007, Bhattacharyyaetal2000, Bhattacharyyaetal2001a}. Various authors suggested ways to constrain the neutron star parameters assuming one of the kHz QPO frequencies as a Keplerian frequency \\citep{Kaaretetal1997, Miller+1998, Zhangetal2007, Zhang2009}. For example, this assumption, and the following two reasonable conditions can constrain the mass ($M$) and the radius ($R$) of a neutron star \\citep{Miller+1998}: \\begin{eqnarray} R \\le r, \\label{freq4} \\end{eqnarray} where $r$ is the radius of the orbit associated with the kHz QPO via the expression of $\\nu_{\\phi}$; and \\begin{eqnarray} r_{\\rm ISCO} \\le r, \\label{freq5} \\end{eqnarray} where $r_{\\rm ISCO}$ is the radius of the ISCO. This is because the first condition gives a mass-dependent upper limit on $R$ via the expression of $\\nu_{\\phi}$; and the second condition gives an upper limit on $M$: $M < c^3/(2\\pi 6^{3/2}G\\nu_\\phi|_r)$ (for Schwarzschild spacetime). Therefore, if 1860 Hz were a Keplerian frequency, then the upper limit of neutron star mass and radius would be $\\approx 1.2 M_\\odot$ and $\\approx 10.5$ km respectively. Such upper limits would support the strange star EoS models \\citep{Chengetal1998, Bhattacharyyaetal2001b}, although the less exotic neutron star EoS models could not be completely ruled out. \\begin{figure}[h] \\centering \\includegraphics[width=9.0cm, angle=0]{fig3.ps} \\begin{minipage}[]{85mm} \\caption{Radial profiles of various frequencies (colour coded) of equatorial circular orbits in Kerr spacetime (see \\S~\\ref{Discussion}). Two angular momentum parameters ($j = 0.0$ (solid) and $j = 0.3$ (dotted)), and neutron star mass $= 1.35 M_\\odot$ are used. Dashed horizontal line exhibits the 1860 Hz frequency. This figure shows that these frequencies, that are often used to explain kHz and some other QPOs, cannot possibly explain a $\\approx 1860$ Hz QPO. }\\end{minipage} \\label{freq1} \\end{figure} The uniquely high frequency of the plausible $\\approx 1860$ Hz QPO and its coexistence with both the kHz QPOs make it extremely interesting. Furthermore, although no such high frequency signal was claimed previously, \\citet{vanderKlis2006} mentioned, ``the distinct impression of the observers is that there is still much hiding below the formal detection levels\". It should, therefore, be worthwhile to report this feature, which will spur a systematic search of such high frequency features in the archival {\\it RXTE} PCA data, as well as in the large area xenon proportional counters (LAXPC) instrument data of the upcoming {\\it Astrosat} space mission. Moreover, next generation X-ray timing instruments, such as the high timing resolution spectrometer (HTRS; proposed for the {\\it International X-ray Observatory}) and the proposed Si pixel detector of the {\\it Advanced X-ray Timing Array}, will have much better capability to detect weak QPOs (see, for example, Fig.~2 of \\citet{Barret+08}; \\citet{Chakrabarty+08}). The plausible $\\approx 1860$ Hz QPO will therefore motivate the future X-ray timing instruments." }, "0911/0911.2448_arXiv.txt": { "abstract": "We compare the relative merits of AGN selection at X-ray and mid-infrared wavelengths using data from moderately deep fields observed by both {\\em Chandra} and {\\em Spitzer}. The X-ray-selected AGN sample and associated photometric and spectroscopic optical follow-up are drawn from a subset of fields studied as part of the Serendipitous Extragalactic X-ray Source Identification (SEXSI) program. Mid-infrared data in these fields are derived from targeted and archival {\\em Spitzer} imaging, and mid-infrared AGN selection is accomplished primarily through application of the IRAC color-color AGN `wedge' selection technique. Nearly all X-ray sources in these fields which exhibit clear spectroscopic signatures of AGN activity have mid-infrared colors consistent with IRAC AGN selection. These are predominantly the most luminous X-ray sources. X-ray sources that lack high-ionization and/or broad lines in their optical spectra are far less likely to be selected as AGN by mid-infrared color selection techniques. The fraction of X-ray sources identified as AGN in the mid-infrared increases monotonically as the X-ray luminosity increases. Conversely, only 22\\%\\ of mid-infrared-selected AGN are detected at X-ray energies in the moderately deep ($\\langle t_{\\rm exp} \\rangle \\approx 100$~ks) SEXSI {\\em Chandra} data. We hypothesize that IRAC sources with AGN colors that lack X-ray detections are predominantly high-luminosity AGN that are obscured and/or lie at high redshift. A stacking analysis of X-ray-undetected sources shows that objects in the mid-infrared AGN selection wedge have average X-ray fluxes in the $2 - 8$~keV band three times higher than sources that fall outside the wedge. Their X-ray spectra are also harder. The hardness ratio of the wedge-selected stack is consistent with moderate intrinsic obscuration, but is not suggestive of a highly obscured, Compton-thick source population. It is evident from this comparative study that in order to create a complete, unbiased census of supermassive black hole growth and evolution, a combination of sensitive infrared, X-ray and hard X-ray selection is required. We conclude by discussing what samples will be provided by upcoming survey missions such as {\\em WISE}, {\\em eROSITA}, and {\\em NuSTAR}. ", "introduction": "\\label{sexsiconclusion_sec:intro} The tight correlation of nuclear black hole mass with the velocity dispersion and mass of the galactic bulge \\citep[e.g.,][]{Magorrian:98,Ferrarese:00,Tremaine:02} implies that the growth and evolution of galaxies is closely linked to the growth and evolution of the supermassive black holes which reside in (at least) all massive galaxies \\citep[e.g.,][]{Kauffmann:00,Heckman:08}. Indeed, recent theoretical work suggests that feedback from active galactic nuclei (AGN) plays a dominant role in establishing the present-day appearances of galaxies, providing a natural, physical explanation for both cosmic downsizing and the possibly related bimodality in local galaxy properties \\citep[e.g.,][]{Scannapieco:05, Hopkins:08, Cattaneo:09}. However, obtaining an unbiased census of black holes in the universe remains challenging, hampering our ability to fully probe this connection. Most surveys for active galaxies are severely biased towards unobscured ({\\em type~1}) AGN since nuclear emission in such sources dominates over host galaxy light at most wavelengths, making type~1 AGN both more readily identifiable and easier to follow up spectroscopically. However, both unified AGN models \\citep[e.g.,][]{Antonucci:93,Urry:95} and the shape of the X-ray background suggest a population of obscured ({\\em type~2}) sources which outnumber the type~1 AGN by up to a factor of ten \\citep[e.g.,][]{Comastri:95, Treister:04, Gilli:07}. Determining the ratio of unobscured to obscured AGN as a function of luminosity and redshift can directly constrain the growth history of supermassive black holes, and help to quantify their influence on the evolution of their host galaxies \\citep[e.g.,][]{Ueda:03,Barger:05,Hasinger:05,Hopkins:07}. Since different search techniques for obscured AGN suffer from different biases, the problem is best addressed through multiwavelength studies. The hard (\\hardrange) X-ray and mid-infrared wavebands provide powerful, complementary methods for identifying and studying AGN over a wide range of intrinsic obscuration. Radiation seen from active galaxies at X-ray energies is primarily due to direct emission from accretion processes near the central supermassive black hole, with higher energy photons less susceptible to absorption, while mid-infrared radiation (rest-frame $\\lambda_0 \\simgt 2~\\mu$m) is typically dominated by emission from dusty obscuring material surrounding the AGN central engine, and is likewise relatively immune to absorption \\citep[e.g.,][]{Krolik:99}. Hard X-ray and mid-infrared surveys of AGN will therefore sample a wide range of intrinsic absorption and, in this regard, will be both less biased and more complete than surveys in the optical \\citep[e.g.,][]{Richards:06} and soft X-ray \\citep[$E \\simlt 2.4$~keV; e.g.,][]{Hasinger:98, Schmidt:98} bands, both of which suffer from strong attenuation by even moderate (\\nh$\\simgt 10^{21}$~cm$^{-2}$) columns of dusty, obscuring material. However, AGN samples selected in the \\hardrange\\ X-ray and infrared bands will also suffer incompleteness. Examples exist of AGN identified from optical spectroscopy which remain undetected even in the deepest X-ray images yet obtained \\citep[e.g.,][]{Steidel:02}; similarly, there are examples of optically selected AGN not identified in infrared surveys (e.g., \\citealt{Stern:05b}). Conversely, X-ray missions have identified AGN whose optical spectra are devoid of AGN signatures (e.g., X-ray bright, optically normal galaxies, or ``XBONGs''; \\citealt{Comastri:02}). Therefore a complete census will necessarily require combining selection techniques from several spectral regimes. The current generation of X-ray telescopes has dramatically advanced our knowledge of the AGN population (see~\\citealt{Brandt:05}\\ for a review). The {\\em Chandra X-ray Observatory} \\citep{Weisskopf:96} provides a large collecting area, a moderate field of view (FOV), and exquisite angular resolution ($<1$\\arcsec) from 0.5 to 8~keV. These capabilities have allowed {\\em Chandra} extragalactic surveys to efficiently select and optically identify large samples of AGN in the 2 -- 8~keV range. Observing in this hard X-ray energy band means that even sources shrouded by considerable obscuring column densities (up to $10^{24}$ cm$^{-2}$) can be detected at low redshift, and sources with even higher column densities can be detected out to $z \\simgt 2$. Previous sensitive X-ray telescopes were primarily restricted to energies below $\\sim 2$~keV and thus missed many of the obscured AGN, a source population that has long been theorized to explain the mismatch in spectral shape between the \\hardrange\\ X-ray background ($\\Gamma \\approx 1.4$) and the unobscured active galaxies which dominate the source counts at soft X-ray energies ($\\Gamma\\approx 1.9$; \\citealt{Nandra:94}). The 2003 launch of the {\\em Spitzer Space Telescope} \\citep{Werner:04} opened a new era in mid-infrared observations, providing orders of magnitude improvement in sensitivity in the $3.6-160~\\mu$m band. The increased sensitivity, combined with the large FOV, allows for the first time efficient long-wavelength survey capabilities. The primary imaging cameras on {\\it Spitzer} are the Infrared Array Camera \\citep[IRAC;][]{Fazio:04}, providing simultaneous imaging at 3.6, 4.5, 5.8, and 8~$\\mu$m, and the Multiband Imaging Photometer for {\\em Spitzer} \\citep[MIPS;][]{Rieke:04}, providing simultaneous imaging at 24, 70, and 160~$\\mu$m. Several methods have been developed to select AGN based on their {\\em Spitzer} colors~\\citep{Lacy:04,Stern:05b,Alonso-Herrero:06}. These methods exploit the difference in the typical spectral energy distribution (SED) of AGN compared to `normal' galaxies. The near-infrared emission of both typical galaxies and vigorously star-forming galaxies is primarily produced by a thermal stellar population, resulting in SEDs peaked near the rest-frame 1.6~$\\mu$m ``bump.'' This bump, which is caused by the minimum in the opacity of the H$^{-}$ ion near 1.6~$\\mu$m \\citep[e.g.,][]{John:88}, is a feature of almost all stellar populations \\citep[e.g.,][]{Wright:94,Simpson:99}, whereas AGN-dominated SEDs have a non-thermal, roughly power-law shape (for $\\lambda \\simlt 10~\\mu$m). At longer wavelengths contributions from stellar blackbody emission is low, while radiation from the obscuring dust near the central engine provides strong, relatively isotropic emission which is quite distinct from stellar light (although the shape of the AGN SED does somewhat depend on viewing angle, e.g., \\citealt[][]{Elitzur:08}). \\citet{Stern:05b}, for example, suggested a selection technique exploiting the fact that, for AGN, the long-wavelength side of the 1.6~$\\mu$m stellar peak does not decline; the technique uses an empirically determined `wedge' in IRAC color-color space ([3.6]$-$[4.5] versus [5.8]$-$[8.0]) that preferentially contains AGN as compared to normal galaxies or Galactic stars. Combining a sample of 10,000 $R<21.5$ spectroscopically identified sources from the AGN and Galaxy Evolution Survey (AGES; Kochanek et al., in prep.) and mid-infrared observations from the IRAC Shallow Survey \\citep{Eisenhardt:04}, \\citet{Stern:05b} defined mid-infrared AGN selection criteria which robustly identify broad- and narrow-lined AGN, with only 18\\% sample contamination from galaxies (17\\%) and stars (1\\%). The true sample contamination is likely lower, since many of the spectroscopically normal galaxies may harbor active nuclei (i.e., are mid-infrared versions of XBONGs). Working from the full spectroscopically defined AGES sample, \\citet{Stern:05b} found that the wedge selects 91\\% of the broad-lined AGN, 40\\% of the narrow-lined AGN, and fewer than 3\\% of the normal galaxies. This paper presents an exploration of the relative strengths of {\\em Chandra} and {\\em Spitzer} as black-hole finders, using a subset of six fields from the Serendipitous Extragalactic X-ray Source Identification program (SEXSI; \\citealt{Harrison:03}, \\citealt{Eckart:05}, and \\citealt{Eckart:06}) for which we have obtained mid-infrared coverage with {\\em Spitzer}. In addition to examining the properties of $\\sim 250$ hard X-ray-selected AGN, we extend our sensitivity to X-ray emission using a stacking analysis on sources with different mid-infrared characteristics. This enables us to make a comprehensive comparison of the relative characteristics of AGN samples selected from their X-ray and mid-infrared properties. We organize the paper as follows: \\S\\ref{sexsiconclusion_sec:experiment} introduces the SEXSI program and presents the complementary mid-infrared observations and the resulting source catalog; \\S\\ref{sexsiconclusion_sec:xrayselected} discusses the mid-infrared properties of the \\hardrange\\ SEXSI sources; \\S\\ref{sexsiconclusion_sec:stacking} presents the mean X-ray properties of X-ray-non-detected {\\em Spitzer} sources; \\S\\ref{sexsiconclusion_sec:discussion} provides a discussion; and \\S\\ref{sexsiconclusion_sec:conclusions} summarizes our conclusions. Luminosity calculations assume a standard cosmology used in our previous work \\citep{Eckart:06}, $\\Omega_0 = 0.3$, $\\Lambda = 0.7$, and $H_0 = 65~ {\\rm km}~ {\\rm s}^{-1}~ {\\rm Mpc}^{-1}$. ", "conclusions": "\\label{sexsiconclusion_sec:conclusions} We have presented an exploration of the relative efficiency of selecting AGN in the X-ray with {\\em Chandra} and in the mid-infrared with {\\em Spitzer}. In this section we summarize our comparison of the selection techniques and make predictions about the AGN samples that will be selected with future X-ray and mid-infrared surveys. \\subsection{Summary} \\label{sexsiconclusion_sec:midIRsummary} Two-thirds of the X-ray-selected AGN are also selected by the \\citet{Stern:05b}\\ {\\em Spitzer} AGN selection criteria. Nearly all of the SEXSI X-ray sources that are classified as either broad-line or narrow-line AGN (i.e., those that exhibit high-ionization optical/UV emission lines in their optical spectra) are consistent with wedge selection within photometric uncertainties. A large fraction of the low luminosity ($L_{\\rm 2-10~keV} < 10^{43.5} \\lumin$), hard X-ray (\\hardrange) AGN identified by {\\em Chandra} will be missed by {\\em Spitzer} selection techniques. These low X-ray luminosity sources tend to lack the high-ionization emission lines that allow AGN identification via optical spectroscopy; thus, the sources that are missed by the {\\em Spitzer} color-color AGN selection will also remain unidentified in optical spectroscopic surveys. These sources tend to lie at lower redshift and have {\\em higher} fluxes in the infrared, on average. This suggests that many of these non-wedge-selected X-ray sources have mid-infrared fluxes that are dominated by stellar (starburst) emission as opposed to AGN activity; indeed, most have optical spectra of ELGs, consistent with the conclusion that emission related to star formation outshines the emission from the active nucleus in some parts of the spectrum (i.e., in the mid-infrared and optical), while the X-ray luminosities of these sources are dominated by X-ray emission from the active nuclei. In addition, we find that while a fair fraction ($\\sim 63$\\%) of the optically bright ($R<21$) wedge-selected {\\em Spitzer} AGN are also identified by {\\em Chandra}, over 80 percent of the optically faint wedge sources are not detected in the X-ray. It is likely that these fainter sources are AGN that are heavily obscured and/or lie at high redshift. We also find that X-ray-selected high-redshift, type-2 quasars are also selected via the $3.6~\\mu$m$-24~\\mu$m selection criteria proposed by \\citet{Martinez-Sansigre:05}; the radio selection criteria used by those authors to eliminate starburst contaminants is not necessary when the sources have X-ray properties indicative of AGN activity. To explore the mean X-ray properties of the X-ray undetected, IRAC-selected AGN candidates, we have stacked the {\\em Chandra} data extracted from the positions of IRAC sources. The stacked wedge sources show significant X-ray signals in the full, soft, and hard X-ray bands. The average hardness ratio of the stacked spectrum of sources inside the wedge is higher than that of sources outside of the wedge, and the average $2-8$~keV flux is three times larger. The hardness ratio of the wedge-selected stacked spectrum is consistent with moderate intrinsic obscuration (\\nh$\\sim 10^{22}-10^{23.7}$~cm$^{-2}$), but is not suggestive of a highly obscured, Compton-thick source population. \\subsection{Conclusions} \\label{sexsiconclusion_sec:finalconclusions} The results of our study illustrate the challenges of identifying obscured and low-luminosity AGN using any one technique. A complete understanding of AGN activity and black hole growth will require multiwavelength data sets: low-luminosity AGN in galaxies with active star formation cannot be selected using mid-infrared color techniques, but are effectively found at low redshift with the current-generation of X-ray telescopes; obscured high-redshift objects are beyond the sensitivity threshold of {\\em Chandra} and {\\em XMM}, but they are effectively identified in the mid-infrared if the AGN luminosity is at the high end of the luminosity distribution. The results from current X-ray and mid-infrared satellites enable us to estimate the progress possible with upcoming space missions that plan to conduct all-sky or pencil beam surveys. The {\\em Wide-field Infrared Survey Explorer} ({\\em WISE;} \\citealt{Mainzer:05}), scheduled for launch at the end of 2009, will provide full-sky mid-infrared imaging reaching a depth of 120~$\\mu$Jy at 3.3~$\\mu$m (similar to IRAC channel~1) and 160~$\\mu$Jy at 4.7~$\\mu$m (similar to IRAC channel~2). At these shallow depths, almost every source redder than [3.3]$-$[4.7] $\\approx 0.5$ will be an AGN; high-redshift galaxies are too faint to make this flux cut, and the coolest brown dwarfs are too rare. The primary contaminant will be actively star-forming galaxies at redshifts of a few tenths, which should be readily identifiable from their optical morphologies and colors in relatively shallow imaging such as is available from the Sloan Digital Sky Survey. By applying a flux cut at the {\\em WISE} depths to the Bo\\\"{o}tes IRAC survey \\citep{Eisenhardt:04,Ashby:09}, we predict that {\\em WISE} should detect approximately 85 AGN candidates per square degree, of which $\\approx 15\\%$ will be low-redshift star-forming galaxy contaminants. {\\em WISE} will therefore provide an unprecedented sample of high-luminosity obscured AGN across a range of redshifts up to $z \\sim 3$ identified based on mid-infrared color selection alone. {\\em WISE} will also detect numerous other low-redshift ($z < 1$), low-luminosity ($L_{\\rm 2-10~keV} \\simlt 10^{43} \\lumin$) active galaxies, but the significant number of open symbols above the {\\em WISE} threshold shown in Figure~\\ref{sexsiconclusion_fig:f36vsfx} shows that mid-infrared colors alone are insufficient to identify them as AGN. The {\\em extended R\\\"ontgen Survey with an Imaging Telescope Array} ({\\em eROSITA;} \\citealt{Predehl:06}) X-ray mission, planned for launch in 2011, will provide full-sky hard X-ray images to a depth of $f_{\\rm 2-10~keV} = 1.5 \\times 10^{-13}~\\fluxu$. Using the sample of XBo\\\"{o}tes sources \\citep{Kenter:05} with fluxes above the {\\em eROSITA} hard-band sensitivity limit, we predict that {\\em eROSITA} should detect $\\approx 4$ hard X-ray sources per square degree. The mission will detect many more AGN in the soft X-ray band because the effective area is much larger at energies below 2~keV, reaching soft-band depths of $f_{\\rm 0.5-2~keV} = 9 \\times 10^{-15}~\\fluxu$. Scaling from the XBo\\\"{o}tes \\softrange\\ sample we predict that {\\em eROSITA} will detect $\\approx 120$ sources per square degree in the soft band. Figure~\\ref{sexsiconclusion_fig:f36vsfx} shows that {\\em eROSITA} will identify a large population of AGN below the {\\em WISE} threshold, most of which we project will be at $z \\simgt 1$. The majority of these sources will be unobscured BLAGN, but a significant population will be narrow-line, obscured AGN (NLAGN and ELG). However, {\\em eROSITA} does not reach sufficient depth to detect many of the low-luminosity, low-redshift AGN detected, but not identified as such, by {\\em WISE}. Clearly {\\em WISE} and {\\em eROSITA} combined will miss (or mis-identify) low-luminosity obscured AGN at all redshifts, most of which will also not be found in optical and soft X-ray surveys. At low redshift only a sensitive X-ray mission extending to E$_x > 10$~keV will effectively uncover this population, which will be visible only in reprocessed light, if at all, at E$_x < 10$~keV. The {\\em Nuclear Spectroscopic Telescope Array} ({\\em NuSTAR;} \\citealt{Harrison:05})\\footnote{\\tt http://www.nustar.caltech.edu.}, scheduled for launch in 2011, will make major advances in studying faint extragalactic sources at energies above 10~keV. {\\em NuSTAR} will be the first focusing, high-energy X-ray satellite and is designed to make targeted observations of the hard X-ray sky from $6-79$~keV. {\\em NuSTAR}, which has a field-of-view comparable to {\\em Chandra}, plans to cover the Bo\\\"{o}tes and Great Observatories Origins Deep Survey (GOODS) fields, and should resolve $\\simgt 50$\\% of the 30~keV X-ray background, reaching $10-40$~keV flux limits of $< 10^{-14}~\\fluxu$. Many of the resolved sources are expected to be low-luminosity, nearby AGN not identified as such in current mid-infrared or \\hardrange\\ X-ray surveys, and some {\\em NuSTAR} sources may be AGN with hard spectra owing to low-efficiency accretion. For studying the high-redshift, low-luminosity population, the {\\em International X-ray Observatory} ({\\em IXO}) will reach \\hardrange\\ flux limits of $\\sim 5 \\times 10^{-17}~\\fluxu$ in a megasecond exposure with its wide-field imager. It could survey up to 10~deg$^2$ to a depth of $10^{-16}~\\fluxu$ and 0.25~deg$^2$ to a depth of $10^{-17}~\\fluxu$, uncovering the largest population of high-redshift AGN yet detected. Predictions for the AGN populations that will be uncovered by these future missions must rely on careful studies of the AGN identified by existing X-ray and infrared surveys. As illustrated in this paper, an understanding of the AGN selection techniques of the various missions, and their relative strengths and weaknesses, is crucial as datasets are synthesized to obtain an understanding of black hole growth and evolution in the universe. Finally, we attempt to put the results of our study in the context of understanding the cosmic history of black hole growth and AGN accretion modes. The primary conclusions of this paper are that heavily obscured luminous AGN are often missed by X-ray selection, while low-luminosity AGN are often missed by mid-infrared selection. For the high-luminosity sources, various work has shown that, at least for the optically-selected unobscured sources, most are emitting at close to their Eddington limit \\citep[e.g.,][]{Kollmeier:06}. Multiwavelength observations are consistent with the heavily-obscured high-luminosity sources having similar intrinsic SEDs, simply with their UV, optical, and X-ray emission suffering from absorption. As for the low-luminosity AGN, there are two possibilities: they could be lower mass black holes again emitting at close to their Eddington limits, or they could be average-sized black holes, with masses similar to the high-luminosity sources, but simply emitting at lower Eddington ratios. Using a new set of empirical, low-resolution SED templates for AGN and galaxies, \\citet{Assef:09} shows that the likelihood of a source to be selected as an AGN using the \\citet{Stern:05b} mid-infrared color criteria is primarily a function of how strong the AGN is relative to the host galaxy --- e.g., the Eddington ratio assuming that, to first order, the host galaxy mid-infrared emission scales with the galaxy total stellar mass which scales with the nuclear supermassive black hole mass. Thus, the fact that we find few low-luminosity sources using the mid-infrared criteria suggests that the low-luminosity sources are {\\em not} emitting at close to their Eddington ratio. This is consistent with the results of \\citet{Babic:07}\\ who find that low-luminosity X-ray sources in the {\\em Chandra} Deep Field-South have a wide range of Eddington ratios, $10^{-5} \\simlt \\log(L_{\\rm bol}/L_{\\rm Edd}) \\simlt 1$. One caveat on this interpretation is that it assumes that the AGN SED is independent of both Eddington ratio and luminosity --- the former of which has recently been questioned by \\citet{Vasudevan:07}. Future work, in particular, higher energy observations with {\\em NuSTAR}, will quantify the extent to which the intrinsic AGN SEDs depend on both luminosity and Eddington ratio." }, "0911/0911.1161_arXiv.txt": { "abstract": "We analyze a solar active region observed by the {\\it Hinode} \\ion{Ca}{2}~H line using the time-distance helioseismology technique, and infer wave-speed perturbation structures and flow fields beneath the active region with a high spatial resolution. The general subsurface wave-speed structure is similar to the previous results obtained from {\\it SOHO}/MDI observations. The general subsurface flow structure is also similar, and the downward flows beneath the sunspot and the mass circulations around the sunspot are clearly resolved. Below the sunspot, some organized divergent flow cells are observed, and these structures may indicate the existence of mesoscale convective motions. Near the light bridge inside the sunspot, hotter plasma is found beneath, and flows divergent from this area are observed. The {\\it Hinode} data also allow us to investigate potential uncertainties caused by the use of phase-speed filter for short travel distances. Comparing the measurements with and without the phase-speed filtering, we find out that inside the sunspot, mean acoustic travel times are in basic agreement, but the values are underestimated by a factor of $20-40\\%$ inside the sunspot umbra for measurements with the filtering. The initial acoustic tomography results from {\\it Hinode} show a great potential of using high-resolution observations for probing the internal structure and dynamics of sunspots. ", "introduction": "Deriving subsurface structures and flow fields of solar active regions is one of the major topics for local helioseismology studies. Using \\ion{Ca}{2}~K observations made at the geographic South Pole in early 1990s, \\citet{duv96} found the first evidence of downdrafts below active regions through measuring acoustic travel times by employing the time-distance helioseismology technique. Later inversions \\citep{kos96} using these travel time measurements found strong converging downflows and areas of the excess of sound-speed beneath growing active regions. With the availability of {\\it Solar and Heliospheric Observatory} / Michelson Doppler Imager \\citep[{\\it SOHO}/MDI;][]{sch95} Doppler observations that are seeing-free and more suitable for local helioseismology studies, more investigations of the sunspot's subsurface structures and flow fields have been carried out by use of various theoretical models: ray-path approximation \\citep[e.g.,][]{kos00, zha01}, Fresnel-zone approximation \\citep[e.g.,][]{jen01, cou04}, and Born approximation \\citep[e.g.,][]{cou06}. Despite the use of different approaches of calculating travel time sensitivity kernels, the results remain largely the same, i.e., for active region subsurface structures, a negative sound-speed perturbation was found up to 5 Mm below the photosphere, and a positive perturbation was seen below that depth. For subsurface flow fields, converging downflows were inferred from the photosphere to about 5 Mm beneath it, and divergent flows were found below this layer. Another local helioseismology technique, ring-diagram analysis, although could not provide subsurface flow maps with such a high spatial resolution, gave results that were in general agreement \\citep{hab04, kom05}. Some direct comparisons of these two techniques were also performed \\citep{hin04}. However, on the other hand, these results were not uncontroversial. By measuring acoustic phase shifts in sunspot penumbra using MDI Doppler observations when sunspots were located near the solar disk limb, \\citet{sch05} demonstrated that the measured acoustic travel times varied around sunspot penumbra when the sunspot was located near the limb, and believed this effect was caused by the inclined magnetic field. While acknowledging the existence of such an effect in oscillation signals observed in Dopplergrams, \\citet{zha06} found out that such an effect did not exist in the oscillations observed in MDI continuum intensity. Additionally, \\citet{raj06} showed that the phase-speed filtering in the time-distance acoustic travel time measurements would also introduce errors in active regions where oscillation amplitudes were suppressed. Using numerical simulations \\citet{par08} confirmed this effect, and concluded that this effect led to underestimation of the mean travel times at short distances. Some recent attempts of using different filtering procedures, e.g., ridge filtering \\citep{bra08}, showed that different filtering might give different measurements. Thus, it is certainly worthwhile examining how filters would change measurement results. On the other hand, it is also arguable whether those time-distance observed acoustic travel time variations in solar magnetic areas could be interpreted as interior sound speed anomalies, or were caused by some surface magnetism effects such as showerglass effect \\citep{lin05a, lin05b}. Furthermore, it is also not clear how the interaction of acoustic waves and magnetic field would effect the phase of these waves, hence the travel time variations in the measured acoustic travel times. It would be very difficult to interpret the measurements if acoustic waves experience some phase shifts at the boundary of unmagnetized and magnetized areas \\citep{cal09}. High spatial resolution observation of sunspots by the Solar Optical Telescope \\citep[SOT;][]{tsu08} onboard the Japanese solar spacecraft {\\it Hinode} \\citep{kos07} not only gives us another possibility of investigating subsurface properties of solar active regions in addition to the already existing helioseismology instruments, but also provides an unprecedented high spatial resolution that has enabled some studies that were improbable using these existing instruments. In this paper, we analyze an active region observed by {\\it Hinode} utilizing the time-distance helioseismology technique, and infer the wave-speed profiles and flow fields beneath this region, and present these results in \\S3. We also analyze the acoustic signals without applying any filters except filtering out solar convection and $f$-modes, and investigate how the phase-speed filtering effects our measurements. Such analyses can hardly be carried out by use of MDI high resolution observations, and these results are presented in \\S4. In \\S5, we discuss and summarize our results. ", "conclusions": "\\subsection{Subsurface Structure and Flow Field} By a time-distance analysis of unprecedented high spatial resolution observations from the {\\it Hinode}/SOT, we have investigated high resolution wave-speed structures and mass flows beneath active region AR10953. For the subsurface wave-speed structure, the inverted results are remarkably similar to previous results based on MDI Dopplergrams with different inversion technique as well as different sensitivity kernels. For subsurface flow fields, the general picture is similar to what has been obtained from MDI observations, i.e., converging downward flows near the surface below the sunspot area. Despite the similarities, the current picture also has some differences compared with the earlier one. One difference is that the downward flows beneath the sunspot are more prominent in the flow fields inferred from this study. Another remarkable result from this study is the mass circulation outside the sunspot, which seems to keep mass conservative and is much more clear than the previous MDI inversions. It is widely believed in theory and in numerical simulations that such downdraft and converging flows near the sunspot surface play important roles in keeping sunspots stable \\citep{par79, hur00, hur08}. However, it is still not quite clear how the overall flow structures around the sunspot's surface and interior look like. It is already well known the existence of penumbral Evershed outflows, outgoing moat flows beyond sunspot's penumbra, and inflows from the inner penumbra and umbra measured by tracking penumbra/umbra dots \\citep[e.g.,][] {sob09}. For the sunspot's interior, there were reports of outflows from f-mode analysis \\citep{giz00}, inflows at the depth of 1.5 - 5 Mm (Zhao et al. 2001), and large scale inflows around active regions up to 10 Mm in depth from ring-diagram analysis \\citep{hab04}. Based on all these observations, \\citet{giz03} proposed a schematic flow structure of the sunspot, and \\citet{hin09} recently proposed a similar one. Both flow structures show outflows near the surface, and inflows below that, with transition depth of these opposite flows undetermined. Here, based on all previous results, as well as on the recent numerical simulation of Evershed flows \\citep{kit09}, we also propose a schematic flow structure of a sunspot (Figure~\\ref{flow_profile}), slightly different from the plots of two previous studies, but essentially similar. However, it is also acknowledged that this flow structure is not consistent with what was recently found by \\citet{giz09} and recent numerical simulations by \\citet{rem09}. Thus, further observational and theoretical studies are required to determine the subsurface dynamics of sunspots. The high resolution subsurface wave-speed maps (Figure~\\ref{cs_horiz}) and flow fields (Figure~\\ref{hr_flows}), present us an unprecedented opportunity to see the sunspot's subsurface structures and dynamics with many details. These results reveal the complexity of the subsurface perturbations and dynamics and their relationship with the sunspot structure. The subsurface image at the depth of $0 - 1.5$ Mm displays a larger wave-speed perturbation, presumably hotter in temperature, near the light bridge area than other areas inside the sunspot, where the wave-speed perturbations are largely negative. This is in agreement with the surface observations that light bridges are formed by upwelling of hot plasma \\citep{kat07}. We also see clear divergent flows from the light bridge in high resolution subsurface flow map at the same depth (Figure~\\ref{hr_flows}). This is consistent with the plasma upwelling that occurs in this area, but not in good agreement with the proper motion tracking results of this region \\citep{lou08}. It is also interesting to see quite a few divergent flow cells inside both the sunspot umbra and penumbra. The cells are bigger than granules and smaller than supergranules, and indicate the existence of a mesoscale convection inside the sunspot. The discovery of these motions is intriguing because they may be a signature of the downdraft vortex rings around magnetic flux bundles, as suggested by \\citet{par92}. The Parker's conjecture was that the observed clustering of magnetic flux bundles in sunspot is at least in part a consequence of hydrodynamic attraction of the downdraft vortex rings. Thus, it is very important to continue the investigations of the subsurface dynamics of sunspots, both observationally by high resolution helioseismology and theoretically by realistic MHD simulations. \\subsection{Analysis without Phase-Speed Filtering} As introduced in \\S1, the phase-speed filtering may introduce some errors in active regions due to the smaller oscillation amplitude in these areas \\citep{raj06}, and using ridge filtering may give different results as using phase-speed filtering in active regions \\citep{bra08}. All of these prompt us a reexamination of the use of phase-speed filtering and an evaluation of how phase-speed filtering alter measured acoustic travel times, in particular, inside active regions. The {\\it Hinode} high resolution observation has helped us to overcome an obstacle that acoustic travel times could hardly be measured in short distances inside active regions if phase-speed filtering was not applied. Our measured distance-dependent acoustic travel time differences relative to the solar quiet region obtained without using the phase-speed filtering, as shown in Figure~\\ref{td_diff}, have evidently demonstrated that for short distances, the acoustic travel time is longer inside both sunspot umbra and penumbra than the quiet region, and for distances longer than $\\sim 13$ Mm, the mean travel time is shorter. This is generally in agreement with the measurements using phase-speed filters, except that the ones with filters would underestimate the values by $20-40\\%$ in sunspot umbra. It is believed that the underestimation is caused by a combination of oscillation amplitude suppression inside active regions and the use of phase-speed filtering. Therefore, we believe that the phase-speed filtering does not change qualitatively the measured travel times, therefore the inferred interior properties of active regions, but would change the inferred values quantitatively. The {\\it Hinode} data also enable us to construct acoustic travel time maps without using filters, and this gives us an opportunity to assess the correctness and accuracy of various filters. Further investigations are particularly important in this regard. \\subsection{Summary} We summarize our results as follows: \\begin{enumerate} \\item By analyzing high resolution {\\it Hinode} \\ion{Ca}{2}~H observations using time-distance helioseismology, we have derived wave-speed structures beneath solar active regions. The inferred structures are remarkably similar to results obtained by various authors analyzing MDI Dopplergram observations, although the instrument, the spectrum line used to observe, the type of observation (intensity and Doppler velocity), the inversion technique, and the sensitivity kernels are mostly different. \\item The subsurface flow field structure is generally in agreement with previous MDI result, but it is clear that the downward flows near the sunspot surface is more prominent, and the mass circulation around the sunspot is more clear and seems self-consistent, although no mass conservation constraint is applied in the inversion. \\item Our analysis without using phase-speed filtering has convincingly demonstrated that the phase-speed filtering does not change travel time measurements qualitatively, but may underestimate travel time anomalies inside active regions. \\item High spatial resolution subsurface flow fields reveal quite a few organized flow structures inside sunspot umbra and penumbra, which are perhaps corresponding to some convection structures or downdraft vortex rings around magnetic flux bundles. In the light bridge area, subsurface hotter temperature is found, and plasma flows divergent from the light bridge is also seen. These initial results provide a basis for further development of high-resolution acoustic tomography of sunspots, and comparison with numerical simulations. \\end{enumerate}" }, "0911/0911.1357_arXiv.txt": { "abstract": "The equivalent width (EW) of the \\lya{} line is directly related to the ratio of star formation rates determined from Ly$\\alpha$ flux and UV flux density [SFR(\\lya{})/SFR(UV)]. We use published data --in the literature EW and SFR(\\lya{})/SFR(UV) are treated as independent quantities-- to show that the predicted relation holds for the vast majority of observed \\lya{} emitting galaxies (LAEs). We show that the relation between EW and SFR(\\lya{})/SFR(UV) applies irrespective of a galaxy's ``true'' underlying star formation rate, and that its only source of scatter is the variation in the spectral slope of the UV continuum between individual galaxies. The derived relation, when combined with the observed EW distribution, implies that the ratio SFR(UV)/SFR(Ly$\\alpha$) is described well by a log-normal distribution with a standard deviation of $\\sim 0.3-0.35$. This result is useful when modelling the statistical properties of LAEs. We further discuss why the relation between EW and SFR(\\lya{})/SFR(UV) may help identifying galaxies with unusual stellar populations. ", "introduction": "\\label{sec:intro} Lyman $\\alpha$ (hereafter Ly$\\alpha$) equivalent width represents a fundamental quantity in the study of Ly$\\alpha$ emitting galaxies. The equivalent width (EW) of the \\lya{} line emitted by galaxies is a sensitive indicator of the initial mass function (IMF) or gas metallicity from which stars form \\citep[e.g.][]{S02,S03}. The existence of large equivalent width (rest-frame $\\mathrm{EW} \\ga 240$\\,\\AA{}) \\lya{} emitters has led to speculation on whether population III formation from pristine gas may actually have been observed \\citep{MR02,J06,DW07}. However, this interpretation depends sensitively on the details of radiative transfer through both the intergalactic medium \\citep[IGM, see e.g.][for a discussion]{S08}, and the interstellar medium \\citep[ISM, e.g.][]{N90,H06,Fi}. It is safe to conclude that at present, no truly convincing candidates for population III galaxies exist. The search for large equivalent width \\lya{} emitters -- and population III galaxy formation -- is a key science driver for the observational community, while modelling \\lya{} radiative transfer in large EW emitters is a challenge for theorists. The main goal of this paper is to draw attention to the fact that \\lya{} EW is frequently discussed independently from the quantity SFR(\\lya{})/SFR(UV) \\citep[e.g.][]{Ajiki03,Fujita03,Venemans04,Shima06,Gronwall07,Tapken07,P08}. This quantity denotes the ratio of the star formation rates derived from the observed \\lya{} flux and rest-frame UV flux density. One can consider the quantity SFR(\\lya{})/SFR(UV) as an ``alternative'' measurement of EW, because the quantities EW and SFR(\\lya{})/SFR(UV) are directly related (see \\S~\\ref{sec:basis} of this paper, and e.g. \\citealp{DW07}, \\citealp{Rauch08}, \\citealp{Dayal08}, \\citealp{Nagamine08}, \\citealp{Nilsson09}). Since EW represents a fundamental property of \\lya{} emitting galaxies, a more detailed investigation of its relation to the ratio SFR(\\lya{})/SFR(UV) is warranted. We derive the relation between EW and SFR(\\lya{})/SFR(UV), and compare with observations in \\S~\\ref{sec:result}. We discuss the implications of our results in \\S~\\ref{sec:conc}. Throughout this paper we denote the rest frame equivalent width by REW, and the observed equivalent width by OEW. The two are related by REW=OEW$/(1+z)$. When we write ``EW'' this refers to both OEW and REW. ", "conclusions": "\\label{sec:conc} We investigate the relation between REW and the ratio of star formation rates derived from \\lya{} flux, and rest-frame UV flux density [SFR(\\lya{})/SFR(UV)] (Eq~\\ref{eq:ratio2}). This relation derives directly from the definition of equivalent width and star formation rate conversion factors, and its only source of scatter is the variation in the slope of the UV continuum at $\\lambda_{{\\rm Ly}\\alpha}< \\lambda <\\lambda_{\\rm UV}$ between individual galaxies. The correlation exists regardless of the assumed star formation calibrators (which themselves depend on the assumed IMF and gas metallicity), or the true star formation rates of these galaxies. Despite their fundamentally tight relation, \\lya{} REW and SFR(\\lya{})/SFR(UV) are often discussed as independent quantities. We investigate their correlation in existing data, and find the vast majority of galaxies to be consistent with the predicted relation (see Figs~\\ref{fig:corr1}). The existence of the relation has interesting applications, which are discussed next. \\subsection{An Empirically Constrained Ly$\\alpha$ Based Star Formation Indicator} A SFR derived from the UV flux density is likely more reliable than a SFR derived from Ly$\\alpha$ flux. Ly$\\alpha$ scatters through the ISM and IGM which makes it hard to determine the amount of extinction. Our relation can be used to derive a more accurate Ly$\\alpha$ based star formation calibrator, if we require that --statistically-- the Ly$\\alpha$ derived SFR should be equal to the UV derived SFR. We introduce the constant $\\mathcal{M}$ such that \\begin{equation} {\\rm SFR}_{\\mathcal{M}}({\\rm Ly}\\alpha)\\equiv \\mathcal{M}\\times {\\rm SFR(Ly}\\alpha{\\rm )}\\equiv {\\rm SFR(UV)}. \\label{eq:m} \\end{equation} The constant $\\mathcal{M}$ ensures that the SFR derived for a certain galaxy from its measured Ly$\\alpha$ flux is equal to that derived from its UV flux density. According to Eq~\\ref{eq:ratio2} $\\mathcal{M}=[C({\\rm REW}/{\\rm REW}_{\\rm c})]^{-1}$; this implies that a galaxy with an unusually large REW has $\\mathcal{M}\\ll 1$. Without this correction one would overestimate the SFR from the Ly$\\alpha$ flux alone. In the most general case, the probability $P(\\mathcal{M})d\\mathcal{M}$ that $\\mathcal{M}$ lies in the range $\\mathcal{M}\\pm d\\mathcal{M}/2$ is given by (see Appendix~\\ref{app:m}) \\begin{eqnarray} P(\\mathcal{M})d\\mathcal{M}=d\\mathcal{M}\\hs\\mathcal{N}\\int_{-\\infty}^{\\infty} P(\\beta) P({\\rm REW}_{\\mathcal M})\\frac{{\\rm REW}_{\\mathcal M}(\\beta)}{\\mathcal{M}}d\\beta, \\label{eq:pm} \\end{eqnarray} with $\\mathcal{N}$ the normalization factor which ensures that $\\int_{0}^{\\infty} P(\\mathcal{M})d\\mathcal{M}=1$, and REW$_{\\mathcal{M}}(\\beta)\\equiv {\\rm REW}_{\\rm c}/(\\mathcal{M}C)$. Furthermore, $P(\\mathrm{REW})d\\mathrm{REW}$ denotes the probability that a LAE has an observed REW in the range REW$\\pm d$REW/2; $P(\\beta)d\\beta$ denotes the probability that a LAE has an observed $\\beta$ in the range $\\beta \\pm d\\beta/2$. Figure~\\ref{fig:pdf} shows the probability distribution $P(\\mathcal{M})$ obtained from Eq~\\ref{eq:pm} ({\\it black solid line}). In this calculation we assume: ({\\it i}) $P(\\beta)d\\beta$ is a Gaussian with $\\bar{\\beta}=1.2$, and $\\sigma_{\\beta}=0.9$. This choice ensures that $\\sim 90\\%$ of the galaxies have $0 \\la \\beta \\la 2.4$ \\citep[cf.][]{Tapken07}; ({\\it ii}) $P(\\mathrm{REW})d\\mathrm{REW}$ is an exponential with a scale length of REW$_{\\rm L}$=76 \\AA\\hs\\citep[][]{Gronwall07}, based on observed Ly$\\alpha$ emitting galaxies at $z=3.1$ with REW$> 20$ \\AA, i.e $P$(REW)dREW$\\propto$ exp(-REW/REW$_{\\rm L})$ for REW$>$REW$_{\\rm min}\\sim$ 20 \\AA, and P(REW)$=0$ otherwise. In Appendix~\\ref{app:dos} we show that the precise shape of the distribution is not very sensitive to the assumed probability density functions (PDFs) of $\\beta$ and REW. \\begin{figure} \\vbox{\\centerline{\\epsfig{file=fig3.eps,angle=270,width=8.0truecm}}} \\caption[]{The {\\it black solid line} shows the probability distribution for $\\mathcal{M}\\equiv$SFR(UV)/SFR(Ly$\\alpha$) as given by Eq~\\ref{eq:ratio2}, in which $P$(REW) and $P$($\\beta$) are determined from the observations (see text). The {\\it red dotted lines} mark the $68\\%$ confidence interval centered on the median $\\mathcal{M}=1.13$ (indicated by the {\\it blue dashed line}). We thus find that$(68\\%$ of LAEs have $\\mathcal{M}=1.1^{+1.4}_{-0.7}$. The data points denote the directly measured distribution of SFR(UV)/SFR(Ly$\\alpha$) collected from the papers used in our analysis. The errorbars denote Poisson uncertainties in these data. The observed data distribution agrees well with our derived PDF. Furthermore, the derived PDF is reproduced quite well by a log-normal distribution which is indicated by the {\\it grey dot-dashed line}, which is easily adopted when modeling statistical properties of LAEs.} \\label{fig:pdf} \\end{figure} Figure~\\ref{fig:pdf} shows that $P(\\mathcal{M})$ peaks at $\\mathcal{M}\\sim 0.6$. The function $P(\\mathcal{M})$ is highly asymmetric and its median is $\\mathcal{M}=1.13$, which is indicated as the {\\it blue dashed line}. That is, SFR(Ly$\\alpha$) $<0.88$ SFR(UV) for $50\\%$ of the galaxies (similarly, SFR(Ly$\\alpha$) $<$ SFR(UV) for $\\sim 57\\%$ of the galaxies). The {\\it red dotted lines} mark the $68\\%$ confidence interval centered on the median value. We therefore find that for $68\\%$ of LAEs $\\mathcal{M}=1.1^{+1.4}_{-0.7}$. In other words, the Ly$\\alpha$ derived SFR lies within a factor of $\\sim 2.5$ from the UV derived SFR for 68\\% of LAEs. Overplotted as the data points is the measured distribution of SFR(UV)/SFR(Ly$\\alpha$) collected from the papers that were used in our analysis. The errorbars denote Poisson uncertainties. This observed distribution agrees quite well with our derived PDF. The {\\it grey dot-dashed line} indicates a log-normal distribution \\begin{eqnarray} P(\\mathcal{M})d\\mathcal{M}=\\frac{1}{\\sigma\\sqrt{2\\pi}}{\\rm exp}\\Big{[} -\\frac{1}{2}\\Big{(}\\frac{[\\log \\mathcal{M}]-x}{\\sigma}\\Big{)}^2\\Big{]}\\frac{d\\mathcal{M}}{\\mathcal{M}\\ln 10}, \\label{eq:log} \\end{eqnarray} with $x=0.04$ and $\\sigma=0.35$. This log-normal distribution provides a decent fit to the derived and observed distribution, and is easily adopted when modeling statistical properties of LAEs. \\citet{Ouchi08} conclude that between $z=3$ and $z=6$ the observed REW PDF is consistent with no redshift evolution. This implies that our derived PDF for $\\mathcal{M}$ also applies at redshifts greater than $z=3$. On the other hand, the measured scalelength of the exponential REW distribution at $z=2.3$ is REW$_{\\rm L}=48$ \\AA\\hs\\citep{Nilsson09}, causing the PDF to broaden and shift to larger $\\mathcal{M}$ (see Appendix \\ref{app:dos}). Table~\\ref{table:param} summarizes the redshift dependence of the parameters describing the log-normal distribution for $\\mathcal{M}$. In both redshift bins the standard deviation of the log-normal distribution is very similar, with $\\sigma \\sim 0.3-0.35$. The abrupt change of the value of $x$ at $z=3$ is clearly a crude approximation of the real redshift evolution of the $\\mathcal{M}$-PDF. We have chosen this parametrization because it corresponds to the simplest description that is consistent with existing data. \\begin{table} \\begin{minipage}{8cm} \\centering \\caption{Fit Parameters for the log-normal PDF for $P(\\log \\mathcal{M})d\\mathcal{M}$.} \\begin{tabular}{l c c} \\hline\\hline & x & $\\sigma$\\\\ $z \\gsim 3$\\hs\\footnote{We assume that the REW-PDF, and hence the $\\mathcal{M}$-PDF, does not to evolve between $z=3-6$ (Ouchi et al. 2008). We take the observed $z=3$ REW-PDF from Gronwall et al. (2007) for $z\\gsim 3$, and the observed $z=2.3$ REW-PDF from Nilsson et al. (2009) for $z\\lsim3$.} & 0.04 & 0.35 \\\\ $z \\lsim 3$ & 0.20& 0.31 \\\\ \\hline\\hline \\end{tabular} \\label{table:param} \\end{minipage} \\end{table} In theoretical work the inverse problem often arises in which a Ly$\\alpha$ luminosity must be obtained from a physical SFR. In such a case one may write $L_{{\\rm Ly}\\alpha}=\\frac{1.1\\times 10^{42}\\hs{\\rm erg}\\hs{\\rm s}^{-1}}{\\mathcal{M}}\\Big{(}\\frac{{\\rm SFR}}{M_{\\odot}\\hs{\\rm yr}^{-1}}\\Big{)}$ and adopt our PDF for $\\mathcal{M}$. This allows one to assign an empirically calibrated, variable Ly$\\alpha$ luminosity to galaxies of a given SFR. This prescription is clearly not perfect, because not all galaxies that are actively forming stars show a Ly$\\alpha$ emission line. For example, only 20-25$\\%$ of Lyman Break galaxies (LBGs) at $z=3$ have a Ly$\\alpha$ line with REW$\\gsim 20\\%$ \\AA\\hs and would classify as a LAE \\citep[e.g.][]{Shapley03}, and the connection between LBGs and LAEs is not well understood. Nevertheless, our suggested prescription for assigning Ly$\\alpha$ luminosities to galaxies of a given SFR is more realistic than the often used one-to-one relation $L_{{\\rm Ly}\\alpha}=1.1 \\times 10^{42}\\hs{\\rm erg}\\hs{\\rm s}^{-1}\\hs {\\rm SFR}/(M_{\\odot}\\hs{\\rm yr}^{-1})$. \\subsection{Outliers in the SFR(Ly$\\alpha$)/SFR(UV)-REW Plane: Signposts for Unusual Galaxies?} ``Normal'' star forming galaxies can emit a maximum REW$_{\\rm max}=240$\\,\\AA. Galaxies with REW$>$REW$_{\\rm max}$ may signal the presence of a galaxy that contains population III stars \\citep[e.g.][]{MR02}. Using Eq~\\ref{eq:ratio2} this corresponds to [SFR(\\lya{})/SFR(UV)]$_{\\rm max} =2.8^{+0.4}_{-0.5}$. The majority of observed galaxies that have quoted uncertainties on SFR(\\lya{})/SFR(UV) are consistent with this upper limit. Objects where nebular emission dominates, such as galaxies containing population III stars \\citep{S02} or cooling clouds \\citep{D09}, may be dominated by the two-photon continuum at 1216\\,\\AA{} $< \\lambda <$ 1600\\,\\AA{}. This results in unusually negative values for $\\beta$. These objects may have been identified in the spectrum itself, because deviations from the correlation are caused by unusual spectral slopes. However, reliable measurements of the continuum just redward of the Ly$\\alpha$ line and at $\\lambda_{\\rm UV}=1400$ \\AA\\hs provide a long baseline in wavelength, which may more clearly reveal the presence of a continuum dominated by two-photon emission. This suggests that outliers in the REW - SFR(Ly$\\alpha$)/SFR(UV) plane may provide a more sensitive probe to cooling clouds or primordial galaxies than the spectrum alone. That additional probe is important especially at high redshifts, where the IGM may transmit only a small fraction of the \\lya{} emitted by galaxies \\citep{igm}. In this case even those star forming galaxies containing population III stars may have REW$<$REW$_{\\rm max}$. This strongly suggests that the determined REW alone is not enough to identify a population III galaxy. {\\bf Acknowledgements} M.D. is supported by Harvard University funds. E.W. acknowledges the Smithsonian Institution for the support of his postdoctoral fellowship. We thank an anonymous referee for helpful, constructive comments that improved the content of this paper." }, "0911/0911.2244.txt": { "abstract": " ", "introduction": "In the early nineties, cosmic shear was predicted to be a promising way to study the distribution of matter in the Universe \\cite{1991MNRAS.251..600B,1992ApJ...388..272K,1991ApJ...380....1M} and since its first detection \\cite{2000MNRAS.318..625B,2000astro.ph..3338K,2000A&A...358...30V,2000Natur.405..143W} it has shown to be a precious mean of investigations of the large-scale structure of the Universe, enabling us to explore dark energy properties or uncover signatures of mode coupling effects \\cite{2008ARNPS..58...99H,1999ARA&A..37..127M,2008PhR...462...67M}. So far cosmic shear surveys have covered only a limited field in the sky. For instance, the CFHTLS,\\footnote{http://www.cfht.hawaii.edu/Science/CFHTLS} which has produced very promising results over the last years (see e.g.~\\cite{2008A&A...479....9F} for a recent account of these observations), is limited to about a 170 squared degree range. With the demonstration of the robustness of cosmic shear observations, (nearly) full-sky surveys such as Pan-STARRS, DES, LSST, JDEM, or Euclid are under preparation. They will open the way to new types of studies. Akin to CMB observations, such surveys will be an excellent tool to explore the physics of the Universe at scales comparable to the Hubble radius, therefore testing genuinely general relativistic effects.\\footnote{See \\cite{2009arXiv0907.0707Y} for a recent account of these effects on galaxy clustering observations.} In particular, the study of mode couplings, already well established on Newtonian scales, can be extended at these very large scales therefore testing the details of our understanding of the origin and formation of the large-scale structure. Such an investigation requires that we know the types of mode couplings that are expected to be seen at such large scales. Calculations have been undertaken to predict the nonlinear growth of metric and density fluctuations after modes reenter the Hubble radius \\cite{Bartolo:2005kv,Boubekeur:2008kn,Fitzpatrick:2009ci}. In the context of the CMB anisotropies, progress has been recently made in understanding the effect of these nonlinearities, from concentrating on the large angular scales \\cite{Pyne:1995bs,Bartolo:2004ty,2009JCAP...08..029B} to the details of the physics of recombination (see for instance \\cite{2006JCAP...06..024B,2009CQGra..26f5006P}). So far the investigations of mode couplings in weak lensing were limited to small angular scales, corresponding to scales much smaller than the angular diameter distance at the source. Accidentally, this distance roughly corresponds to the Hubble radius at the source. Thus, on these scales one can consistently neglect general relativistic effects that are suppressed by the ratio between the scale probed and the Hubble scale. On small angular scales the dominant contribution to the cosmic shear comes from fluctuations of the gravitational potential transverse to the line of sight. Perturbations along the line of sight average out and do not yield appreciable effects. In this regime the dominant geometrical mode couplings were identified more than a decade ago in \\cite{1997A&A...322....1B}. They include the Born correction and the lens-lens coupling. In the so-called Born approximation one integrates the lensing distortion over an unperturbed photon path. One can consider the correction due to the fact that the photon path is perturbed. The lens-lens coupling consists in the correction due to the deformation of a distant lens caused by a foreground one. The consequences of these effects have been extensively described in the literature and they have been found to have an impact on both the shear power spectrum and higher-order statistical observables such as the bispectrum \\cite{Cooray:2002mj,2003MNRAS.344..857T,Dodelson:2005zj,2006JCAP...03..007S,Hilbert:2008kb,Krause:2009yr}. As the shape distortion probes the reduced shear rather than the shear itself, there is another correction associated to the nonlinear conversion between these two quantities \\cite{Dodelson:2005rf,Schneider:1997ge,Krause:2009yr}. Finally, another nonlinear effect is the source-lens clustering, due to the fact that the source of the lensed light is itself a perturbed field with specific clustering properties correlated with the lens \\cite{1998A&A...338..375B,2002A&A...389..729S}. For the current surveys restricted to a limited angular field all these types of couplings are undoubtedly the dominant ones. In view of full-sky surveys one needs to go beyond the small-angle approximation and probe scales of the order of the angular diameter distance to the source. In this case fluctuations along the line of sight are not negligible and terms other than those described above may become important. Furthermore, as we are probing scales comparable to the Hubble size, one needs to undertake a full general relativistic treatment. This is important in order to compute accessible higher-order observables in full or almost full-sky surveys. In particular, in such surveys, it becomes necessary if one wants to compute the lensing bispectrum in the squeezed configuration, when one of the scales probed is taken to be much larger than the other two. We present here the exhaustive calculation of the weak lensing cosmic shear at second order including all general relativistic contributions and without relying on the small-angle or thin-lens approximations. However, we do not include in our study the effect of source-lens clustering \\cite{1998A&A...338..375B,2002A&A...389..729S} and other intrinsic effects in the alignment and ellipticity of galaxies. We work in the so-called generalized Poisson gauge without specifying the matter content of the Universe. We assume that there are no primordial vector and tensor perturbations. However, we will take into account vector and tensor components of the metric generated at second order from scalar fluctuations. In this gauge we derive the reduced -- i.e.~observable -- shear by solving the Sachs equation \\cite{1961RSPSA.264..309S} which describes the distortion of the cross section of an infinitesimal bundle of light rays in the geometric-optics limit. The advantage of using the Sachs equation instead of the geodesic equation is that it deals only with physically observable quantities. As the resolution of the Sachs equation is extremely tedious and involves a large number of terms we will develop tests that allow us to check the validity of its solution. In particular, some of the contributions to the second-order shear that we compute -- and that are usually neglected in the small-angle approximation -- can be compared to those expected from the lensing shear at linear order in a universe with spatial curvature. Using the solution of the Sachs equation we will compute the reduced shear by adding the nonlinear corrections coming from the relation between this quantity and the shear itself. Finally, as in the Poisson gauge hypersurfaces of constant redshift are inhomogeneous, we take into account the corrections due to the inhomogeneity of the redshift of the source \\cite{2008PhRvD..78l3530B}. The final observable reduced shear field that we obtain is a gauge invariant quantity, although its separate contributions are not necessarily so. As it is customary for CMB polarization, we express the reduced shear in terms of spin-2 operators on the sky. In particular, angular gradients on the sky will be written in terms of spin raising and lowering operators, whose eigenfunctions are the well-known spin-weighted spherical harmonics. The plan of the paper is the following. In Sec.~\\ref{sec:weak} we give the outline of our calculation. In particular, we describe the Sachs equation and how the transverse size of a propagating beam can be related to the observable reduced shear including the effects of the source redshift inhomogeneities. We will solve the Sachs equation at first order in Sec.~\\ref{sec:linear} while the full second-order calculation will be presented in Sec.~\\ref{sec:shear_2nd}. We discuss and comment on our results in Sec.~\\ref{sec:conclusion} in the context of cosmic shear surveys regarding the generation of $B$ modes and the expected contributions to the cosmic shear bispectrum. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{sec:conclusion} In this article we have derived the expression of the reduced cosmic shear up to second order in the perturbations with full-sky validity. Our main result is summarized in eq.~(\\ref{final_result}). As it is expressed in terms of spin-2 operators on the sphere it can be decomposed as sum of spin-weighted spherical harmonics on the sky. Indeed, this description ensures that our observable has a genuine spin-2 behavior on the celestial sphere. Our result is written in terms of the metric perturbations in the generalized Poisson gauge. These are the scalar potentials $\\phi$ and $\\psi$ and the vector and tensor components of the metric generated at second order, respectively $\\omega_i$ and $h_{ij}$. Let us first comment on the first two terms on the right-hand side of eq.~(\\ref{final_result}). Remarkably, the contribution from scalar perturbations from the sum of these two terms can be expressed in terms of the Weyl potential $\\Psi = (\\phi+\\psi)/2$ only. As explained, this is due to the fact that null geodesics are conformally invariant. These two terms contain the well-known second-order corrections due to lens-lens coupling and departure from the Born approximation, which dominate in the small-angle approximation. On larger angular scales new couplings become important. These are an intrinsic contribution which is a purely general relativistic effect at second order, a coupling between the gravitational potential at the source with the lens and corrections due to couplings between the lens and the photon time-delay. We have checked that these contributions can be independently reconstructed from the calculation of the shear at first order in a universe with a radially dependent spatial curvature. Other checks, such as the invariance under a homogeneous time shift and a conformal transformation can be used to verify the validity of these new corrections. Another scalar correction appears in the form of products of two spin-1 fields and comes from the couplings between two photon deflections. Finally, besides the scalar contributions, the shear gets a contribution from spin-2 quantities defined from the vector and tensor components of the metric generated at second-order. Note that the separation between all these contributions is not gauge invariant. In Poisson gauge, the correction due to the coupling between the photon redshift perturbation and the lens cannot be written in terms of $\\Psi$ only. Indeed, the integrated contribution to the photon redshift -- the integrated Sachs-Wolfe effect -- is a time integral over $\\dot \\Psi$ but the intrinsic contributions -- Sachs-Wolfe and Doppler effects -- are expressed in terms of the Newtonian gravitational potential and the velocity along the line of sight and do not depend on $\\Psi$ only. We are now in the position to explore the phenomenological consequences of these results in view of the future (partially) full-sky lensing surveys. In particular, the new corrections that we have computed should become relevant in deriving the lensing bispectrum on large angular scales. For instance, to compute the bispectrum in the squeezed limit one needs to take one of the three modes to be much smaller than the other two, corresponding to angular scales comparable to the depth of the survey. As the lensing is a cumulative effect integrated along the line of sight, it is difficult, at this stage, to precisely guess the relative importance of the various contributions. In particular, although the lens-lens coupling terms are a priori larger by a factor $\\sim l^2$ compared to the others, they may be damped by geometrical factors. We leave these investigations for the future." }, "0911/0911.4955_arXiv.txt": { "abstract": "We constrain the cosmological density of cosmic string loops using two observational signatures -- gravitational microlensing and the Kaiser-Stebbins effect. Photometry from RXTE and CoRoT space missions and pulsar timing from Parkes Pulsar Timing Array, Arecibo and Green Bank radio telescopes allow us to probe cosmic strings in a wide range of tensions $G\\mu/c^2=10^{-16}\\div10^{-10}$. We find that pulsar timing data provide the most stringent constraints on the abundance of light strings at the level $\\Omega_s \\sim 10^{-3}$. Future observational facilities such as the Square Kilometer Array will allow one to improve these constraints by orders of magnitude. ", "introduction": "} Cosmic strings are now a widely recognized part of cosmological theory. Cosmic stings appear naturally in a multitude of inflationary models as topological defects from the early Universe (e.g. \\cite{Allen1990,Copeland2009}, for more studies see references in \\cite{Chernoff2009}). Similar objects commonly referred to as cosmic superstrings can also be produced in fundametal string and M-theories \\cite{daviskibble, polchinski}. In the present study we will not differentiate between the two classes, because their observational signatures considered in this paper are the same. The key parameter of a cosmic string is its tension $\\mu$, which is assumed to be related to the effective energy scale of the string-producing theory $\\Lambda$ by \\cite{Vilenkin1994} $$ \\frac{G\\mu}{c^2}\\sim\\frac{\\Lambda^2}{M_\\mathrm{Planck}^2}. $$ Earliest theories of string formation placed time of their generation to the Grand Unification Theory epoch and therefore their tension seemed to be of order $10^{-6}$. Initially, possible tensions of string were constrained from both sides: $10^{-11} 10^{-13}$ because of lesser enhancement of density of heavy strings. These results clearly demonstrate that the density enhancement in the Galaxy improves the limits on cosmic string abundance that can be set by test sensitive to the local population of loops as suggested by~\\cite{Chernoff2009}. However, the same approach can also be applied to extragalactic sources, most notably quasars. The probability of lensing by string loops for these sources is lower due to absence of enhancement in the extragalactic case. On the other hand, the huge distance to the source can easily compensate for the relative deficit of string density. Moreover, these observations are insensitive to theoretical uncertainty in predicting $\\eta$. As an example we consider the 'Einstein cross' quasar QSO~2237+0305. The Optical Gravitational Lensing Experiment ({\\it OGLE}) data for this source currently span more than 7 years of high-quality photometry sampled every few days~\\cite{ogle2237} and no obvious signature of string lensing are seen in these light curves\\footnote{The light curve for image~B does show an approximately twofold increase in brightness starting near $JD\\sim2453500$ but present data is far from being sufficient to draw any firm conclusions. Unfortunately, no observations can resolve this source at present to the level needed to rule out the possibility of lensing by cosmic string.}. This non-detection can be used to place some limits on the density of cosmic strings in the range of tensions $3\\cdot10^{-12}10\\,\\msun$ can be found in our models as far as $10\\,\\pc$ from the cluster center. However, we find that in order to measure the correct slope of the MF, we only need the stars inside the tidal radius of $1\\,\\pc$. No turn-over at the low mass end of the IMF can be seen. This means that the turn-over seen by \\citet{SBG05} cannot be explained by mass segregation (unlike the flat IMF at the high-mass end). Deriving the IMF from our simulation data is not hampered by incompleteness, selection effects, and mass determination from observed luminosities. All these effects may bias the observed star counts to produce a turn-over, however, the turn-over appears at $\\sim6\\,\\msun$ where the data is still 50\\% complete. So, either the IMF is indeed truncated or another effect has to be considered. One possible explanation is that tidal stripping preferably removes low-mass stars from the cluster. This effect is not included in our current simulations but will be tested in a follow-up paper. Alternatively, \\citet{ESM09} pointed out that local variations in extinction can account for (some of) the flattening of the obserevd MF. The same effect could cause lower-mass stars to be lost preferentially. Then the completeness function, which only takes into account crowding and sensitivity effect, would not be sufficient to correct the low-mass end of the IMF. \\begin{figure} \\includegraphics[width=84mm,angle=0]{plots/rcore.eps} \\caption{The core radius evolution with time of model IKW03S05 with $\\rvir=0.7\\,\\pc$ and $\\NMS=150$. The black line with errorbars shows the average of ten random realizations of the same model (grey lines). Core collapse happens at $\\sim4\\,\\myr$. } \\label{fig:rcore} \\end{figure} In Fig.~\\ref{fig:rcore} we show the core radius evolution with time for one of our best fit models. At its current age of $3.5\\,\\myr$ the cluster is a little more than halfway to core collapse which will occur at $\\sim4\\,\\myr$. This result is in agreement with the findings of \\citet{PZGC07}. We have performed a large number of $N$-body simulations in order to find the best fitting model for the Arches cluster. The available observational data has been used to constrain the free parameters in our model. In a systematic analysis, we compared the total mass, the cumulative mass profile, and the present-day MF and defined fitness parameters for each of the three observables. The main conclusion from our analysis is that the Arches cluster, despite of being born in an extreme environment, has formed with a standard Salpeter IMF. Due to mass segregation, the slope of the observed MF is flattened inside a radius of $0.4\\,\\pc$. This radius was imposed a limit from the obversavational data used, and we estimate that a limiting radius of $\\sim1\\,\\pc$ would be required for the observed MF to match the underlying IMF. We conclude from our best fitting models that the Arches cluster was born with an initial virial radius of $0.76 \\pm 0.12\\,\\pc$ and an initial total mass of about $(5.1 \\pm 0.8) \\cdot 10^4\\,\\msun$ assuming a lower mass limit of $0.5\\,\\msun$ for the Salpeter IMF. The lower mass limit cannot be constrained well from our models, giving rise to additional uncertainties in determining the initial cluster mass. The King model concentration parameter of the best fitting models is $W_0 = 3-5$, however, the $W_0 = 7$ produced also reasonable fits so that this parameter is not very well constrained. We neglected the Galactic potential and also stellar evolution for the simulations in this paper. These processes will be included in a following paper to get a more realistic model of the Arches cluster, however this first step was needed in order to reduce the number of free parameters for the these (computationally more expensive) simulations, which further constrain the dynamical evolution and tidal effects acting on the Arches cluster, and thereby the initial conditions of this nearby starburst, such as the cluster mass, the orbital motion and the IMF at the birth of the cluster. {\\bf Acknowledgments} SH and SPZ are grateful for the support from the NWO Computational Science STARE project \\#643.200.503 and NWO grant \\#639.073.803. AS acknowledges funding from the German Science Foundation (DFG) Emmy-Noether-Programme under grant \\mbox{STO 496-3/1}." }, "0911/0911.1788_arXiv.txt": { "abstract": "This is the fourth of a series of papers in which we derive simultaneous constraints on cosmological parameters and X-ray scaling relations using observations of the growth of massive, X-ray flux-selected galaxy clusters. Our data set consists of 238 clusters drawn from the \\ROSAT{} All-Sky Survey, and incorporates extensive follow-up observations using the \\Chandra{} X-ray Observatory. Here we examine the constraints on neutrino properties that are enabled by the precise and robust constraint on the amplitude of the matter power spectrum at low redshift available from our data. In combination with cluster gas mass fraction, cosmic microwave background, supernova and baryon acoustic oscillation data, and incorporating conservative allowances for systematic uncertainties, we limit the species-summed neutrino mass, \\Mnu{}, to $<0.33\\eV$ at 95.4 per cent confidence in a spatially flat, cosmological constant (\\LCDM) model. In a flat \\LCDM{} model where the effective number of neutrino species, \\Neff{}, is allowed to vary, we find $\\Neff=3.4_{-0.5}^{+0.6}$ (68.3 per cent confidence, incorporating a direct constraint on the Hubble parameter from Cepheid and supernova data). We also obtain results with additional degrees of freedom in the cosmological model, in the form of global spatial curvature (\\Omegak) and a primordial spectrum of tensor perturbations ($r$ and \\nt{}). The results are not immune to these generalizations; however, in the most general case we consider, in which \\Mnu{}, \\Neff{}, curvature and tensors are all free, we still obtain $\\Mnu<0.70\\eV$ and $\\Neff=3.7 \\pm 0.7$ (at, respectively, the same confidence levels as above). These results agree well with recent work using independent data, and highlight the importance of measuring cosmic structure and expansion at low as well as high ($z \\sim 1100$) redshifts. Although our cluster data extend to redshift $z=0.5$, the direct effect of neutrino mass on the growth of structure at late times is not yet detected at a significant level. ", "introduction": "\\label{sec:introduction} Observations of neutrino flavor oscillation have conclusively shown that the neutrino mass eigenstates are non-degenerate \\citep[e.g.][]{Fukuda98,Fukuda02,Ahmad02,Ahn03,Ahn06,Eguchi03,Sanchez03,Giacomelli04,Aharmim05}. Although, these observations can place tight constraints on the differences in squared mass, measuring the absolute mass scale remains challenging. Current laboratory efforts focus on tritium beta decay \\citep[e.g.][]{Lobashev03,Kraus05} and neutrinoless double beta decay \\citep[e.g.][]{Aalseth99,Klapdor01,Arnaboldi05,Arnold05}; the latter, if observed, would additionally indicate that neutrinos are Majorana rather than Dirac fermions. The squared mass differences measured from flavor oscillations place a lower bound on the sum of the three masses, $\\Mnu=\\sum_i m_i$, at $\\sim 0.056$ (0.095)$\\eV/c^2$ in the normal (inverted) hierarchy, while current tritium beta decay results provide an upper bound on the mass of the electron neutrino at $\\sim 2\\eV/c^2$ (thus on \\Mnu{} at $\\sim 6\\eV/c^2$).\\footnote{Henceforth, we set $c=1$.} Because neutrinos play a prominent role in the early Universe, cosmological observations are also sensitive to their properties (for a review, see \\citealt{Lesgourgues06}; see also \\secref{sec:background}). The primary effect of non-zero neutrino mass on cosmological observables is to suppress the formation of cosmic structure on intermediate and small scales. Comparison of the Cosmic Microwave Background (CMB), which reflects large-scale structure at early times, with measurements of the intermediate- or small-scale structure in the local Universe thus provides a way to constrain the absolute mass scale of the neutrino \\citep[e.g.][]{Fukugita00}. This approach has the disadvantage that any neutrino properties inferred are at some level sensitive to our incomplete understanding of cosmology, in particular dark energy and inflation. It has long been recognized, however, that using complementary measurements of structure, as described above, significantly reduces the sensitivity of the results to such necessary assumptions (e.g. \\citealt*{Allen03a}; \\citealt{Tegmark04}). Nevertheless, it is imperative that any systematic uncertainties affecting the measurements of cosmic structure be properly accounted for in order to obtain robust results. The amount of structure in the local Universe is generally described by the parameter $\\sigma_8$, defined as the present day root-mean-square fluctuation of the linearly evolved matter density field, smoothed by a spherical top-hat window of comoving radius $8h^{-1}\\Mpc$; here $h=H_0/100\\km\\second^{-1}\\Mpc^{-1}$ is the normalized Hubble parameter. At present, the most robust observation for measuring this quantity is arguably the abundance of massive galaxy clusters;\\footnote{We note that current cluster measurements do not constrain $\\sigma_8$ independent of spectral index of the power spectrum, \\ns{}. However, the best fitting value of $\\sigma_8$ does not vary rapidly with \\ns{}; moreover, \\ns{} is well constrained by CMB data in all cosmological models considered here.} recent advances in cluster simulation \\citep*[e.g.][]{Nagai07}, comparisons of different mass measurement techniques \\citep[e.g.][]{Bradac08,Mahdavi08,Newman09}, and the availability of robust mass proxies (e.g. \\citealt{Allen04,Allen08}; \\citealt*{Kravtsov06}) have significantly reduced systematic uncertainties associated with the measurement of cluster masses. Consequently, recent estimates of $\\sigma_8$ based on independent analyses of galaxy cluster data (both optically and X-ray selected clusters) are in very good agreement [$\\sigma_8 \\sim 0.8$; \\citealt{Mantz08,Mantz09} (hereafter \\cosmopaper{}); \\citealt{Henry09,Vikhlinin09a,Rozo10}]. Recent analyses of cosmic shear data are also in good agreement \\citep{Benjamin07,Fu08}, and provide compatible constraints on the neutrino mass to our own \\citep{Tereno09}. Galaxy redshift surveys have also produced comparable results \\citep*{Thomas09}. The number of neutrino species participating in weak interactions is known to be three to high precision \\citep{Amsler08}. However, the possibility remains that additional ``sterile'' species exist. Results consistent with the existence of a fourth neutrino were reported by the Liquid Scintillator Neutrino Detector \\citep[LSND;][]{Aguilar01}; these results are disfavored by the MiniBooNE experiment \\citep{MiniBooNE09,MiniBooNE09a}, although the interpretation remains somewhat ambiguous at present. This is another question that cosmological observations can address. In particular, the synthesis of light elements is sensitive to the number of relativistic species present in the early Universe, since these determine the expansion rate at that time; however, observations of primordial deuterium and helium abundances are challenging and are subject to large systematic uncertainties. An independent probe is provided by the CMB, in combination with other cosmological data, as described in \\secref{sec:background}. In this work, we consider only the case where the neutrino species (whatever their number) have approximately degenerate mass; in this case, \\Mnu{} and the effective number of neutrino species, \\Neff{}, are sufficient to describe the cosmological effect of neutrinos. More general mass splittings, and in particular the case favored by the initial LSND results, in which a sterile neutrino is significantly more massive than the other species, require a more complete treatment \\citep[as in, e.g.,][]{Crotty04}. In this paper, we apply the statistically rigorous analysis method of \\cosmopaper{} and the X-ray flux-limited cluster samples and follow-up observations described in \\citet[][hereafter \\scalingpaper{}]{Mantz09a} to the problem of inferring neutrino properties from cosmological data. These data (collectively referred to here as the cluster X-ray Luminosity Function, or XLF) provide a robust means to measure $\\sigma_8$; our analysis method includes generous systematic allowances and accounts fully for all parameter degeneracies. In obtaining our results, we also incorporate CMB data and measurements of cosmic distance in the form of cluster gas mass fractions (\\fgas{}), type Ia supernova (SNIa) fluxes and Baryon Acoustic Oscillation (BAO) data (see \\secref{sec:data} and references therein). We note that \\citet{Reid09a} obtain very similar results to our own by importance sampling 5-year {\\it Wilkinson Microwave Anisotropy Probe} (\\WMAP{}) results using a prior based on the analysis of optically selected clusters by \\citet{Rozo10}. The basic cosmology that we consider in this paper is the spatially flat, cosmological constant (\\LCDM{}) model parametrized by the mean baryon density, \\Omegab{}; the mean total matter density including baryons, neutrinos and cold dark matter (CDM), \\Omegam{}; the Hubble parameter, $H_0$; the normalization of the matter power spectrum, $\\sigma_8$; the spectral index of the primordial scalar power spectrum, \\ns{}; and the optical depth to reionization, $\\tau$. Here the mean densities refer to the present day (redshift $z=0$), since their values at other times are then determined by the Friedmann equation, and $\\sigma_8^2$ is the $z=0$ variance in the matter density field at scales of $8h^{-1}\\Mpc$, as defined above. This simple and commonly used model assumes $\\Mnu=0$ and $\\Neff=3.046$, the predicted value for the three weakly interacting neutrinos. In addition to freeing \\Mnu{} and \\Neff{}, we will consider the effect of marginalizing over other parameters with which they are degenerate. However, since the flat \\LCDM{} model is known to provide a good fit to currently available cosmological data, we are conservative in incorporating these additional degrees of freedom. In particular, we consider generalizing the description of dark energy, either by allowing global spatial curvature or by varying the dark energy equation of state in a flat universe, and including the effects of primordial tensor perturbations. The former case is parameterized by the effective curvature energy density, \\Omegak{}, or the equation of state parameter, $w$, while in the latter case we simultaneously marginalize over the tensor-to-scalar ratio, $r$ (defined with respect to wavenumber $k=0.002h\\Mpc^{-1}$, as in, e.g., \\citealt{Komatsu09}), and the tensor power spectral index, \\nt{}. In our results, we will consistently quote one-sided limits (upper bounds) on \\Mnu{} and $r$ at the 95.4 per cent confidence level and two-sided constraints on all other parameters at 68.3 per cent confidence. ", "conclusions": "\\label{sec:conclusion} We have applied measurements of the growth of massive galaxy clusters (detailed in Papers~I and II), in combination with other cosmological data, to the problem of constraining the species-summed neutrino mass, \\Mnu{}, and effective number, \\Neff{}. Our results show that a robust measurement of $\\sigma_8$ from clusters significantly improves limits on \\Mnu{} from current data, and reduces their sensitivity to assumptions about the cosmological model. In a simple \\LCDM+\\Mnu{} cosmology, the addition of the XLF data improves the 95.4 per cent confidence upper limit on \\Mnu{} to 0.33\\eV{}, compared with 0.61\\eV{} from the combination of CMB, \\fgas, SNIa and BAO data. In a more general model, marginalized over spatial curvature, primordial tensors and the effective number of neutrinos, and incorporating a prior on the Hubble parameter, the XLF data improve the limit from 1.75\\eV{} to 0.7\\eV{}. The results indicate that this improvement is due entirely to the ability of the cluster data to constrain $\\sigma_8$ (at $z=0$); while measurements of the ($z$-dependent) growth of structure in principle contain additional information about \\Mnu{}, current data are not sufficient to exploit it. The cluster data additionally improve constraints on the effective number of relativistic species, from $\\Neff=3.6_{-0.6}^{+0.7}$ to $\\Neff=3.4_{-0.5}^{+0.6}$ in a simple \\LCDM+\\Neff{} model (68.3 per cent confidence). For the more general model, including curvature, tensors and neutrino mass, only a marginal improvement is observed. The results obtained here are compatible with other recent estimates based on galaxy clusters \\citep{Reid09a,Vikhlinin09a}, reflecting the good agreement in recent $\\sigma_8$ constraints based on X-ray and optically selected clusters \\citep[\\cosmopaper{};][]{Henry09,Rozo10}. Although current cosmological data are not sufficient to rule out the existence of sterile neutrino species or distinguish between the normal and inverted mass hierarchies, they continue to provide some of the tightest constraints available, in particular on the mass scale. We also consider the prospects for further improvement, finding that a few per cent level measurement of the Hubble parameter will significantly improve the constraints on models where \\Neff{} is free. A similarly improved determination of $\\sigma_8$, combined with improved CMB data from {\\it Planck}, will further tighten limits on the neutrino mass scale,\\footnote{We note that the 7-year \\WMAP{} results, which were released while this work was in revision, already offer some improvement. Specifically, \\citet{Komatsu10} find $\\Mnu<0.58\\eV$ by combining 7-year \\WMAP{} data with the \\citet{Riess09} $H_0$ prior and the BAO results of \\citet{Percival10}, a noticeable improvement over the limit $\\Mnu<0.66\\eV$ obtained by \\citet{Sekiguchi09} from the same auxiliary data sets combined with WMAP5.} perhaps providing the first hints of non-zero mass from cosmological data. More precise measurements of the growth of structure at late times, and the extension of such data to higher redshifts, will provide a powerful, new mechanism to constrain neutrino mass." }, "0911/0911.5625_arXiv.txt": { "abstract": "Comet 17P/Holmes was observed for linear polarisation using the optical polarimeter mounted on the 1.2m telescope atop Gurushikhar peak near Mt. Abu during the period November-December 2007. Observations were conducted through the IHW narrow band (continuum) filters. During the observing run the phase angle was near $13^{\\circ}$ at which the comet showed negative polarisation. On the basis of the observed polarisation data we find comet 17P/Holmes to be a typical comet with usual dust characteristics. We note that radial rate of change of brightness in coma in red band is higher than that in blue band; it has decreased by a factor of 3.6 and 2.5 respectively in red and blue bands during the November - December run, indicating relative increase in the abundance of smaller dust particles out ward. Radial brightness variation seen near the nucleus on November 6 is indicative of the presence of a blob or shocked region beyond 10\" from the nucleus which has gradually smoothened by December 13. The brightness distribution is found steeper during November 5-7 as compared to on December 13. ", "introduction": "Photopolarimetric observations of comet 17P/Holmes were made during the period November 5-7 and on December 13, 2007 with a two channel photo polarimeter \\citep{Deshpande1985,Joshi1987}, which has been fully automated recently \\citep{ganesh2008p}, mounted on the 1.2m telescope of Mt. Abu Observatory operated by Physical Research Laboratory (PRL), Ahmedabad. {\\bf PRL polarimeter works on rapid modulation principle. A rotating super-achromatic half-wave plate in front of a fixed Wollaston prism modulates the polarized component of the incident light at a frequency of 10Hz. Another identical Panchratnam half-wave plate is put between the rotating half-wave plate and the Wollaston prism to eliminate any wavelength dependence of optic axis of half-wave plate. Rotating half plate rotates at discrete steps with 1 ms sampling time. One full rotation of half-wave plate is completed in 96 step and thus one modulation cycle involves 24 steps (i.e. 24ms) and the data are folded and accumulated in 24 bins. With the rapid modulation, the atmospheric scintillation effects (eg. sky transparency fluctuations) are eliminated. \\\\} The instrument is equipped with IAU's International Halley Watch ($IHW$)) continuum filters (3650/80\\AA, 4845/65\\AA, 6840/90\\AA~) \\citep{Osborn1990} and Johnson-Cousins' BVR broad band filters. The $IHW$ filters acquired for Comet Halley have been used for these observations. These filters have been in regular use for observations of several other comets \\citep{Joshi1987,Sen1991,Ganesh1998,Joshi2002, Joshi2003}. The filters have been carefully stored in dry atmospheric conditions to preserve their transmission characteristics. The observations made with the same set of filters facilitate better comparison with other comets observed earlier, hence their continued use is justified.\\\\ {\\bf The online data reduction performed after each integration invokes a least square fit to the counts (comet-sky) obtained from the two photo-multiplier tubes as a function of the rotating half-wave plate position. The mean error in polarisation is estimated from actual deviation of the counts from the fitted curve. To further reduce the error, several such measurements were taken and averaged. The error bars represent this error and instrumental error (obtained by observing several zero polarisation standards during each night and found to be $<0.03\\%$). Regarding the sky polarisation, on an average it was $\\sim 5-6\\% (\\pm 3\\%)$ but the sky as seen through the 26\" aperture is $\\sim 4$ mag fainter compared to the comet observed through the same aperture. These numbers were consistent during the observing run. Hence, the error communicated due to the sky to the comet data is negligible.} Nonetheless, to take care of the sky polarisation, observations were made alternately centred on the photo Centre of the comet and on a region of the sky more than 30 arcmin away from the comet (along the anti-tail direction). All the observations were made under dark sky conditions. The errors in the position angle are obtained using the equation 8.5.4 given by \\citet{Serkowski1974}.\\\\ The observations were taken with apertures (non-metallic diaphragms centred on the photo-centre of the comet) of different sizes- 10\", 20\", 26\" and 54\" with the projected diameter varying from 11750 to 69925 km (see Table \\ref{obstab}) to study the behaviour of the dust as a function of radial distance from the comet nucleus. All the observations were made under dark sky conditions, and the comet was much brighter than the sky resulting in negligible contribution by the sky to the observed degree of polarisation of the comet. Nonetheless, to take care of the sky polarisation, observations were made alternately centred on the photo Centre of the comet and on a region of the sky more than 30 arcmin away from the comet (along the anti-tail direction).\\\\ Polarisation standard 9 Gem was observed to calibrate the observed position angle. Comet's IHW magnitudes were obtained using the observations of solar type stars, namely HD29461, HD76151. Polarisation values, corrected position angle and IHW magnitudes in continuum bands are given in Table \\ref{obstab}. \\begin{table*} \\caption{Polarisation observations of comet 17P/Holmes. Listed entries are Julian date(JDT), Heliocentric range(r), Geocentric range($\\Delta$), phase angle($\\alpha$), filter, aperture(arcsec), Diameter(km), total integration time(IT sec), degree of polarisation($P\\%$), error in polarisation(${\\epsilon}_P\\%$), position angle($\\theta^\\circ$) in equatorial plane, magnitude at the time of observations. JDT=JD-2454400} \\begin{tabular}{|l|l|l|l|l|l|l|l|l|l|l|l|l|} \\hline Date&JDT& r&${\\Delta}$&${\\alpha}$& Filter(\\AA)& Ap(\")&Dia(km)& IT(sec)& $P\\%$ & ${\\epsilon}_p\\%$ &${\\theta^\\circ}$ & mag \\\\ \\hline Nov 5&10.32687 & 2.485602 & 1.620252 & 13.9 & 6840 & 10 & 11751 & 400 & -1.17 & 0.50 & 19 & 10.05\\\\ &10.33699 & 2.485642 & 1.620252 & 13.9 & 4845 & 10 & 11751 & 300 & -1.70 & 0.51 & 4 & 11.09\\\\ &10.34764 & 2.485684 & 1.620251 & 13.9 & 3650 & 10 & 11751 & 500 & -2.72 & 2.14 & 81 & 12.45\\\\ &10.35990 & 2.485732 & 1.620251 & 13.8 & 6840 & 26 & 30553 & 400 & -1.22 & 0.21 & 24 & 8.24\\\\ &10.36958 & 2.485770 & 1.620252 & 13.8 & 4845 & 26 & 30553 & 400 & -1.09 & 0.18 & 22 & 9.28\\\\ &10.38014 & 2.485791 & 1.620252 & 13.8 & 3650 & 20 & 23502 & 100 & -1.25 & 0.76 & 86 & 10.58\\\\ Nov 6&11.28874 & 2.489386 & 1.620319 & 13.6 & 6840 & 26 & 30554 & 500 & -1.60 & 0.27 & 14 & 8.66\\\\ &11.30002 & 2.489430 & 1.620321 & 13.6 & 4845 & 26 & 30554 & 500 & -0.95 & 0.19 & 20 & 9.72\\\\ &11.31650 & 2.489495 & 1.620323 & 13.6 & 3650 & 26 & 30554 & 900 & -1.51 & 0.70 & 32 & 10.98\\\\ &11.33392 & 2.489564 & 1.620327 & 13.6 & 3650 & 10 & 11751 & 400 & -2.60 & 2.15 & 15 & 12.56\\\\ &11.34823 & 2.489620 & 1.620330 & 13.6 & 4845 & 10 & 11751 & 800 & -0.96 & 0.43 & 22 & 11.25\\\\ &11.36806 & 2.489698 & 1.620334 & 13.6 & 6840 & 10 & 11751 & 700 & -1.10 & 0.54 & 43 & 10.16\\\\ &11.38504 & 2.489765 & 1.620338 & 13.6 & 6840 & 20 & 23503 & 600 & -1.49 & 0.29 & 20 & 9.01\\\\ &11.40166 & 2.489831 & 1.620343 & 13.6 & 4845 & 20 & 23503 & 400 & -1.36 & 0.27 & 20 & 10.05\\\\ &11.41631 & 2.489888 & 1.620348 & 13.6 & 6840 & 54 & 63460 & 400 & -1.09 & 0.14 & 13 & 7.21\\\\ Nov 7&12.30185 & 2.493380 & 1.620604 & 13.4 & 6840 & 26 & 30559 & 600 & -1.60 & 0.23 & 15 & 8.67\\\\ &12.32505 & 2.493472 & 1.620613 & 13.4 & 4845 & 26 & 30560 & 600 & -1.17 & 0.16 & 17 & 9.73\\\\ &12.33843 & 2.493524 & 1.620618 & 13.4 & 3650 & 26 & 30560 & 500 & -1.02 & 0.71 & 97 & 11.03\\\\ Dec 13&48.25186 & 2.639940 & 1.784706 & 12.9 & 6840 & 10 & 12944 & 600 & -3.96 & 1.77 & 97 & 11.69\\\\ &48.26690 & 2.640003 & 1.784838 & 12.9 & 4845 & 10 & 12944 & 600 & -1.39 & 0.76 & 96 & 12.16\\\\ &48.28473 & 2.640077 & 1.784996 & 12.9 & 3650 & 10 & 12946 & 600 & -2.88 & 3.46 &142 & 12.98\\\\ &48.29981 & 2.640141 & 1.785130 & 12.9 & 3650 & 26 & 33662 & 300 & -3.20 & 4.12 &131 & 12.72\\\\ &48.31233 & 2.640193 & 1.785241 & 12.9 & 3650 & 20 & 25895 & 300 & -3.52 & 3.55 &138 & 12.84\\\\ &48.32681 & 2.640254 & 1.785370 & 12.9 & 3650 & 54 & 69923 & 400 & -4.56 & 3.53 &107 & 12.49\\\\ &48.33570 & 2.640291 & 1.785449 & 12.9 & 4845 & 54 & 69926 & 300 & -1.79 & 0.88 &114 & 11.55\\\\ &48.34532 & 2.640331 & 1.785535 & 12.9 & 4845 & 26 & 33669 & 400 & -1.97 & 0.83 &110 & 11.92\\\\ &48.35515 & 2.640372 & 1.785623 & 12.9 & 4845 & 20 & 25901 & 400 & -1.60 & 0.79 & 64 & 12.02\\\\ \\hline \\end{tabular} \\label{obstab} \\end{table*} ", "conclusions": " \\begin{itemize} \\item Like other comets, 17P/Holmes shows negative polarisation at low phase angle (ie below $20^\\circ$). Based on the polarisation values, we believe that 17P/Holmes is not unusual comet as claimed by \\citet{rosenbush2009} and the dust seems to be of the same nature as found in other comets. The discrepancy could possibly be due to the gas emission contamination in the broad band observations.\\\\ \\item Though there is indication that the value of polarisation decreases towards longer wavelength, we infer that, within the errors, the wavelength dependence of polarisation as observed through different sized apertures is typical of comets.\\\\ \\item Radial distribution of brightness in the coma shows higher variation with time in the red wave band as compared to that in the blue wave band; it has decreased by a factor of 3.6 and 2.5 respectively in red and blue bands indicating increase in the relative abundance of smaller dust particles away from the nucleus.\\\\ \\item Radial brightness distribution shows the non-uniform distribution of material near the comet nucleus on November 6 which has gradually smoothened by December 13. Also the brightness distribution has become steeper from November 5-7 run to December 13 run. \\end{itemize}" }, "0911/0911.2469_arXiv.txt": { "abstract": "The horizontal branch (HB) morphology of globular clusters (GCs) is most strongly influenced by metallicity. The second parameter phenomenon, first described in the 1960's, acknowledges that metallicity alone is not enough to describe the HB morphology of all GCs. In particular, astronomers noticed that the outer Galactic halo contains GCs with redder HBs at a given metallicity than are found inside the Solar circle. Thus, at least a second parameter was required to characterize HB morphology. While the term `second parameter' has since come to be used in a broader context, its identity with respect to the original problem has not been conclusively determined. Here we analyze the median color difference between the HB and the red giant branch (RGB), hereafter denoted $\\dvi$, measured from Hubble Space Telescope (HST) Advanced Camera for Surveys (ACS) photometry of 60 GCs within $\\sim$20 kpc of the Galactic Center. Analysis of this homogeneous data set reveals that, after the influence of metallicity has been removed from the data, the correlation between $\\dvi$ and age is stronger than that of any other parameter considered. Expanding the sample to include HST ACS and Wide Field Planetary Camera 2 (WFPC2) photometry of the 6 most distant Galactic GCs lends additional support to the correlation between $\\dvi$ and age. This result is robust with respect to the adopted metallicity scale and the method of age determination, but must bear the caveat that high quality, detailed abundance information is not available for a significant fraction of the sample. Furthermore, when a subset of GCs with similar metallicities and ages are considered, a correlation between $\\dvi$ and central luminosity density is exposed. With respect to the existence of GCs with anomalously red HBs at a given metallicity, we conclude that age is the second parameter and central density is most likely the third. Important problems related to HB morphology in GCs, notably multi-modal distributions and faint blue tails, remain to be explained. ", "introduction": "} It has been clear for decades that metallicity is the most influential factor governing the HB morphologies of Galactic GCs: metallicity is the first parameter. The earliest photographic color-magnitude diagrams (CMDs) of GCs by, e.g.\\,\\citet{ar52,sa53}, revealed that HB stars in metal-rich GCs tend to lie on the red side of the RR Lyrae instability strip while HB stars in metal-poor GCs lie primarily on the blue side. As more and more CMDs were assembled, exceptions to this rule were uncovered. \\citet{sa60} noted that M~13 and M~22 display HB morphologies appropriate for metal-poor GCs, despite the fact both GCs were of intermediate metallicity, and suggested that a difference in age might be responsible. \\citet{sa67} noticed that the HB of NGC~7006 is redder than either M~13 or M~3, despite the fact that all three GCs appeared to have very similar metallicities. Such anomalies suggested the need for a second parameter that could account for differences in HB morphology not obviously caused by metallicity. \\citet{va65} summarized the problem, stating that metallicity is not sufficient to explain the extent of observed HB morphologies but that differences in age or He enrichment could explain the observed variations. Both of these suggestions remain valid up to the present time. \\citet{va67} analyzed the integrated colors of 49 GCs with $UBV$ photometry and concluded that `at least two parameters (one of which is metal abundance) are required to describe globular clusters.' The second parameter phenomenon took on a greater significance with the seminal work of \\citet{sz78}. Searle \\& Zinn recognized that GCs with unusually red HBs are relatively rare in the inner regions of the Galactic halo (Galactocentric radius, $\\rgc \\la 8$ kpc) but become increasingly common at greater $\\rgc$ (see their Figure 10). Searle \\& Zinn used this fact to argue that the inner halo was assembled early and in a fairly short time while the outer halo was assembled over an extended period of time. Current galaxy formation scenarios envision the outer regions of the Galaxy as the accumulated debris of the many accretion events that shaped the early evolution of the Galaxy. In that context, understanding the origin and existence of age and metallicity gradients in the Galactic GC population is as relevant now as it was at the time when Searle \\& Zinn first introduced their halo formation scenario. Subsequent efforts relating to the second parameter problem fall into two general categories: those that attempt to measure the age difference between two (or a few) GCs with similar metallicities but markedly different HB morphologies and those that investigate the ensemble properties of a large sample of GCs. One canonical pair in the former category is NGC~288, with a blue HB, and NGC~362, with a red HB. The first CCD-based, differential photometric study of these clusters was conducted by \\citet{bo89}. Bolte concluded that NGC~288 is $\\sim$3 Gyr older than NGC~362. The subsequent work of \\citet{gr90}, \\citet{sa90}, and \\citet{va90} echoed this result. By contrast, \\citet{vd90} reached a very different conclusion. Using the brightest RGB star in each GC, \\citet{vd90} corrected the photometry of NGC~288 and NGC~362 for their relative distances and found their main sequence turnoffs to be roughly coincident, thereby suggesting a negligible age difference. \\citet{be01} found NGC~288 to be 2$\\pm$1 Gyr older than NGC~362 using three different techniques; in a follow-up, \\citet{ca01} found that an age difference of 2 Gyr was plausible if both GCs have $\\feh \\sim -1.2$ but that their synthetic HB models were unable to match the detailed HB morphology of either GC using canonical assumptions of average mass loss and dispersion on the HB. In considering differences between second parameter pairs, it is important to understand just how similar the abundances are in both clusters. \\citet{sh00} performed a detailed comparison of abundances in red giants with 13 stars in NGC~288 and 12 in NGC~362. These authors concluded that NGC~288 has a lower $\\feh$ than NGC~362 by 0.06 dex, that the average $\\afe$ ratios are very similar, and that both GCs exhibit variations among O, Na, and Al. The other canonical second parameter pair is M~13, with a blue HB, and M~3, with an intermediate HB. M~13 and M~3 are $\\sim$0.2 dex more metal poor than NGC~288 and NGC~362. \\citet{va90} found evidence for an age difference but the quality of their CMDs did not allow a more definitive statement. \\citet{ca95} estimated an upper limit to the age difference of $\\sim$3 Gyr and concluded that it was insufficient to explain the HB morphologies, assuming both GCs have the same chemical composition. Alternatively, \\citet{jo98} suggested the difference in their HB morphologies was due to a difference in their He abundances. Johnson \\& Bolte suggested that M~13 had a He mass fraction $\\sim$0.05 greater than M~3. However, \\citet{sw98} showed that a difference of $\\Delta$Y $\\sim 0.05$ at constant $\\feh$ would make the level of the HB brighter by $\\sim$0.2 mag. Such a difference between M~3 and M~13 was ruled out by the photometry of \\citet{re01}, who also concluded that M~13 is older than M~3 by $1.7\\pm0.7$ Gyr. \\citet{sn04} compared spectra of 28 red giants in M~3 with 35 in M~13 and found the two GCs' mean $\\feh$ values the same within their 1-$\\sigma$ error bars. However, \\citet{sn04} reported differences in light element abundance variations, particularly O, for which M~13 displays a larger range of variation than M~3 by $\\sim$0.5 dex. A discussion of the search for the second parameter in GC-to-GC comparisons would not be complete without including the work of \\citet{st99} and \\citet{do08b}. These investigations used HST WFPC2 photometry to measure the ages of the outer halo GCs Palomar~3, Palomar~4, and Eridanus \\citep{st99} and AM-1 and Palomar~14 \\citep{do08b} relative to inner halo GCs with similar metallicities, M3 and M5\\footnote{There is substantial evidence that M~5 is actually an outer halo GC currently near its perigalacticon, see \\citet{sc96} and \\citet{di99}.}. Each of the five outer halo GCs has a redder HB morphology than its comparison inner halo GC. Both studies concluded that the outer halo GCs are $\\sim$1.5-2 Gyr younger than M3 and M5, provided the chemical compositions of the outer halo GCs are comparable to their inner halo counterparts. A recent study by \\citet{ko09} found the abundances of Pal~3 are essentially the same as inner halo GCs of similar metallicity and thus the relative age comparison with M~3 by \\citet{st99} was justified. Overall, the outer halo GCs' chemical compositions remain poorly understood by comparison with the inner halo, see e.g.\\,\\citet{pvi05}. \\citet{ca00} found the reported age differences between Pal~4 and Eridanus, with red HBs, and M~5, with an intermediate HB, too small to explain the difference in HB morphologies unless all three GCs are younger than 10 Gyr, assuming standard assumptions of mass loss on the RGB. \\citet{ca01b} concluded that it was possible to explain the difference in HB morphology between Pal~3 and M~3 if the former has less HB mass dispersion and is younger than the latter. These examples indicate that if age is the second parameter, then it is our lack of understanding of mass loss that confuses GC-to-GC comparisons as first pointed out by \\citet{ro73}. The review by \\citet{ca05} includes a thorough discussion of different mass loss prescriptions and their efficacy. Given the HB morphologies and relatively young ages of the most distant outer halo GCs\\footnote{The exception in the outer halo is NGC~2419. A number of dedicated HST photometric studies have focused on this massive, distant GC. \\citet{ha97} and \\citet{sa08} both found NGC~2419 to be an old, metal poor GC with a blue HB.}, it is important to include them when considering the properties of the entire Galactic GC population, especially in the context of Searle \\& Zinn's halo formation scenario. The second category of second parameter studies are those that consider the properties of a large sample of Galactic GCs, e.g.\\,\\citet{sz78}. In the first CCD-era study, \\citet{sa89} compiled properties for 31 Galactic GCs. Among other things, \\citet{sa89} examined the variation of HB morphology with age for GCs in a narrow range of metal abundance and showed that GCs with red HBs are significantly younger than those with blue HBs by as much as 5 Gyr. The subsequent studies of \\citet{ch92}, \\citet{sa95}, and \\citet{ch96} updated and reaffirmed this result. \\citet{ro99} found that the GCs with $\\rgc > 8$ kpc are, on average, younger than those with $\\rgc < 8$ kpc, which is consistent with age as the second parameter. By contrast, \\citet{ri96} examined the CMDs of 36 GCs, found an age dispersion of $\\sim$1 Gyr with no significant age gradient in the Galactic halo, and concluded that this age range was too small for metallicity and age alone to explain HB morphology in GCs. Theoretical efforts have provided further insights into the complexities of HB morphology. In particular, several studies have applied the synthetic HB model developed by \\citet{ro73}. \\citet{ldz90}, \\citet{ca93}, and \\citet{ldz94} used synthetic HB models to explore the interplay of HB morphology, metallicity, and age in the CMD. These studies demonstrated considerable degeneracy in the HB morphology--metallicity diagram and, in particular, \\citet{ca93} argued that unless absolute values of the He abundance, $\\afe$ ratios, and RGB mass loss were known, HB morphology did not constitute a reliable age constraint. Nevertheless, \\citet{ldz94} concluded that there was evidence for an age dispersion of $\\la 5$ Gyr among Galactic GCs of similar metallicity but markedly different HB morphologies. A number of theoretical studies have also focused on second parameter pairs or triads but, unfortunately, the lack of a firm theoretical understanding of mass loss along the RGB has impaired the these efforts, as first pointed out by \\citet{ro73}. See \\citet{ca05} for a recent review. As it pertains to HB morphology, the use of the term 'second parameter' has expanded to cover a broad range of complex behaviors. In particular, this includes the faint extent of the blue HB tail in some GCs. \\citet{fp93} analyzed 53 GCs and found that the length of the HB in the CMD--and the extent of the blue tail--is correlated with central density. \\citet{bu97} also reported a link between an extended blue tail and central density in GCs and concluded that `environment is ``a'' second parameter'. \\citet{sm04} showed that there is a correlation between central density and HB morphology for intermediate metallicity GCs ($-1.7 < \\feh < -1.3$) where the second parameter effect is most pronounced. \\citet{rb06} analyzed the properties of 54 Galactic GCs with homogeneous photometry from an HST WFPC2 Snapshot Survey \\citep{pi02} and concluded that more massive GCs tend to have more extended blue HBs. \\citet{rb06} recognized a link between the effective temperature of the hottest HB star in a GC and its mass (as inferred from its integrated luminosity) and suggested that self-pollution could explain the existence of faint blue tails in preferentially massive GCs. Evidence abounds for the presence of chemical abundance variations in all GCs \\citep{gr04,car09a,car09b} and the possible correlation between the degree of abundance variations and the faint extent of the blue HB \\citep{car07}. However, it is unlikely that the existence of faint, blue tails in the HBs of some GCs is directly related to the appearance of anomalously red HBs in others. It is also unlikely that chemical abundance variations will unduly affect age estimates of most GCs, provided the total metal content is constant across all stars \\citep{pie09}. For cases in which the total metal abundance is not constant, or there are distinct stellar populations present in the CMD, age estimates are necessarily more complicated and careful analysis of each of these GCs is needed. See \\citet{pi09} for a recent summary of GCs with multiple stellar populations visible in the CMD, several of which were discovered with photometry from the ACS Survey of Galactic GCs. The complex issue of multiple stellar populations in GCs remains poorly understood and the extent to which multiple-population GCs permeate the Galactic GC population is unknown at present. To summarize, the presence of metal poor GCs with red HBs that primarily reside in the outer Galactic halo is well-known observationally. Although the present study is focused on the Galactic GCs, including some of those associated with the Sagittarius dwarf, there is ample evidence to suggest that metal poor GCs with red HBs are also found in the Magellanic Clouds \\citep{jo99,gl08} and Fornax \\citep{bu99}. \\citet{mg04} summarize our current knowledge of the GC populations in the Galaxy and its satellites in the HB morphology--metallicity diagram. Despite an abundance of observational evidence, no consensus has been reached as to what parameter(s) are responsible for the appearance of relatively red HBs in metal poor GCs. Age is frequently offered as a candidate but considerable doubt still remains because the age difference claimed--or required by theoretical studies of HB morphology--is too large to satisfy the observations. The existence of a homogeneous database of deep, high quality photometry from the ACS Survey of Galactic GCs \\citep[Paper I in this series]{sa07} has motivated a re-examination of HB morphology and its relation to a variety of GC properties. The paper is organized as follows: $\\S$\\ref{data} describes the data sample; $\\S$\\ref{hbm} describes the methods that were employed to determine the HB morphologies; $\\S$\\ref{gcages} describes the sources of GC ages and provides some discussion of complicating factors in age determination; $\\S$\\ref{results} presents the analysis performed on the assembled data and discusses the important results; and, finally, $\\S$\\ref{conclusion} provides a summary of the salient points. ", "conclusions": "} HB morphologies characterized by $\\dvi$ of 66 Galactic GCs were examined to determine the sensitivity of HB morphology to a variety of different factors. Deep, homogeneous photometry from the ACS Survey of Galactic GCs accounts for 60 of these while the remaining 6 GCs are the most distant Galactic GCs known. The complete sample is the largest examined to date and consists solely of high-quality HST photometry. It spans the full range of $\\rgc$ and almost the entire range of metallicity present in the Galactic GC population. $\\dvi$ values and two independently measured sets of ages were joined with other quantities from the literature to assess the relative importance of these quantities on HB morphology. The data were split into two groups, roughly equal in number, based on $\\rgc$. The tight relationship between metallicity and HB morphology in the inner halo group ($\\rgc < 8$ kpc) was characterized by a fitting function and this trend was subtracted off of the outer halo group. The difference between fit and measured $\\dvi$ was then compared to a variety of parameters, of which only age showed a significant correlation. The age correlation does not rely on the presence of the 6 most distant GCs in the analysis, though their presence does strengthen the result. Hence we conclude that, after metallicity, age has the most influence on $\\dvi$. The age spread among the bulk of GCs in the Galactic halo was found to be 2-2.5 Gyr, though there are a few younger outliers such as Pal~12 and Ter~7. Further analysis, in which both metallicity and age were restricted, provided strong evidence that central luminosity density ($\\rho_0$) is the third most influential parameter on $\\dvi$." }, "0911/0911.3659_arXiv.txt": { "abstract": "We derive a relation for the steepening of blazar $\\gamma$-ray spectra between the multi-GeV {\\it Fermi} energy range and the TeV energy range observed by atmospheric \\v{C}erenkov telescopes. The change in spectral index is produced by two effects: (1) an intrinsic steepening, independent of redshift, owing to the properties of emission and absorption in the source, and (2) a redshift-dependent steepening produced by intergalactic pair production interactions of blazar $\\gamma$-rays with low energy photons of the ``intergalactic background light'' (IBL). Given this relation, with good enough data on the mean \\gray SED of TeV Selected BL Lacs, the redshift evolution of the IBL can, in principle, be determined independently of stellar evolution models. We apply our relation to the results of new {\\it Fermi} observations of TeV selected blazars. ", "introduction": "Stecker \\& Scully (2006) (SS06) derived a simple analytic expression for the change in spectral index of a TeV \\gray source in the redshift range between 0.05 and 0.4. They showed that the change in the spectral index caused by intergalactic absorption is given by an approximately linear relation in redshift, {\\it i.e.}, $\\Delta \\Gamma_{a} \\simeq C + Dz$. The purpose of this letter is to generalze this relation by including the effect of intrinsic steepening in the source spectra between the mutli-GeV energy range observed by {\\it Fermi} and the TeV range observed by atmospheric \\v{C}herenkov telescopes. Our general result is roughly independent of the specific model of the intergalactic background light (IBL) used, because it only depends on the shape of the average galaxy spectral energy distribution (SED) on the near IR side of the starlight peak that determines the absorption in the TeV energy range. We compare our relation for the specific baseline and fast evolution models of Stecker, Malkan \\& Scully (2006) (SMS06) with the results of recent {\\it Fermi} observations of 13 TeV selected BL Lac AGN. We also show how it can be used to independently determine the redshift evolution of the IBL. ", "conclusions": "We have derived a simple analytic approximation for determining the steepening in the spectra of TeV selected BL Lac AGN and compared our results with recent observational data on 13 TeV selected BL Lac AGN. Our relation is in excellent agreement with the observational data and provides a framework for understanding and interpreting both these and observational data. SS06 have shown that the effect of intergalactic absorption on the spectra of AGN in the energy range 0.2 TeV $ < E_{\\gamma} <$ 2 TeV and the redshift range $0.05 < z < 0.4$ is a simple power-law to power-law steepening. Absorption in this energy range is primarily produced by interactions with near infrared photons from on the low energy sides of the starlight peaks in galaxy SEDs. The {\\it shape} of the resulting peak in the intergalactic SED produces an approximately logarithmic energy dependence for the function $\\tau (E_{\\gamma})$. This energy dependence should be roughly the same for all models of the IBL. Therefore, our general result of a linear $\\Delta\\Gamma = (C+K) + Dz$ relationship is roughly independent of the specific IBL model used because it only depends on the shape of the average galaxy SED on the near IR side of the IBL starlight peak. The parameters $C, D$ and $K$ will be different for the different models since the absolute value of $\\tau$ varies from one IBL model to another. Given our derived relation between $\\Delta \\Gamma$ and redshift, and with good enough data on the mean \\gray SED of TeV Selected BL Lacs used to determine $K$, the redshift evolution of the IBL can, in principle, be determined independently of stellar evolution models." }, "0911/0911.3974_arXiv.txt": { "abstract": "The SPace Infrared telescope for Cosmology and Astrophysics (SPICA) is a proposed mid-to-far infrared (4-200 $\\mu$m) astronomy mission, scheduled for launch in 2017. A single, 3.5m aperture telescope would provide superior image quality at 5-200 $\\mu$m, and its very cold ($\\sim$5 K) instrumentation would provide superior sensitivity in the 25-200 $\\mu$m wavelength regimes. This would provide a breakthrough opportunity for studies of exoplanets, protoplanetary and debris disk, and small solar system bodies. This paper summarizes the potential scientific impacts for the proposed instrumentation. ", "introduction": "The SPace Infrared telescope for Cosmology and Astrophysics (SPICA) is a proposed mission for mid-to-far infrared (MIR/FIR) astronomy, consisting of a single 3.5-m aperture space telescope with cooled ($\\sim$5 K) instrumentation (see Nakagawa 2008). This mission would provide a significant step forward in detection sensitivity in the 4 to 200 $\\mu$m wavelength regime, which would revolutionize our understanding of how galaxies, stars and planets form, and how interactions between complex astrophysical processes have ultimately led to the formation of our own Solar system and the emergence of life on Earth. The role of SPICA would complement other forthcoming space infrared missions; Herschel and JWST. Herschel will provide improved capabilities at wavelengths longer than 57 $\\mu$m, while its relatively warm temperature will produce a modest improvement over the Spitzer Space Telescope. JWST will provide the best performance at wavelengths shorter than 25 $\\mu$m, and will have the highest angular resolution. SPICA would (1) achieve the best sensitivity at 25--200 $\\mu$m, due to its cold mirror ($\\sim$5 K); (2) achieve a clean point-spread function at 5--200 $\\mu$m due to its non-segmented mirror, thereby providing the best performance for high-contrast coronagraphy; and (3) fill the gap in wavelength coverage between Herschel and JWST. SPICA would offer a unique opportunity for studying exoplanets, planet formation, circumstellar disks, and small bodies in the solar system. Table 1 summarizes the instruments proposed to date. Those selected for the mission well be chosen based on technical constraints (weight, volume, heat dissipation etc.) and scientific impact. In \\S 2, we describe potential scientific achievements in the above areas of research. In \\S 3 we briefly describe the project schedule, including the selection process for the instruments. ", "conclusions": "SPICA would provide significant advances in the study of exoplanets, protoplanetary disks, debris disks and small bodies in the solar system. Its imaging capability would be able to detect debris disks with an unprecedented sensitivity. These capabilities would allow to measure the total flux of small bodies in the solar system, useful for the study of their size distribution and albedos. SPICA's spectral capability would probe the gas and dust associated with circumstellar disks, and the atmospheres of the transiting planets. Its coronagraphic capability would lead to the discovery of a number of exoplanets, and the subsequent investigation of their atmospheres and the geometry and distribution of ice in the spatially resolved circumstellar disks. Possible achievements with the individual instruments are summarized in Fig 4. The conceptual design of SPICA is currently underway, and we will begin the definition phase of the mission in 2009. This phase, which will include the final decisions for the instruments, will be completed in 2011. \\\\ \\\\ {\\it Acknowledgement} --- We are grateful to Dr. Yuri Aikawa and Akira Kouchi for useful discussions. We thank Dr. Jennifer Karr for reading our manuscript, and the editor and anonymous referee for useful comments." }, "0911/0911.3145_arXiv.txt": { "abstract": "From the analysis of measurements of the linear polarisation of visible light coming from quasars, the existence of large-scale coherent orientations of quasar polarisation vectors in some regions of the sky has been reported. Here, we show that this can be explained by the mixing of the incoming photons with nearly massless pseudoscalar (axion-like) particles in extragalactic magnetic fields. We present a new treatment in terms of wave packets and discuss its implications for the circular polarisation. ", "introduction": "In this work~\\cite{paper}, we are looking at the effect that axion-photon mixing can have on the polarisation of light emitted by distant astronomical sources. In particular, the observation of redshift-dependent large-scale coherent orientations of AGN polarisation vectors~\\cite{hutsemekers} can, at least qualitatively, be reproduced as a result of such a mixing of incoming photons with extremely light axion-like particles as they travel inside external magnetic fields. These observations (see Figure~\\ref{fig:hutsemekers}) are based on good-quality measurements in visible light of the linear polarisation for a sample of 355 quasars from different groups, working on different instruments. \\begin{center} \\begin{figure}[h] \\includegraphics[trim = 10mm 10mm 0mm 0mm, clip, width=0.8\\textwidth]{fig_map_nc.ps} \\caption{Maps of the same region of the sky in right ascension ($x$ axis) and declination ($y$ axis), both given in degrees, for AGN characterised by different redshifts, $z$. These polarisation vectors are thus for objects approximately along the same line of sight but at different distances from us. In the low-$z$ case, the average direction is $\\bar{\\theta}\\approx 79^\\circ$ while at high-$z$, it would be $\\bar{\\theta}\\approx 8^\\circ$ ($\\theta$ being counted from North to East)~\\cite{hutsemekers}. } \\label{fig:hutsemekers} \\end{figure} \\end{center} This has been discussed in terms of axion-photon mixing by several authors, in the case of plane waves, and a prediction from this mixing, in this particular case and in general, is a circular polarisation comparable to the linear one and, hence, observable~\\cite{planewaves,reviewHLPW}. Here, we present a treatment in which light is described by wave packets and show that the circular polarisation can be suppressed with respect to that which is predicted in the plane wave case. ", "conclusions": "We have reviewed axion-photon mixing in the case of plane waves and have briefly presented our new formalism in terms of wave packets. The main consequence of this new treatment is the net decrease of circular polarisation with respect to what is predicted using plane waves. From this we conclude that the lack of circular polarisation in the light from AGN does not rule out the ALP-photon mixing. \\begin{theacknowledgments} A.~P. would like to thank the IISN for funding and to acknowledge constructive discussions on physical and numerical matters with Fredrik Sandin and Davide Mancusi. D.~H. is senior research associate FNRS. \\end{theacknowledgments}" }, "0911/0911.3235_arXiv.txt": { "abstract": "Motivated by recent observations from Pamela, Fermi and H.E.S.S., we consider dark matter decays in the framework of supersymmetric SU(5) grand unification theories. An SU(5) singlet $S$ is assumed to be the main component of dark matters, which decays into visible particles through dimension six operators suppressed by the grand unification scale. Under certain conditions, $S$ decays dominantly into a pair of sleptons with universal coupling for all generations. Subsequently, electrons and positrons are produced from cascade decays of these sleptons. These cascade decay chains smooth the $e^{+}+e^{-}$ spectrum, which permit naturally a good fit to the Fermi LAT data. The observed positron fraction upturn by PAMELA can be reproduced simultaneously. We have also calculated diffuse gamma-ray spectra due to the $e^{\\pm}$ excesses and compared them with the preliminary Fermi LAT data from 0.1 GeV to 10 GeV in the region $0^{\\circ}\\leq l\\leq360^{\\circ},10^{\\circ}\\leq|b|\\leq20^{\\circ}$. The photon spectrum of energy above 100 GeV, mainly from final state radiations, may be checked in the near future. ", "introduction": "Electron, proton, photon, neutrino and their antiparticles are stable, at least on the cosmological time scale. Detection of these particles from cosmic rays provides an interesting window to look into the deep universe. Recently, the PAMELA experiment reported a significant excess in the positron fraction $e^{+}/(e^{+}+e^{-})$ between $10$ GeV and $100$ GeV \\cite{Adriani:2008zr}. On the other hand, the measured antiproton to proton flux ratio appears to be consistent with predictions \\cite{Adriani:2008zq}. More recently, the Fermi LAT collaboration observed a smooth $e^{+}+e^{-}$ spectrum with high accuracy. It is found to be falling as $E^{-3.0}$ from $20$ GeV to $1$ TeV \\cite{Abdo:2009zk}, much harder than the predictions of conventional models. The H.E.S.S. collaboration measured the $e^{+}+e^{-}$ spectrum from $600$ GeV up to several TeV \\cite{Aharonian:2009ah}, which is consistent with the Fermi data in overlapping regions and steepens at about $1$ TeV towards higher energy. These excesses of electrons and positrons could be due to unidentified astrophysical sources, e.g., nearby pulsars or supernova remnants \\cite{Stawarz:2009ig,Piran:2009tx,Grasso:2009ma}. However, an explanation via dark matter (DM) annihilation or decay is, arguably, a much more interesting possibility, at least from the perspective of particle physics. The electron and positron spectra alone, even with higher precision and broader energy range, cannot decisively decide which explanation is more plausible \\cite{Malyshev:2009tw}. Hopefully, the energy spectrum and the angular dependence of cosmic gamma rays \\cite{Bertone:2008xr,Zhang:2008tb,Bergstrom:2008ag,Ibarra:2009nw,Zhang:2009kp}, to be measured by the Fermi LAT in the near future, may provide a more definite answer. For the DM interpretation, the mass of the DM should be around several TeV, to provide the $e^{\\pm}$ excesses from $20$ GeV to $1$ TeV and steepen sharply above $1$ TeV. Furthermore, traditional WIMP DM candidates usually produce extra antiprotons. As Pamela does not observe any deviation on antiproton spectrum from the anticipation, WIMP DMs are now disfavored as potential sources of the observed cosmic-ray excesses. Still, there are plenty of freedoms for both DM annihilation and decay to reproduce the experimental $e^{\\pm}$ spectra reasonably \\cite{Meade:2009iu,Bergstrom:2009fa}. For DM annihilation, a large boost factor in the order of $10^{2}$ to $10^{3}$ is needed for the theory to be consistent with the relic abundance measured by the WMAP \\cite{Dunkley:2008ie}. As the clumpiness property of the DM distribution falls far short of such a large factor, one usually resorts to nonperturbative Sommerfeld \\cite{Hisano:2004ds,Hisano:2006nn,Cirelli:2007xd,ArkaniHamed:2008qn} or Breit\\textendash{}Wigner \\cite{Feldman:2008xs,Ibe:2008ye,Guo:2009aj} enhancement in model buildings. For DM decays, the lifetime should typically be around the order of $10^{26}s$ to fit the $e^{\\pm}$ data\\cite{Meade:2009iu,Shirai:2009fq,Mardon:2009gw}, which is much longer than the lifetime of the universe. Therefore the DM decay rates will not affect the relic abundance appreciably. The energetic $e^{\\pm}$ flux produced from DM annihilations/decays would inevitably emit gamma rays. These gamma rays depend on the DM density as $\\rho^{2}$ for annihilations and $\\rho$ for decays. This will lead to different angular dependence of the gamma ray spectrum, which may be measurable in the near future to differentiate these two scenarios. The gamma ray spectrum can also be used to differentiate DM explanations from astrophysical ones. In this paper, we will focused on DM decays. Notice that a lot of suppression will be needed for a TeV scale particle to have a lifetime $\\sim10^{26}s$. If it decays via dimension four operators, tremendous fine tunings will be needed. If it decays via dimension six operators, it still needs to be suppressed by a scale $\\sim10^{16}$ GeV, which turns out to coincide with the grand unification theory (GUT) scale \\cite{Langacker:1991an,Amaldi:1991cn,Ellis:1990wk}. In the same spirit of Refs. \\cite{Arvanitaki:2008hq,Arvanitaki:2009yb}, we will take a singlet as the dark matter candidate and provide a detailed analysis in the frame of supersymmetric SU(5) GUT. To be consistent with the Pamela antiproton measurement, squark masses are assumed to be heavier than that of the SU(5) singlet $S$, so the $S$ decay would be quark phobic. The $S$ then decays dominantly into slepton pairs with a universal coupling for all generations. These sleptons decay quickly into leptons and lightest supersymmetric particles (LSPs), if R-parity is conserved. In this framework, we have obtained a reasonable fit to all $e^{\\pm}$ data from Pamela, Fermi and H.E.S.S.. The $e^{\\pm}$ fluxes from $S$ decays are inevitably accompanied by hard photons: coming from final state radiations (FSR) of cascade decays, including $S\\to\\tilde{\\tau}\\to\\tau\\to\\pi^{0}\\to2\\gamma$ and the inverse Compton scattering (ICS) on the interstellar radiation field (ISRF). The gamma ray fluxes could have Galactic and extragalactic origins. We have calculated all these gamma ray spectra and compared them with the recent Fermi LAT measurement in the region $0^{\\circ}\\leq l\\leq360^{\\circ},10^{\\circ}\\leq|b|\\leq20^{\\circ}$ \\cite{Porter:2009sg}. This paper is organized as follows. The supersymmetric SU(5) model plus a singlet $S$ is presented in Section II, where we have also discussed the possible decay channels of $S$ in some detail. In Section III, a reasonable fit is obtained to reproduce the observed $e^{\\pm}$ fluxes, by tuning relevant parameters in the model. Section IV is devoted to the study of gamma-ray spectra from $e^{\\pm}$ excesses. Finally we conclude with a summary in section V. The component field structure of the dimension six effective operators will be presented in the Appendix. In this paper, we have used the Navarro-Frenk-White (NFW) halo model \\cite{Navarro:1996gj} for DM distribution and the MED propagation model \\cite{Donato:2003xg,Delahaye:2007fr}. For other halo and propagation models, the conclusions are similar. In addition, all computations on astrophysical effects are performed semi-analytically instead of using the GALPROP program% \\footnote{Web page: http://galprop.stanford.edu/web\\_galprop/galprop\\_home.html% }. ", "conclusions": "In this paper we have studied the DM decay in supersymmetric SU(5) models. An SU(5) singlet $S$, instead of LSP, is assumed to be the dominant component of DM. With R-parity conservation and a spontaneously broken $Z_{2}$ symmetry, the singlet $S$ can decay into visible particles through dimension six effective operators suppressed by the GUT scale. Assuming the squarks to be heavier than $S$, $S$ decays dominantly into a pair of sleptons through the effective operator $\\widetilde{s}^{*}(\\widetilde{l_{L}}^{*}\\square\\widetilde{l_{L}}+\\widetilde{l_{R}}\\square\\widetilde{l_{R}}^{*})$. Typically, the lifetime of $S$ is around $10^{26}$ s, much longer than the age of the Universe. Since the decay products of $S$ do not contain any quarks, our model is consistent with the Pamela antiproton measurement automatically. For illustration, we have chosen $M^{DM}=6.5$ TeV, $M_{GUT}=10^{16}$ GeV, $M_{\\widetilde{e}}=380$ GeV, $M_{\\tilde{\\mu}}=370$ GeV, $M_{\\widetilde{\\tau}}=330$ GeV and $M_{LSP}=300$ GeV. With this parameter set, we have shown that the $S$ decays can account for the PAMELA, H.E.S.S. and Fermi LAT $e^{\\pm}$ excesses. Numerically, we have adopted the NFW profile for the dark matter distribution and the MED propagation model for the cosmic ray propagation. To interpret these results, one should keep in mind that there exist substantial astrophysical uncertainties about $e^{\\pm}$ background, $e^{\\pm}$ propagation and the DM distribution. When $e^{\\pm}$ propagate from the decay position to the Earth, hard photons are emitted inevitable due to inverse Compton scattering and final state radiation. Future measurements of the diffuse gamma ray may distinguish DM explanations from astrophysical explanations by looking into the energy and angular distributions. We have calculated the gamma ray spectra in our model. The predicted photon spectrum are compared to the preliminary Fermi LAT measurement from $0.1$ GeV to $10$ GeV in the region $0^{\\circ}\\leq l\\leq360^{\\circ},~10^{\\circ}\\leq|b|\\leq20^{\\circ}$, which seems to be consistent with each other. The total gamma ray spectrum are dominated by photons from Galactic final state radiation for the photon energy above $100$ GeV, which may be tested by Fermi LAT in the near future." }, "0911/0911.0416_arXiv.txt": { "abstract": "We present three-dimensional simulations on a new mechanism for the detonation of a sub-Chandrasekhar CO white dwarf in a dynamically unstable system where the secondary is either a pure He white dwarf or a He/CO hybrid. For dynamically unstable systems where the accretion stream directly impacts the surface of the primary, the final tens of orbits can have mass accretion rates that range from $10^{-5}$ to $10^{-3} M_{\\odot}$ s$^{-1}$, leading to the rapid accumulation of helium on the surface of the primary. After $\\sim 10^{-2}$ $M_{\\odot}$ of helium has been accreted, the ram pressure of the hot helium torus can deflect the accretion stream such that the stream no longer directly impacts the surface. The velocity difference between the stream and the torus produces shearing which seeds large-scale Kelvin-Helmholtz instabilities along the interface between the two regions. These instabilities eventually grow into dense knots of material that periodically strike the surface of the primary, adiabatically compressing the underlying helium torus. If the temperature of the compressed material is raised above a critical temperature, the timescale for triple-$\\alpha$ reactions becomes comparable to the dynamical timescale, leading to the detonation of the primary's helium envelope. This detonation drives shockwaves into the primary which tend to concentrate at one or more focal points within the primary's CO core. If a relatively small amount of mass is raised above a critical temperature and density at these focal points, the CO core may itself be detonated. ", "introduction": "White dwarfs (WDs), the end point of stellar evolution for most stars, are extremely common, with about $10^{10}$ of them residing within the Milky Way \\citep{Napiwotzki:2009p3570}. WDs are frequently observed in binary systems with normal stellar companions, and, less frequently, with compact stellar companions. Double degenerate (DD) systems are those in which the companion is another WD. Typically, DD systems are formed via common envelope evolution \\citep{Nelemans:2001p1711}. Often, the final result of this evolutionary process is a binary consisting of a carbon-oxygen (CO) WD and a lower-mass helium WD companion \\citep{Napiwotzki:2007p1710}. For decades, Type Ia supernovae (SNe) have been employed as standard candles. Still, even the preferred mechanism exhibits substantial variability \\citep{Timmes:2003p2073, Kasen:2009p3427}. Complicating this issue further is that WDs may not need to exceed the Chandrasekhar limit to be capable of exploding; other mechanisms include collisions between WDs in dense stellar environments such as the cores of globular clusters \\citep{Rosswog:2009p3556}, tidal encounters with moderately massive black holes \\citep{Rosswog:2008p3059, Rosswog:2009p3553, RamirezRuiz:2009p3071}, or as we are introducing in this paper, the accretion of dense material from a He WD companion. At the distance of tens of WD radii, gravitational radiation can bring two WDs close enough for the less-massive WD (secondary) to overfill its Roche lobe and begin transferring mass to the more-massive WD (primary). Because of the mass-radius relationship of WDs, mass transfer is often unstable and can eventually lead to a merger \\citep{Marsh:2004p1880, Gokhale:2007p2850}. For binary mass ratios close to unity the circularization radius $R_{\\rm h}$ drops below the radius of the primary $R_{1}$, and thus the accretion stream will directly impact the primary's surface. Most studies of dynamically unstable DD systems focus on the evolution of the post-merger object and, assuming that the final object has a mass larger than the Chandrasekhar limit, on how the merged remnant may eventually lead to a Type Ia SNe \\citep{Livio:2000p3540, Yoon:2007p1003}. These models all assume that the rapid accretion that precedes coalescence is uneventful. In this \\textit{Letter}, we present three-dimensional hydrodynamics simulations that demonstrate that explosive phenomena are an inevitable results of the high accretion rates characteristic of the final stages of a merger in a DD system. In some cases, these explosive phenomena may lead to the complete detonation of the CO primary. \\begin{figure*}[tb] \\centering\\includegraphics[width=0.5\\linewidth,clip=true,angle=-90]{fig1.eps} \\caption{Setup of run Ba in FLASH, with several annotated regions of interest. The color scheme shows $\\log T$ through a slice of the orbital plane. The cyan circle shows the spherical outflow boundary condition centered about the secondary, while the cyan wedge shows a cross-section of the cone used as the mass inflow boundary. The dashed contour shows the system's Roche surface. The white boxes with labels are described in Section \\ref{accinsta}.} \\label{slicefig} \\end{figure*} ", "conclusions": "\\label{results} Surface detonations are present in both of the longer runs with a 0.9 $M_{\\odot}$ CO primary, although run B detonates much later in its evolution than run C. The geometry and evolution of the detonations are similar in both runs. A small region of the surface helium is heated to a temperature of $\\sim 2 \\times 10^{9}$ K, leading to a detonation front that expands outwards from the ignition site along the primary's surface (Figure \\ref{volfig}, second column). Because the helium layer is toroidal rather than spherical, the front runs out of fuel as it propagates towards the primary's poles. Consequently, the detonation splits into two fronts that run clockwise and counterclockwise around the equator of the primary (Figure \\ref{volfig}, third column). These detonation fronts eventually run into each other along a longitudinal line that is opposite to the original ignition site (Figure \\ref{volfig}, fourth column). The highest temperatures are produced in this convergence region. For run A, the primary's surface gravity is too low to compress the helium layer above the critical temperature necessary for thermonuclear runway. No detonation was observed in run Ba, despite being initialized using a checkpoint from run B just prior to the observed detonation in B. This is because of the stochastic nature of the ignition mechanism --- the particular dense knot of material that led to the surface detonation in B has a different shape in Ba, and did not have a favorable geometry for igniting the helium torus. And because Ba has twice the linear resolution of run B, we could only afford to evolve it for a short period of time. \\begin{figure}[tb] \\centering\\includegraphics[width=\\linewidth,clip=true]{fig4.eps} \\caption{Maximum temperature $T$ in a mass shell as a function of interior mass $M\\left(< r\\right)$ and time $t$ for run B. Each shell lies on a surface of equal gravitational potential. The shading is logarithmically binned, with red corresponding to $T_{\\max}$ and blue corresponding to $T_{\\min} = T_{\\max}/10^{1.5}$. The dashed line shows the region where $X_{\\rm He} > 0.02$, while the dotted line shows where $X_{\\rm Ne} > 10^{-3}$. For convenience, we include a zoomed view of the surface detonation as an inset.} \\label{radialfig} \\end{figure} If a helium surface detonation occurs as the result of stream instabilities in a DD system, the resulting transient event could potential resemble a dim Type Ia SNe \\citep{Bildsten:2007p2743, Perets:2009p3232}. The results of our simulations shows a large degree of variability which ultimately depends on the geometry of the dense knots when they strike the surface of the primary in the impact zone. However, if the post-detonation temperature is not much larger than $2 \\times 10^{9}$ K or the geometry of the detonation fronts are not favorable, it is possible to only synthesize intermediate-mass $\\alpha$-elements (Table \\ref{outcomes}). The prospect of igniting the CO core itself is attractive because the typical densities of moderate-mass WDs ($M \\sim 1.0 M_{\\odot}$) are $\\sim 10^{6}-10^{7}$ g cm$^{-3}$, comparable to the densities found in the DDT for Type Ia SNe \\citep{Khokhlov:1997p3443}. \\cite{Fink:2007p1886} (hereafter FHR) show that when $0.1 M_{\\odot}$ of helium is ignited at the He/CO interface of a moderate-mass WD, strong shocks are launched deep into the CO core, which converge at a focusing point. This focusing produces a region where the temperature and density meet the criteria for explosively igniting carbon \\citep{Niemeyer:1997p1901, Ropke:2007p3332}, thus leading to a full detonation of the CO core. The linear resolution of our simulations is substantially coarser than the highest-resolution two-dimensional simulations of FHR, and we are certainly under-resolving the true density and temperature peaks. However, FHR's results appear to be rather optimistic when applied to our model for the following reasons. First, FHR assumes perfect mirror symmetry, which dramatically increases the amount of focusing by forcing the shocks to converge at a single point rather than a locus of points. Second, FHR does not use a nuclear network and makes the simplifying assumption that the entire He envelope is burned to Ni, which naturally overestimates the energy injected into the CO core when compared to the more realistic case where burning is incomplete. Because of these differences, the conditions for double detonation are still not met even in our run with the most favorable shock geometry and strongest surface detonation (run C). In the focusing region of run C, the temperature and density are $2.5 \\times 10^{8}$ K and $2.0 \\times 10^{7}$ g cm$^{-3}$, respectively. Conditions for CO core detonation may be more propitious in systems containing a slightly more massive primary, but ultimately the successful detonation of the core depends on the precise distribution of temperature and density within the high-pressure regions of the convergence zone \\citep{Seitenzahl:2009p1881}. The mechanism reported here for the detonation of a sub-Chandrasekhar CO WD in a dynamically unstable binary is not tied to a particular mass scale and therefore allows for considerably more diversity. As outlined above, mass transfer between a pure He WD or a He/CO hybrid and a CO WD before the merger provides a novel pathway to ignite CO WDs. Even if a critical amount of mass is not raised above the conditions required for CO detonation, a peculiar underluminous optical transient should signal the last few orbits of a merging system." }, "0911/0911.5233_arXiv.txt": { "abstract": "In recent years we have been witnessing the discovery of one extrasolar gas giant after another. Now the time has come to detect more low-mass planets like Super-Earths and Earth-like objects. An interesting question to ask is: where should we look for them? We have explored here the possibility of finding Super-Earths in the close vicinity of gas giants, as a result of the early evolution of planetary systems. For this purpose, we have considered a young planetary system containing a Super-Earth and a gas giant, both embedded in a protoplanetary disc. We have shown that, if the Super-Earth is on the internal orbit relative to the gas giant, the planets can easily become locked in a mean motion resonance. This is no longer true, however, if the Super-Earth is on the external orbit. In this case we have obtained that the low-mass planet is captured in a trap at the outer edge of the gap opened by the giant planet and no first order mean motion commensurabilities are expected. Our investigations might be particularly useful for the observational TTV (Transit Timing Variation) technique. ", "introduction": " ", "conclusions": "" }, "0911/0911.5696_arXiv.txt": { "abstract": "The Transit Timing Variation (TTV) method relies on monitoring changes in timing of transits of known exoplanets. Non-transiting planets in the system can be inferred from TTVs by their gravitational interaction with the transiting planet. The TTV method is sensitive to low-mass planets that cannot be detected by other means. Here we describe a fast algorithm that can be used to determine the mass and orbit of the non-transiting planets from the TTV data. We apply our code, {\\tt ttvim.f}, to a wide variety of planetary systems to test the uniqueness of the TTV inversion problem and its dependence on the precision of TTV observations. We find that planetary parameters, including the mass and mutual orbital inclination of planets, can be determined from the TTV datasets that should become available in near future. Unlike the radial velocity technique, the TTV method can therefore be used to characterize the inclination distribution of multi-planet systems. ", "introduction": "In Nesvorn\\'y \\& Morbidelli (2008) and Nesvorn\\'y (2009) (hereafter NM08 and N09) we developed and tested a fast inversion method that can be used to characterize planetary systems from the observed Transit Timing Variations (TTVs; Agol et al. 2005; Holman \\& Murray 2005). See NM08 and N09 for a technical description of the method. Here we use this new method to solve the TTV inversion problem for an arbitrary planetary system. The results provide a baseline for studies of real exoplanetary systems for which TTVs will be detected. Examples of past work that would greatly benefit from the application of the fast inversion algorithm discussed here include Steffen \\& Agol (2005), Agol \\& Steffen (2007), Miller-Ricci et al. (2008) and Gibson et al. (2009). In \\S2, we briefly describe the TTV inversion method. In \\S3, we apply it to a case with coplanar planetary orbits. Inclined planetary orbits are discussed in \\S4. We show, for example, that the mutual inclination of planetary orbits can be determined from TTVs. This important parameter, which may be used to test planet-migration theories (e.g., Rasio \\& Ford 1996, Goldreich \\& Sari 2003), is not typically available from other existing planet-detection methods. ", "conclusions": "The method developed here can be used to analyze TTVs found for any of the potentially hundreds of planets expected to be discovered by {\\it Kepler} (Beatty \\& Gaudi 2008). {\\it Kepler} should be able to detect transit timing variations of only a few seconds (Holman \\& Murray 2005), which should easily exist in many systems, extrapolating from the radial velocity planets (Agol et al. 2005, Fabrycky 2008). Perhaps the most interesting result that comes out of this work is that the shape of the TTV signal is generally sensitive to the orbital inclination of the non-transiting planetary companion. Thus, the TTV method can provide means of determining mutual inclinations in systems in which at least one planet is transiting. This parameter cannot be determined by other planet-detection methods. TTVIM algorithm can be easily extended to incorporate uncertainties in the transiting planet's parameters. This can be done by sampling dimensions that correspond to the additional parameters. For example, in N09 we extended the NM08 method to the case with $e_1\\neq0$. This may be especially relevant to the transiting planets that will be discovered by Kepler because these planets are expected to have wider orbits, which are less susceptible to the circularizing effects of tides. The low CPU cost of the TTVIM algorithm is the key element which will make such studies possible." }, "0911/0911.4659_arXiv.txt": { "abstract": "{ We report on the results of new simulations of near-infrared (NIR) observations of the Sagittarius~A* (Sgr~A*) counterpart associated with the super-massive black hole at the Galactic Center.}{Our goal is to investigate and understand the physical processes behind the variability associated with the NIR flaring emission from Sgr~A*. } { The observations have been carried out using the NACO adaptive optics (AO) instrument at the European Southern Observatory's Very Large Telescope and CIAO NIR camera on the Subaru telescope (13 June 2004, 30 July 2005, 1 June 2006, 15 May 2007, 17 May 2007 and 28 May 2008). We used a model of synchrotron emission from relativistic electrons in the inner parts of an accretion disk. The relativistic simulations have been carried out using the Karas-Yaqoob (KY) ray-tracing code.} {We probe the existence of a correlation between the modulations of the observed flux density light curves and changes in polarimetric data. Furthermore, we confirm that the same correlation is also predicted by the hot spot model. Correlations between intensity and polarimetric parameters of the observed light curves as well as a comparison of predicted and observed light curve features through a pattern recognition algorithm result in the detection of a signature of orbiting matter under the influence of strong gravity. This pattern is detected statistically significant against randomly polarized red noise. Expected results from future observations of VLT interferometry like GRAVITY experiment are also discussed.}{ The observed correlations between flux modulations and changes in linear polarization degree and angle can be a sign that the NIR flares have properties that are not expected from purely random red-noise. We find that the geometric shape of the emission region plays a major role in the predictions of the model. From fully relativistic simulations of a spiral shape emitting region, we conclude that the observed swings of the polarization angle during NIR flares support the idea of compact orbiting spots instead of extended patterns. The effects of gravitational shearing, fast synchrotron cooling of the components and confusion from a variable accretion disk have been taken into account. Simulated centroids of NIR images lead us to the conclusion that a clear observation of the position wander of the center of NIR images with future infrared interferometers will prove the existence of orbiting hot spots in the vicinity of our Galactic super-massive black hole.} ", "introduction": "\\begin{figure*}[!htb] \\begin{minipage}{\\textwidth} \\centering{\\includegraphics[width=\\textwidth]{12473f1.jpg}} \\end{minipage} \\caption{Sgr A* as it was observed in NIR $L^{'}$-band (3.8 $\\mu$m) on 3 June 2008 between 05:29:00 - 09:42:00 ~(UT time). ({\\it a-f}) show the observed images of Sgr A* after 0, 50, 91, 135, 155 and 212 minutes off the start of observation. The images show that when Sgr A* is in its flaring state the flux changes up to 100\\% in time intervals of the order of only tens of minutes.} \\label{label} \\end{figure*} The nearest super-massive black hole candidate ($\\sim~4\\times10^6 M_\\odot$) lies at the center of our galaxy, as inferred from motions of stars near the Galactic Center (Eckart \\& Genzel 1996, 1997; Eckart et al. 2002; Sch\\\"{o}del et al. 2002; Eisenhauer et al. 2003; Ghez et al. 2000, 2005, 2008; Gillessen et al. 2009). With a luminosity of $10^{-9}-10^{-10} L_{Edd}$, where $L_{Edd}$ is its limiting Eddington luminosity, Sagittarius A*, the radio source associated with this SMBH, is one of the most extreme sub-Eddington sources accessible to observations. However, X-ray and near-infrared (NIR) flares are routinely detected with high spatial and spectral resolution observations (Baganoff et al. 2001; Porquet et al. 2003, 2008; Genzel et al. 2003; Eckart et al. 2004, 2006a-c, 2008a-c; Meyer et al. 2006a,b, 2007; Yusef-Zadeh et al. 2006a,b, 2007, 2008). These short bursts of increased radiation last normally for about 100 minutes and occur four to five times a day (see Fig. 1 for a typical behavior of Sgr~A* during a flaring phase in NIR bandwidth). Recent NIR and X-ray observations have revealed the non-thermal nature of high frequency radiation from Sgr~A* (Eckart et al. 2006a-c, 2008a-c; Gillessen et al. 2006; Hornstein et al. 2007). Sgr~A* is probably visible in the NIR regime only during its flaring state. The short time scale variabilities seen during several observed NIR and X-ray flares argue for an emitting region not bigger than about ten Schwarzschild radii ($r_s=\\frac{2GM}{c^2}=2r_g=1.2\\times10^{12}\\big(\\frac{M}{4\\times10^6 M_{\\odot} }\\big)$ cm) of the associated super-massive black hole (Baganoff et al. 2001; Genzel et al. 2003). We have scaled the relevant physical distances according to the gravitational radius ($r_g$) throughout this paper. The NIR flares are highly polarized and normally have X-ray counterparts, which strongly suggests a synchrotron-self-Compton (SSC) or inverse Compton emission as the responsible radiation mechanism (Eckart et al. 2004, 2006a,b; Yuan et al. 2004; Liu et al. 2006). Several observations have already confirmed the existence of a time lag between the simultaneous NIR/X-ray flares and the flares in the lower frequencies. This is interpreted as a sign for cooling down via adiabatic expansion (Eckart et al. 2006a, 2008b,c; Yusef-Zadeh et al. 2006a,b, 2007, 2008; Marrone et al. 2008, Zamaninasab et al. 2008a). \\begin{figure*}[!t] \\begin{minipage}{1.\\textwidth} \\centering{\\includegraphics[width=\\textwidth]{12473f2.jpg}} \\end{minipage} \\caption{Our sample of light curves of Sgr~A* flares observed in NIR $K_s$ band (2.2$\\mu$m) polarimetry mode. The events were observed on 2004 June 13 (a), 2005 July 30 (b), 2006 June 1 (c), 2007 May 15 (d), 2007 May 17 (e) and 2008 May 28 (f). In each panel, the top shows the de-reddened flux density measured in mJy (black), the middle shows the polarization angle (blue) and the bottom shows the degree of linear polarization (red). The gaps in the light curves are due to the sky background measurements.} \\label{events} \\end{figure*} \\begin{table*}[btp] \\begin{center} \\begin{tabular}{lcccccccr}\\hline \\hline Date (Telescope) & Spectral domain & UT start & UT stop & Max. flux & Min. flux& Average flux & Average polarization \\\\ & & time & time & & & sampling rate & sampling rate \\\\ \\hline \\hline \\\\ 2004 June 13 (NACO)& 2.2$\\mu$m & 07:20:02 & 09:15:08 & 5.19 mJy & 2.21 mJy & 1.2 min & 3 min \\\\ \\\\ & & & & \\\\ 2005 July 30 (NACO)& 2.2$\\mu$m & 02:07:36 & 07:03:39 & 8.19 mJy & 1.23 mJy & 1.2 min & 3 min \\\\ & & & & \\\\ & & & & \\\\ 2006 June 1 (NACO)& 2.2$\\mu$m & 04:26:03 & 10:44:27 & 19.33 mJy & 0.72 mJy & 1.5 min & 2 min \\\\ & & & & \\\\ \\\\ 2007 May 15 (NACO)& 2.2$\\mu$m & 09:08:14 & 09:42:12 & 22.27 mJy & 1.70 mJy & 1.5 min & 2 min \\\\ & & & & \\\\ \\\\ 2007 May 17 (NACO)& 2.2$\\mu$m & 04:42:14 & 09:34:40 & 11.87 mJy & 1.86 mJy & 1.5 min & 2 min \\\\ & & & & \\\\ \\\\ 2008 May 28 (CIAO)& 2.15$\\mu$m & 09:22:51 & 13:00:37 & 7.70 mJy & 0.97 mJy & 3.3 min & 10 min \\\\ & & & & \\\\ \\hline \\hline \\end{tabular} \\end{center} \\caption{Observations log.} \\label{table1} \\end{table*} The other feature related to these NIR/X-ray flares are the claimed quasi-periodic oscillations (QPOs) with a period of $20\\pm5$ minutes, which have been reported in several of these events (Genzel et al. 2003; Belanger et al. 2006; Eckart et al. 2006b,c; 2008a; Meyer et al. 2006a,b, Hamaus et al. (2009)). Short periods of increased radiation (the so called \"NIR flares\", normally around 100 minutes) seem to be accompanied by QPOs. All the studies mentioned above probed this 20$\\pm$5 minutes quasi-periodicity, by performing a sliding window analysis with window lengths of the order of the flaring time. Recently, Do et al. (2009) argued that they did not find any significant periodicity at any time scale while probing their sample of observations for a periodic signal. Their method is based on the Lomb-Scargle periodogram analysis of a sample of six light curves and comparing them with several thousands of artificial light curves with the same underlying red-noise. One must note that the suggested QPOs are transient phenomena, lasting for only very few cycles (50-100 minutes). This kind of behavior, along with the inevitable uncertainty in the red noise power law index determination, makes a clear and unambiguous detection ($>5\\sigma$) of a periodic signal very difficult. Whenever a flare of Sgr~A* was observed with polarimetry, it was found that it is accompanied by significant polarization which varies on similarly short timescales as the light curve itself. By carrying out an analysis only on the total flux, some pieces of the observed information are ignored. It is already known that polarimetric data have been shown to be able to reveal substructure in flares, even when the light curve appears largely featureless (e.g. see Fig. 4 in Eckart et al. 2006b). The other main advantage of polarimetric observations is that, in addition to the flux density light curve, one can analyze the changes in the observed degree of polarization and the changes in polarization angle during flaring time as well. The claimed quasi-periodicity has been interpreted as being related to the orbital time scale of the matter in the inner parts of the accretion disk. According to well-known observed high frequency quasi-periodic oscillations (HFQPOs) in X-ray light curves of stellar mass black holes and binaries (Nowak \\& Lehr 1998), this interpretation is of special interest since it shows a way to better understand the behavior of accretion disks for a wide range of black hole masses. The recent unambiguous discovery of a one hour (quasi-)periodicity in the X-ray emission light curve of the active galaxy RE J1034+396 provides further support to this idea and extends the similarity between stellar-mass and super-massive black holes to a new territory (Gierli\\'{n}ski et al. 2008). Although the origin of the observed QPOs in the sources associated with black holes is still a matter of debate, several magnetohydrodynamic (MHD) simulations confirmed that it could be related to instabilities in the inner parts of the accretion disks, very close to the marginally stable orbit of the black hole ($r_{mso}$), and also possibly connected with the so-called \"stress edge\" (Hawley 1991; Chan et al. 2009b). If the flux modulations are related to a single azimuthal compact over-dense region (hereafter: \"hot spot\"), orbiting with the same speed as the underlying accretion disk, one can constrain the spin of the black hole by connecting the observed time scales of QPOs to the orbital time scale of matter around the black hole: $T = 2.07 (r^{\\frac{3}{2}}+a)(\\frac{M}{4\\times10^6M_\\odot})$~min (Bardeen et al. 1972, where $ -1\\leq a\\leq1$ is the black hole dimensionless spin parameter and $r$ is the distance of the spot from the black hole). The characteristic behavior of general relativistic flux modulations produced via the orbiting hot spots have been discussed in several papers (see e.g. Cunningham \\& Bardeen 1973; Abramowicz et al. 1991; Karas \\& Bao 1992; Hollywood et al. 1995; Dov\\v{c}iak 2004, 2007). In this paper, we have used a spotted accretion disk scenario to model the observed patterns of our sample of NIR light curves. Several authors have proposed different models in order to explain the flaring activity of Sgr~A*. These models cover a wide range of hypotheses like disk-star interactions (Nayakshin et al. 2004), stochastic acceleration of electrons in the inner region of the disk (Liu et al. 2006), sudden changes in the accretion rate of the black hole (Liu et al. 2002), heating of electrons close to the core of a jet (Markoff et al. 2001; Yuan et al. 2002), trapped oscillatory modes in the inner regions of the accretion disk in the form of spiral patterns or Rossby waves (Tagger \\& Melia 2006; Falanga et al. 2007; Karas et al. 2008), non-axisymmetric density perturbations which emerge as the disk evolves in time (Chan et al. 2009b), non-Keplerian orbiting spots falling inward inside the plunging region created via magnetic reconnections (Falanga et al. 2008), and also comet like objects trapped and tidally disrupted by the black hole (\\v{C}ade\\v{z} et al. 2006; Kosti\\'{c} et al. 2009). Observational data render some of these models unlikely. The star-disk interaction model is unable to produce the repeated flux modulations and also the high degree of polarization since it mainly deals with thermal emission. Tidal disruption of comet-like objects are also unable to reproduce the observed rate of flares per day, since the estimated capture rate of such objects for the Sgr~A* environment is at least one order of magnitude lower (\\v{C}ade\\v{z} et al. 2006). Nevertheless, several viable models exist and make different predictions that can be distinguished observationally. For example, one important characteristic prediction of hot spot models is about the wobbling of the center of the images (Broderick \\& Loeb 2006a,b; Paumard et al. 2006; Zamaninasab et al. 2008b; Hamaus et al. 2009). Significant effort has been already devoted to measure this possible position wander of Sgr~A* in the mm, sub-mm and NIR regimes (Eisenhauer 2005b, 2008, Gillesen 2006, Reid et al. 2008, Doelleman et al. 2008). In this paper we discuss how NIR polarimetry and the next generations of VLT interferometry (VLTI) and Very Long Baseline Interferometry (VLBI) experiments can provide data to support or reject certain models for the accretion flow/outflow related to Sgr~A*. Obtaining accurate data on the accretion flow of Sgr~A* can lead us to a better understanding of the physics of strong gravitational regimes, formation of black holes and their possible relation to the galaxy formation process in a cosmological context. In Sect. 2 we present a complete sample of NIR light curves observed in the polarimetry mode. A brief description about the details of the observation and data reduction methods is provided. We discuss the quasi-periodicity detection methods and present the results of a correlation analysis between the flux and polarimetric parameters. A general description of our model setup and results of simulations are discussed in Sect. 3. We show how NIR polarimetry can be used as a way to constrain physical parameters of the emitting region (like its geometrical shape) in Sect. 4 and Sect. 5. In Sect. 6 we mainly discuss the predictions that the future NIR interferometer (GRAVITY) is expected to reveal and how different assumptions in the model parameters can modify the results. In Sect. 7 we summarize the main results of the paper and draw our conclusions. ", "conclusions": "We used a sample of NIR flares of Sgr~A* observed in polarimetry mode to study the nature of the observed variability. Using the z-transformed discrete correlation function algorithm, we found a significant correlation between changes in the measured polarimetric data and total flux densities. This provides evidence that the variations probably originate from inner parts of an accretion disk while a strong gravity's effects are manifested inside them. In order to obtain this information polarization data is indispensable since the corresponding signals are very difficult or even impossible to extract from simple total intensity light curves only (Do et al. 2008). The presence of significant signals from orbiting matter then calls for detailed modeling of the NIR polarized light curves in order to analyze the distribution of the emitting material and the magnetic field structure within the accretion disk (e.g. Eckart et al. 2006a, 2008a, Meyer et al. 2006ab,2007). In addition, the pattern recognition algorithm we employed in this paper is an efficient tool to search for flare events that carry the signature of strong gravity in light curves which are significantly longer than the orbiting time scale over which such an event can typically occur. Other methods that make simultaneous use of the entire light curve (like e.g. the Lomb-Scargle algorithm) tend to dilute these signals and lower the possibility for significant detections dramatically (Do et al. 2008, Meyer et al. 2009). In order to constrain the physical properties of the emitting region we employed a relativistic disk model with azimuthal over-densities of relativistic electrons. A combination of a synchrotron mechanism and relativistic amplifications allows us to fit the real observed data and make predictions about astrometric parameters of the accretion disk around Sgr~A*. The modeled light curves show the same correlation between the flux and polarimetric data as the one deduced from observations. Simulations have been carried out in a way that they cover a wide range of parameters, including the effects of gravitational shearing inside the accretion disk, the heating and cooling time scales, the inclination and the spin of the black hole. It is discussed how the observed swings in the polarization angle support the idea of a compact source for the emission, instead of radially extended spiral shapes. Furthermore, we present a model in which the observed NIR polarization angle can lead to confining the expected region for the expected outflow/wind from Sgr~A*. The model also predicts that when observations will be able to resolve the position of such an outflow, the magnetic field structure inside the accretion disk could be confined. Finally, the centroid paths of the NIR images are discussed. In comparison with the results by Broderick \\& Loeb (2006a,b) and Hamaus et al. (2009), we have shown that the geometrical structure of the emitting region (elongation of the spot according to the Keplerian shearing, multi-component structures, spiral arms, confusion caused by the radiation from the hot torus) can affect the expected centroid tracks. While all the mentioned geometries are able to fit the observed fluxes, we show how the future NIR interferometer GRAVITY on the VLT can break these degeneracies. The results of simulations propose that focusing GRAVITY observations on the polarimetry mode could reveal a clear centriod track of the spot(s). We conclude that even though a non-detection of centroid shifts can not rule out the multi-components model or spiral arms scenarios, a clear position wander in the center of NIR images during the flares will support the idea of bright long-lived spots orbiting occasionally around the central black hole. This possible detection opens a new window for testing the physics very close to the edge of infinity." }, "0911/0911.3401_arXiv.txt": { "abstract": "Previous solar system constraints of the Brans-Dicke (BD) parameter $\\omega$ have either ignored the effects of the scalar field potential (mass terms) or assumed a highly massive scalar field. Here, we interpolate between the above two assumptions and derive the solar system constraints on the BD parameter $\\omega$ for {\\it any} field mass. We show that for $\\omega=O(1)$ the solar system constraints relax for a field mass $m \\gsim 20 \\times m_{AU}= 20\\times 10^{-27}GeV$. ", "introduction": " ", "conclusions": "" }, "0911/0911.1859_arXiv.txt": { "abstract": "{We here propose the time drift of subtended angles as a new possible cosmological probe. In particular, with the coming era of microarcsecond astrometry, our proposal can be used to measure the Hubble expansion rate of our universe in a direct way.} \\begin{document} ", "introduction": "Research in cosmology has become extraordinarily lively in the past quarter century. In particular, the use of Type Ia supernovae as standard candles lead to the discovery that the expansion of our universe is accelerating, implying that the energy of the universe may be dominated by some sort of exotic matter, with a ratio of pressure to density less than one third, dubbed dark energy\\cite{Weinberg}. However, besides the cosmological constant, there are various other dark energy models devised to explain the mysterious accelerating expansion. On the other hand, the current accelerating expansion of our universe can also be accounted for by either the modification of Einstein gravity or the violation of Copernican principle. With the other cosmological probes, some of them has been ruled out, but there still remain a number of theoretical models surviving such observational tests\\cite{FTH}. In this era of empirical cosmology, it is thus significant to propose some new observational programs that may shed light on our understanding of the universe in a way independent from other known cosmological probes. By comparing and combining results from very different methods of determining cosmological parameters, we may both obtain stronger constraints than any method alone would impose, and test techniques against one another to identify signatures of systematic effects. In particular, in a field so afflicted by systematic errors as cosmology, having many independent but complementary techniques is the best way to ensure that we are on the right track to explore the genuine mechanism underlying the evolution of our universe. We here contribute to such a theme by proposing the time drift of subtended angles as a new cosmological probe. As depicted in Fig.\\ref{td}, the ordinary cosmological probes are usually related to the sky survey along the past light cone of today. However, we have another way to acquire the evolution information of our universe by measuring the time drift of cosmological objects. The classical exemplification is Sandage-Loeb test\\cite{Sandage,Loeb}. In next section we shall popularize the spirit of time drift by confining ourselves onto the case of subtended angles and argue that the time drift of subtended angle may provide us with a new cosmological probe, just like Sandage-Loeb test. Some discussions will be presented in the last section. \\begin{figure} \\includegraphics[width=12cm,height=10cm] {td.eps}\\\\ \\caption{Time drift versus sky survey in the redshift space.}\\label{td} \\end{figure} ", "conclusions": "We have argued that the time drift of subtended angles may be used as a new promising cosmological probe to measure the Hubble expansion rate of our universe in a direct way. However, so far the analysis of its feasibility rests on an order of magnitude estimate. It is thus important to see whether the realistic data will distort such an estimate dramatically. On the other hand, although it is hard to imagine that the virialized systems will undertake a size evolution to the same order as the Hubble expansion, it may be expected that at least a few of them will violate this general belief. Nevertheless we may employ the other side of the same coin by instead using Eq.(\\ref{coin}) to determine the size evolution rate for those exceptional objects and non-virialized ones if we know the evolution law of our universe from other cosmological probes, and vice versa, since the subtended angle by definition entangles the whole universe with those luminous objects in it. It is interesting to ask whether we can disentangle them intrinsically. One way is to combine it with the time drift of the redshift difference from the far and near points of the object to obtain an evolution free cosmological probe, reminiscent of Alcock-Paczynski test\\cite{AP}. Another way out is to go directly for the time drift of angle diameter distance, which by expression is immune to the size evolution effect totally. All of these issues are worthy of further investigation but beyond the scope of this paper, we expect to report them elsewhere in the near future." }, "0911/0911.1297_arXiv.txt": { "abstract": "We study relativistic stars in the context of scalar tensor theories of gravity that try to account for the observed cosmic acceleration and satisfy the local gravity constraints via the chameleon mechanism. More specifically, we consider two types of models: scalar tensor theories with an inverse power law potential and f(R) theories. Using a relaxation algorithm, we construct numerically static relativistic stars, both for constant energy density configurations and for a polytropic equation of state. We can reach a gravitational potential up to $\\Phi\\sim 0.3$ at the surface of the star, even in f(R) theories with an ``unprotected\" curvature singularity. However, we find static configurations only if the pressure does not exceed one third of the energy density, except possibly in a limited region of the star (otherwise, one expects tachyonic instabilities to develop). This constraint is satisfied by realistic equations of state for neutron stars. ", "introduction": "One of the most challenging tasks for cosmology and fundamental physics today is to try to understand the apparent acceleration of the Universe. Beyond the minimal assumption of a pure cosmological constant, two main approaches have been explored. The first one consists in assuming some unknown form of matter, called {\\it dark energy}, characterized by an equation of state $P\\simeq -\\rho$. The second approach is more radical, as it tries to explain the present observations as the manifestation of a modified theory of gravity, which mimicks general relativity on solar system scales, but significantly deviates from it on cosmological scales. It turns out that it is rather difficult to construct a theory of gravity that, while being internally consistent, can account for the cosmological observations and be compatible with the present gravity constraints deduced from laboratory experiments and from solar and astrophysical systems. A class which has attracted a lot of attention is the so-called $f(R)$ gravity theories where the standard Einstein-Hilbert gravitational Lagrangian, proportional to the scalar curvature $R$, is replaced by a function of $R$ while the matter part of the Lagrangian is left unchanged (see e.g. \\cite{review} for a recent review). After several detours, it has been realized that viable $f(R)$ theories must satisfy stringent conditions in order to avoid instabilities and to satisfy the present laboratory and solar system constraints, and a few models have been carefully constructed to meet these requirements \\cite{Hu:2007nk,Appleby:2007vb,Starobinsky:2007hu}. To explore the full viability of these theories, it is important to go one step further by studying their behaviour in the strong gravity regime, such as reigns in the core of the most relativistic stars, namely neutron stars. In this context, it has been claimed in \\cite{Frolov,KM1} that very relativistic stars do not exist in the models \\cite{Hu:2007nk,Appleby:2007vb,Starobinsky:2007hu} because of the presence of an easily accessible singularity. In a recent work \\cite{BL1}, we have shown that this claim does not hold by constructing numerically static relativistic stars. Note that relativistic stars have also been studied in \\cite{Kainulainen:2007bt} but within different $f(R)$ models. $f(R)$ models can also be seen as a subclass of scalar-tensor theories. In particular, viable $f(R)$ models rely on the so-called chameleon mechanism. For this reason, it is interesting to extend the study of relativistic stars to chameleon models. In this work, we show that the behaviour of chameleon and $f(R)$ models is quite similar. In particular, in both cases, the scalar field in the innermost part of the star sticks to the minimum of its effective potential (if this minimum exists). If the equation of state is such that $\\rho-3P<0$, which occurs in the central part of highly relativistic constant energy density stars, then there is no minimum. As we show here, it is nevertheless possible to construct numerically static stars, up to some critical value of the central energy density. We believe that this is due to tachyonic instabilities, associated with a negative effective squared mass, which develop and prevent the existence of a static star configuration. However, this problem does not apply to realistic neutron stars: although there is a large uncertainty on the equation of state deep inside a neutron star, the equations of state that have been proposed in the literature verify $\\rho-3P>0$ throughout the star. To construct a simple approximation of a realistic neutron star, we have used a polytropic equation of state. { Although, according to our previous work \\cite{BL1} and the present one, the singularity of the models \\cite{Hu:2007nk,Appleby:2007vb,Starobinsky:2007hu} does not seem so far to be an obstacle for relativistic stars, it appears to be problematic for cosmology~\\cite{Frolov}.} This has motivated the construction of regularized versions of these models by adding for instance an extra $R^2$ term~ \\cite{Starobinsky:2007hu,Capozziello:2009hc,Dev:2008rx,Abdalla:2004sw,Thongkool:2009js}. A more sophisticated model was also proposed recently in~\\cite{Appleby:2009uf}. We will consider these ``cured'' $f(R)$ theories and compare their behaviour in relativistic stars with their ``singular'' counterparts. This paper is organized as follows. In the next section, we derive the main equations governing static and spherically symmetric configurations in scalar-tensor theories. Section III is devoted to relativistic stars in chameleon models. We then consider, in Section IV, the $f(R)$ models, including the ``cured'' models that have recently been introduced to solve the cosmological singularity problem of some of these models. We finally conclude in Section V. ", "conclusions": "We have studied in this work relativistic stars in scalar tensor and $f(R)$ theories that use the chameleon mechanism. The behaviour of the scalar field is extremely similar in the two types of models. The main properties are the following. Deep inside the star, the scalar field follows very closely the minimum of its effective potential (if it exists). As a consequence, if the minimum changes as a function of the radius, the scalar field will closely follow it. If $\\trho-3\\tP<0$, which can occur for instance in the central region of very compact stars with constant energy density, there is no minimum for the effective potential. It is however possible to find numerical solutions for constant energy stars with both $w<1/3$ and $w>1/3$ regions, but only up to some maximum gravitational potential. We have given qualitative arguments to show that one expects tachyonic instabilities if most of the star is characterized by $w>1/3$. As already emphasized in our previous work \\cite{BL1}, our results invalidate the claim that highly relativistic stars cannot be constructed in $f(R)$ theories. Numerically, the construction of a relativistic stars in $f(R)$ gravity is a challenging task because the scalar field value is extremely close to the singularity in the center of the star. We have also constructed relativistic stars in the so-called ``cured'' $f(R)$ models, which have been advocated to solve a cosmological singularity problem. We have also illustrated how the screening effect of the chameleon mechanism manifests itself for relativistic stars, by computing numerically the effective coupling of the star as well as the post-Newtonian parameter $\\tilde\\gamma$. \\vskip 1cm {\\bf Note:} while this paper was being completed, another paper~\\cite{Cooney:2009rr} on relativistic stars in $f(R)$ theories appeared on the arXiv, although in a different context. \\vskip 1cm" }, "0911/0911.5361_arXiv.txt": { "abstract": "The misalignment between the orbital plane of a transiting exoplanet and the spin axis of its host star provides important insights into the system's dynamical history. The amplitude and asymmetry of the radial-velocity distortion during a planetary transit (the Rossiter-McLaughlin effect) depend on the projected stellar rotation rate $v\\sin I$ and misalignment angle $\\lambda$, where the stellar rotation axis is inclined at angle $I$ to the line of sight. The parameters derived from modelling the R-M effect have, however, been found to be prone to systematic errors arising from the time-variable asymmetry of the stellar spectral lines during transit. Here we present a direct method for isolating the component of the starlight blocked by a planet as it transits the host star, and apply it to spectra of the bright transiting planet HD 189733b. We model the global shape of the stellar cross-correlation function as the convolution of a limb-darkened rotation profile and a gaussian representing the Doppler core of the average photospheric line profile. The light blocked by the planet during the transit is a gaussian of the same intrinsic width, whose trajectory across the line profile yields a precise measure of the misalignment angle and an independent measure of $v\\sin I$. We show that even when $v\\sin I$ is less than the width of the intrinsic line profile, the travelling Doppler ``shadow'' cast by the planet creates an identifiable distortion in the line profiles which is amenable to direct modelling. Direct measurement of the trajectory of the missing starlight yields self-consistent measures of the projected stellar rotation rate, the intrinsic width of the mean local photospheric line profile, the projected spin-orbit misalignment angle, and the system's centre-of-mass velocity. Combined with the photometric rotation period, the results give a geometrical measure of the stellar radius which agrees closely with values obtained from high-precision transit photometry if a small amount of differential rotation is present in the stellar photosphere. ", "introduction": "The recent discoveries of strong projected spin-orbit misalignments in the transiting exoplanets XO-3b (\\citealt{hebrard2008_xo3_rm}, \\citealt{winn2009xo3_rm}), HD 80606b (\\citealt{winn2009hd80606rm}, \\citealt{pont2009hd80606rm}) and WASP-14b \\citep{johnson2009wasp-14rm}, and retrograde orbital motion in WASP-17b \\citep{anderson2009wasp-17} and HAT-P-7b (\\citealt{winn2009hat-p-7rm}; \\citealt{narita2009hat-p-7rm}) indicate violent dynamical histories for a significant fraction of the population of hot-Jupiter planets. These misalignments are measured by obtaining densely-sampled radial-velocity observations during transits. The starlight blocked by the planet during a transit possesses the radial velocity of the obscured part of the stellar surface. The radial-velocity centroid of the light emanating from the visible parts of the stellar disk exhibits a reflex motion disturbance --the Rossiter-McLaughlin (RM) effect -- whose form depends on the projected stellar equatorial velocity $v\\sin I$, the impact parameter $b$ of the planet's path across the stellar disc, the projected misalignment angle $\\lambda$ between the orbital axis and the stellar spin axis, and the relative sizes of the star and planet. Detailed semi-analytic formulations for determining the anomalous departure of the line centroid from the stellar reflex orbit have been developed by \\citet{ohta2005rm}, \\citet{gimenez2006rm} and \\citet{hirano2009rm}. The process of measuring and interpreting this reflex velocity shift is not, however, straightforward. If the stellar rotation profile is even partially resolved by the spectrograph, the missing light within the planet's silhouette introduces an asymmetry into the line spread function. For instruments such as SOPHIE and HARPS, radial velocities are derived using a gaussian fit to a cross-correlation function (CCF) computed from the stellar spectrum and a weighted line mask (\\citealt{baranne96elodie}; \\citealt{pepe2002harps}). Such methods work extremely well for the symmetric and time-invariant profile shapes encountered outside transit. When applied to profiles with a time-varying degree of intrinsic asymmetry, however, they yield velocity measures that depart systematically from the velocity centroid of the visible starlight. \\citet{winn2005rm} overcame this difficulty for their iodine-cell observations of HD 209458b using a direct modelling approach which they have employed in subsequent papers. They built models of both the unobscured stellar profile and the light blocked by the planet into their analysis of the line-spread function, deriving semi-empirical corrections to the model velocities to enable meaningful comparison with the data. Nonetheless, their study of the RM effect in HD 189733 \\citep{winn2006hd189733} shows a clear pattern of correlated radial-velocity residuals during the transit. In a more recent HARPS study of the RM effect during a transit of HD 189733b, \\citet{triaud2009rm} noted an almost identical systematic pattern of correlated residuals between the radial velocities of the gaussian fits to the CCFs, and the line centroid velocities of the best-fitting model computed using the formulation of \\citet{gimenez2006rm}. In developing an analytic model of this velocity anomaly for cross-correlation spectra, \\citet{hirano2009rm} have shown that such errors are exacerbated as $v\\sin I$ increases and the asymmetry becomes more extreme. If uncorrected, the anomaly can lead to significant over-estimation of the value of $v\\sin I$, particularly for rapidly-rotating planet-host stars. The motivation for the present study is to develop a comprehensive model of the changes in CCF morphology that occur before, during and after a transit, in terms of the projected stellar equatorial rotation speed, the spin-orbit misalignment angle, the intrinsic width of the combined stellar CCF and instrumental broadening function, and the stellar limb-darkening coefficient. In principle this approach is similar to that of \\citet{winn2005rm}, but we use it to decompose the observed CCF into its various components and track them directly. A similar methodology has already been used to model the RM effect in the binary stars V1143 Cyg \\citep{albrecht2007} and DI Her \\citet{albrecht2009}. To illustrate the effectiveness of this approach as an alternative to gaussian-fitting in the study of transiting exoplanets, we re-analyse the same HARPS observations of HD 189733b described by \\citet{triaud2009rm}. We show that the spectral signature of the light blocked by the planet is clearly discernible in the residuals when a model of the unobscured starlight is subtracted from the data, and track it directly to derive new measurements of the projected stellar rotation velocity and spin-orbit misalignment angle. ", "conclusions": "In principle we could determine a lower limit on the stellar radius from $v\\sin I$ and the photometric rotation period, which was determined by \\citet{henry2008hd189733} to be $11.953\\pm 0.009$ days. The very close alignment of the projected orbital and spin axes in the plane of the sky makes it highly probable that the inclination $I$ of the stellar spin axis to the line of sight should be very close to the inclination $i=85.5$ degrees of the planet's orbit. Assuming $\\sin I = \\sin i = 0.997$, our value for $v\\sin I$ yields a direct estimate of the stellar radius $R_*=0.732\\pm 0.007$ R$_\\odot$, independently of the usual assumptions involving isochrone fits to the stellar density and effective temperature. This is significantly smaller than the $R_*=0.77\\pm 0.01$ R$_\\odot$ determined by \\citet{triaud2009rm}, and marginally less than the $R_*=0.75\\pm 0.01$ R$_\\odot$ found by \\citet{pont2007hst}. We suspect that the discrepancy arises from differential rotation of the stellar photosphere. \\citet{gaudi2007rm} pointed out that under some circumstances, the RM effect could be used to measure stellar differential rotation. As with most previous studies of this kind, we have measured $v\\sin I$ under the assumption that the star rotates as a solid body. In all stars for which differential rotation has been measured directly, however, a solar-like pattern is found: the equator rotates faster than the poles, with the departure from the equatorial spin rate having a sine-squared dependence on latitude. The typical difference in rotational angular frequency $\\Delta\\Omega = \\Omega_{\\rm equator}-\\Omega_{\\rm pole}$ is found to be of order 0.04 radian per day for rapidly-rotating stars K dwarfs similar to HD 189733, and only weakly dependent on rotation rate \\citep{barnes2005diffrot}. This is very similar to the solar value of $\\Delta\\Omega$. If the starspots giving rise to the photometric modulation signal are concentrated at the same stellar latitude as the planet's path across the stellar disc, we should derive the correct value of the stellar radius. Unfortunately we have no way of determining the dominant latitude from which the modulation signal originates. If the orbital and spin axes are closely aligned, the impact parameter of the planet's path across the star carries it across a stellar latitude of 39$^\\circ$ in the hemisphere facing away from the observer. Although \\citet{pont2007hst} noted the passage of the planet across one or two isolated dark spots, these spots are significantly foreshortened. It is reasonable to expect the modulation signal to arise mainly from spots at intermediate to low latitude in the observer's hemisphere. A difference in latitude of 20 or 30 degrees is sufficient to give a difference in angular velocity $\\delta\\Omega\\simeq 0.015$ radians per day between the main spot belts and the latitudes traversed by the planet. This discrepancy yields a stellar radius that is too small by about 3 percent, which is sufficient to reconcile our radius estimate with those of \\citet{pont2007hst} and \\citet{triaud2009rm}. We conclude that direct modelling of the cross-correlation function during a planetary transit is feasible even for a host star such as HD~189733 whose $v\\sin I$ is comparable to the intrinsic line width. The profile decomposition method described here yields values and error estimates for the stellar spin rate and orbital obliquity that agree closely with the results of previous studies of this system. The parameter values are free of the systematic errors that occur when velocity measurements are derived from gaussian fits to line profiles with inherent time-variable asymmetry. This circumvents the need for semi-empirical corrections when modelling the Rossiter-McLaughlin effect. The method makes the fullest use of all the information present in the cross-correlation function: $v\\sin I$ is tightly constrained by the shape of the stellar rotation profile as well as by the spectral signature of the starlight blocked by the planet. The stellar radius obtained from the photometric rotation period and the stellar $v\\sin I$ agrees well with values obtained from high-precision transit photometry, if a modest amount of differential rotation is present. The method described here could in principle be extended to observations for which an iodine cell is used to track the spectrograph PSF and wavelength scale. The analysis of such observations requires a reference spectrum of the host star, taken with the same instrument without the iodine cell. An artificial reference spectrum could be generated by convolving a limb-darkened rotation profile with the the spectrum of a narrow-lined star of the same spectral type. The Doppler signature of the light blocked by the planet could be mimicked by scaling, shifting and subtracting the same narrow-lined spectrum from the broadened stellar spectrum; the result would then be used as the reference spectrum. A very similar procedure was described by \\citet{winn2005rm} for calibrating the departure of the velocities measured during a transit of HD 209458b from the predictions of the \\citet{ohta2005rm} model, using high-resolution solar spectra." }, "0911/0911.0896_arXiv.txt": { "abstract": " ", "introduction": " ", "conclusions": "\\label{sec5} In the last decade ground based as well as satellite observations probed with greater accuracy the nature of the magnetic field of the sun. The SOHO \\cite{soho}, TRACE \\cite{trace}, GONG \\cite{gong} experiments \\footnote{ SOHO stands for Solar and Heliospheric Observatory; TRACE stands for TRansition Region and Coronal Explorer; GONG stands for Global Oscillation Network Group.} helped in deepening solar observations. In spite of these direct probes, it is fair to say that the 22-year magnetic cycle of the Sun cannot be claimed to be fully understood in terms of the so-called $\\alpha\\omega$ mechanism. The GONG observations give, for instance, the profile of the angular velocity of the sun not only on the surface but also in the interior of the sun. In spite of this valuable determination other effects may play a role so that it is fair to say that the situation is not clear \\cite{sun}. The sun is the closest and better observed astrophysical object and still the features of dynamo amplification are under debate. In the case of large-scale magnetism we face a similar situation. It seems that we will probably never be able to know the initial conditions of the galactic dynamo as accurately as we pretend to understand the initial conditions of solar dynamos where, still, crucial puzzles remain. The hopes of clarifying this problem in the near future might not be so forlorn. Indeed, as repeatedly argued, CMB anisotropies and polarization \\cite{mga,dyn} are a powerful window on pre-decoupling physics and, consequently, on pre-decoupling magnetic fields. Waiting for more direct observational evidences it is plausible to speculate that a moderate dynamo action combined with compressional amplification could indeed bridge the regime of the initial conditions with the observed large-scale magnetic fields. Even in this simplified framework important theoretical problems persist and they have to do with the way the Universe becomes a good conductor at the end of inflation. The early variation of the gauge couplings is a potential candidate for producing large-scale magnetic seeds for galaxies and clusters. However, more effort is certainly required especially in modeling the finite conductivity effects and the self-consistent evolution of the whole system of equations. \\newpage" }, "0911/0911.4967.txt": { "abstract": "{ We consider the application of quantum corrections computed using renormalization group arguments in the astrophysical domain and show that, for the most natural interpretation of the renormalization group scale parameter, a gravitational coupling parameter $G$ varying $10^{-7}$ of its value across a galaxy (which is roughly a variation of $10^{-12}$ per light-year) is sufficient to generate galaxy rotation curves in agreement with the observations. The quality of the resulting fit is similar to the Isothermal profile quality once both the shape of the rotation curve and the mass-to-light ratios are considered for evaluation. In order to perform the analysis, we use recent high quality data from nine regular disk galaxies. For the sake of comparison, the same set of data is modeled also for the Modified Newtonian Dynamics (MOND) and for the recently proposed Scalar Tensor Vector Gravity (STVG). At face value, the model based on quantum corrections clearly leads to better fits than these two alternative theories.} ", "introduction": "\\label{intro} The dynamics of many galaxies looks like as if the main part of their masses is distributed in a different way than their light is, with heavier densities than those expected from the emitted radiation by the stars and gas. Curiously, although galaxies are far from being ``hairless\", like fundamental particles or black holes, these masses discrepancies follow considerably strong patterns, including a series of correlations with the luminous matter, e.g. \\cite{urc, btf}. Usually, in order to solve this missing matter problem, a qualitatively new kind of matter, dark matter (DM), is evoked. Among all kinds of dark matter, current cosmological observations (like the large scale structure and the microwave background anisotropies) favor the collisionless and ``cold'' type of dark matter (CDM). However the CDM paradigm faces some difficulties at galactic scales, like the angular momentum issues on galaxy formation, the discrepancies on the number of satellite galaxies, and the cuspy dark matter density profile (see \\cite{primack} for a recent review). The CDM paradigm has not yet yielded a satisfactory dark matter profile for spiral galaxies from the simulations of cosmological evolution. The problem with the cuspy dark matter distribution in galaxies was already pointed in \\cite{flores, moore1994}, and in particular both the Navarro-Frenk-White \\cite{nfw} and the Moore \\cite{moore} dark matter profiles have a cusp on the dark matter density at the galactic center whose dynamical consequences do not match the observations; the discrepancies with the observations becoming particularly clear for the Low Surface Brightness galaxies. The problematics associated with this cuspy dark matter profile, and possible solutions, were systematically analyzed in many references, including \\cite{blokrubin, spano, Gentile, ThingsRot, ohthings, Donato}, see \\cite{corecuspblok} for a recent review. It is well known that the parameters of the dark matter profiles that best fit the observed rotation curves (like the Isothermal \\cite{BBS} or the Burkert\\footnote{Considering the CDM simulations, the Burkert profile has merits over the Isothermal one, in particular for its large $R$ behavior and its finite total mass; see also \\cite{burkertCDM}. } \\cite{burkert} ones, which have a core and were proposed on phenomenological grounds) display many correlations with the baryonic matter behavior, e.g. \\cite{mcgaughdmlsb, corecor, Donato, BBS, burkert}, see also \\cite{Cthesis}. These correlations between a galaxy rotation curve and its baryonic matter suggest another interpretation to the missing mass problem in galaxies, namely a change of the gravitational law itself instead of introducing DM. This has lead to diverse proposals, including modifications on the law of inertia \\cite{mond} and its extensions to General Relativity (e.g., \\cite{revteves}), proposals with extra-dimensions \\cite{extrad}, actions of General Relativity with an extra affine connection or a new graviton \\cite{ebi}, conformal gravity \\cite{Mannheim}, non-symmetric metrics \\cite{BrownMoffat}, among others possibilities. MOND \\cite{mond, sandersmcgaugh} is by far the most studied example of the last approach. Although it seems to be consistent with the general features of the galactic dynamics, its concordance with observations (if the original recipe is applied) started to be questioned in the light of the higher precision recent observations and the improvements on the stellar mass-to-light ratios constraints, e.g. \\cite{Gentile, m33m31, tevesgalaxy, mondmilky, mondearlydisk}. One can suppose that the deviation from the gravitational (either Einstein or Newton) law is due to the quantum effects, such as semiclassical corrections, effects of quantum gravity, consequences of string theory physics, extra dimensions, branes or some other (maybe yet unknown) model of ``quantum gravity''. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\vskip 1mm The corrections to the Newton law is a relatively common feature of the different models of ``quantum gravity'', including higher derivative quantum gravity \\cite{frts82}, different versions of the effective low-energy quantum theory of gravitational field \\cite{AntMot,don,Wood}, so-called non-perturbative quantum gravity based on the hypothesis of the existence of the non-Gaussian UV fixed point \\cite{Reuter} and, finally, in the semiclassical approach to quantum gravity \\cite{book} (see also references therein). The application of these corrections has been elaborated recently in the cosmological \\cite{nova,CCfit,Gruni} and astrophysical \\cite{Gruni} areas. One can see the recent paper \\cite{DCCR} for the detailed discussion of the quantum field theory backgrounds of these quantum effects. Let us emphasize, from the very beginning, that in the most cases the present day state of art in all mentioned approaches to quantum gravity does not enable one to really calculate the relevant quantum contributions to the Newton law in a unique and consistent way. At the same time, we have many reasons to believe that these quantum corrections can be non-zero and, in principle, may have a measurable effect. Typically, the theoretical estimate for the quantum contribution to the gravitational law involves more or less strong arbitrariness, related to the dependence of the quantum corrections on some dimensional parameter (scale parameter in the renormalization group-based approaches, for instance) from one side, and to the physical interpretation of this parameter from the other side. In the present paper we will consider a relatively general model of quantum corrections which is mainly controlled by covariance. Furthermore we will use a new identification of the scale parameter $\\mu$ in the astrophysical setting (rotation curves problem), which we believe to be the more natural compared to the one \\ ($\\mu \\sim 1/r$) \\ considered before in \\cite{Bertol,Gruni,Reuter} (see also \\cite{Mazzitelli-94}). It turns out that this new identification of scale leads to the surprisingly good result for the rotation curves of the galaxies, which is very close in quality to the output of the mainstream approach based on the DM paradigm. The previous papers on the subject were based on the renormalization group equations coming from the higher-derivative quantum gravity \\cite{Bertol} (unfortunately, the proper renormalization group equations \\cite{frts82} are ambiguous in this case, making their application to astrophysics doubtful \\cite{nova}), on the assumption of the Appelquist and Carazzone decoupling of the quantum corrections in the low-energy domain \\cite{babic,nova,CCfit} and on the assumption of the asymptotic safety and existence of the non-Gaussian fixed point in the four-dimensional quantum gravity \\cite{Reuter}. It is remarkable that all those, in fact rather different approaches, converge in predicting qualitatively similar logarithmic running of the effective Newton constant with the scale $\\mu$. It was shown in \\cite{Gruni} (see also previous qualitatively similar consideration in \\cite{Bertol}) that this logarithmic behaviour, together with the mentioned above identification of $\\mu^{-1}$ to the distance from the center of the galaxy, $r$, can partially explain the main features of the rotation curves without invoking the DM. The astrophysical applications of \\cite{Bertol}, \\cite{Gruni} and \\cite{Reuter} were performed for the point-like model of the galaxy, which is not supposed to be realistic. It was pointed out in \\cite{Gruni} that it would be very interesting to make a more detailed investigation, taking the case of a plane or thin disk distribution of mass in the galaxy. We present the corresponding analysis and show that it meets certain obstacles at both theoretical and phenomenological levels. In order to make a conclusive consideration of the astrophysical application of the renormalization group method, one has to ask whether the $\\mu \\sim 1/r$ identification is the unique possible one, or there are some alternative options. One can remember that the identification of scale in the cosmological setting is a nontrivial issue, which was treated differently by different authors in different papers \\cite{CCcosm,nova,babic,CCfit,Reuter-t,Reuter}. Furthermore, was an interesting attempt to construct the regular scale-fitting procedure \\cite{Guberina-scale}. The output of this procedure was the most natural identification of $\\mu$ with the Hubble parameter $H$, which is the energy characteristic of the external metric leg of the Feynman diagram, corresponding to the quantum correction to the expansion of the universe. Indeed, similar energy characteristic can be constructed in the case of the galaxies rotation curves, and it is indeed different from $1/r$. Moreover, this ``right'' setting of the energy scale leads to the immediate and dramatic improvement of the phenomenological analysis results. After all, we gain a chance to explain the rotation curves on the basis of quantum corrections to the classical action of gravity, without using DM. The paper is organized as follows. In section 2 we discuss very general features of quantum contributions and establish their likely form, which is based on the general covariance and on the proper existence of these contributions, following \\cite{Gruni}. As we have already mentioned above, this existence can not be either proved or disproved at the present-day state of art in this area \\cite{DCCR} and it is legitimate to apply a phenomenological approach. Let us note that the covariance enables one to put serious restrictions on the possible forms of quantum corrections in the cosmological setting, the astrophysical ones can be deduced starting from the cosmological setting result and the assumption of a unique effective action for gravity. The main new element in section 2 is the new definition of the energy scale, which is appropriate for the application to the rotation curves problem. In section 3 we present the numerical results of our model applied to the data of nine regular galaxies divided into two samples of data. For comparison purposes, our results are shown together with the numerical results of three other proposals, obtained from exactly the same data and procedures. These three proposals are: dark matter (modeled with the Isothermal profile \\cite{BBS}), MOND \\cite{mond} and STVG \\cite{stvg, BrownMoffat}. The first is the leading candidate for the missing matter problem in galaxies and all the universe altogether, MOND is the paradigmatic model for galaxy rotation curves without dark matter and STVG is a recently proposed model for astrophysics and cosmology without dark matter that also uses the variation of the gravitational coupling parameter $G$ to fit galaxy rotation curves. Finally, in section 4 we draw our Conclusions and discuss the perspectives of the theory of quantum corrections (especially the ones based on the renormalization group) as a potential competitor of the standard $\\Lambda$CDM model, at both astrophysical and cosmological scales. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{conclusions} We presented a model, motivated by quantum corrections to the Einstein-Hilbert action, that introduces small inhomogeneities in the gravitational coupling across a galaxy (of about 1 part in $10^7$) and can generate galaxy rotation curves in agreement with the observational data, without the introduction of dark matter as a new kind of matter. Considering the samples of galaxies evaluated in this paper, the quality of the rotation curve fits and the associated mass-to-light ratios from our model is a bit lower, or about the same, than the Isothermal profile quality, but with one less free parameter. We also compared the results of our model with MOND \\cite{mond} and STVG\\cite{stvg,BrownMoffat}, and at face value our model yielded clearly better results. Our results can be seen as a next step compared to the previous models motivated by renormalization group effects in gravity, \\cite{Gruni, Reuter:2007de}. Their original analyses could only yield a rough estimate on the galaxy rotation curves, since they were restricted to modeling a galaxy as a single point. Trying to extend this approach to real galaxies, we have shown that the phenomenologically optimized proper scale for the renormalization group phenomenology is not of a geometric type, like the inverse of the distance, but is proportional to the Newtonian potential with null boundary condition at infinity. This setting is shown to be in agreement with both theoretical expectations and with observations. The essential feature for the rotation curve fittings is the formula (\\ref{vsvg}), which is by itself a remarkably simple formula that provides a very efficient description of galaxy rotation curves. We have shown that it can be derived from the assumption of a gravitational coupling parameter with a very small departure from the constant $G_0$, such that, in the weak field limit, it depends on the logarithm of the Newtonian potential, e.g. $G(\\Phi_\\Newt) = G_0 \\[ 1 - \\beta \\, \\ln (\\Phi_\\Newt / \\Phi_0) \\]$. In the last equation, $\\beta$ is a positive and very small ($\\sim 10^{-7}$) effective parameter that necessarily depends on the distribution of mass of the system. The precise relation between $\\beta$ (or $\\alpha \\nu$ using the theory of quantum corrections) with the galaxy parameters is yet to be unveiled. To this end, from a phenomenological approach, a larger sample of galaxies would be necessary to find meaningful results. On the other hand, using the Tully-Fisher law, it is not hard to guess that $\\beta$, for disk galaxies, should scale with the mass approximately as $\\sqrt{M}$, in which $M$ is the baryonic mass of the galaxy. The results for $\\alpha$ in tables \\ref{resultsA} and \\ref{resultsB}, clearly show that it indeed increases with the galactic mass. In order to avoid premature statements, we think that the nature of $\\beta$ (or $\\alpha$) should be explored in more detail in the forthcoming papers. In the present paper we have addressed the issue of generating galaxies rotation curves by developing and extending the proposal of \\cite{Gruni}. It turns out that, at least for the sample galaxies we dealt with here, one can explain these curves without invoking the dark matter concept. Does it mean that we can really be free of dark matter in constructing a realistic cosmological scenario? The answer to this question is probably negative. It is well known that there many other issues associated with dark matter, or the lack of it, that are necessary to take into account. In particular, the results on the density perturbations and related issues such as cosmic microwave background radiation (CMBR) and the large scale structure (LSS) data, baryon acoustic oscillations, big bang nucleosynthesis, gravitational lensing and others (see, e.g. \\cite{Ross} for a recent review). In all these observational and experimental issues the standard way of explaining the date is to assume the existence of dark matter. At the same time, one can not underestimate the fact that the rotation curves may be explained without dark matter effect. One can imagine, for instance, the scenario with essentially smaller amount of dark matter which has slightly different set of properties compared to the usual CDM model. In this case the rotation curves will be explained by summing up the effect of quantum corrections and the one of the dark matter content. The preliminary analysis shows this possibility can not be ruled out \\cite{AToJFa}. Finally, there is a chance to address all mentioned issues trading a large amount of cold dark matter content by another one, which can have qualitatively other origin, and possibly invoking the quantum corrections to the cosmological perturbations spectrum (see, e.g., \\cite{den, GCC}). \\vspace{.2in} {\\it Note added.} After the first version of this work was finished, we became aware, through \\cite{brownstein}, and contrary to the expectations in \\cite{stvg}, that there are small and perhaps sensible differences between the fits of the rotation curves using either the STVG or the MSTG formulations. If precision on the nomenclature and results is at stake, our results on the model developed by Moffat and collaborators apply to the older formulation, namely MSTG. \\vspace{.2in}" }, "0911/0911.1224_arXiv.txt": { "abstract": "\\noindent We study the linear differential equation $\\dot{x} = Lx$ in $1$:$1$ resonance. That is, $x \\in \\fR^4$ and $L$ is $4 \\times 4$ matrix with a semi-simple double pair of imaginary eigenvalues $(i\\beta,-i\\beta,i\\beta,-i\\beta)$. We wish to find all perturbations of this linear system such that the perturbed system is stable. Since linear differential equations are in one to one correspondence with linear maps we translate this problem to $\\gl(4,\\fR)$. In this setting our aim is to determine the stability domain and the singularities of its boundary. The dimension of $\\gl(4,\\fR)$ is $16$, therefore we first reduce the dimension as far as possible. Here we use a versal unfolding of $L$ ie a transverse section of the orbit of $L$ under the adjoint action of $\\Gl(4,\\fR)$. Repeating a similar procedure in the versal unfolding we are able to reduce the dimension to $4$. A $3$-sphere in this $4$-dimensional space contains all information about the neighborhood of $L$ in $\\gl(4,\\fR)$. Considering the $3$-sphere as two $3$-discs glued smoothly along their common boundary we find that the boundary of the stability domain is contained in two right conoids, one in each $3$-disc. The singularities of this surface are transverse self-intersections, Whitney umbrellas and an intersection of self-intersections where the surface has a self-tangency. A Whitney stratification of the $3$-sphere such that the eigenvalue configurations of corresponding matrices are constant on strata allows us to describe the neighborhood of $L$ and in particular identify the stability domain. ", "introduction": "\\label{sec:intro} \\subsection{Setting}\\label{sec:introset} Suppose that the ordinary differential equation \\begin{equation}\\label{eq:orig} \\dot{x} = F_{\\mu}(x) \\end{equation} has a stationary point at $x=0$ for all $\\mu$. In this equation $x \\in \\fR^n$, $\\mu \\in \\fR^p$ and $F_{\\mu}(x)\\frac{\\partial}{\\partial x}$ is a vector field on $\\fR^n$ smoothly depending on $x$ and $\\mu$. Furthermore suppose that $A: \\fR^p \\to \\gl(n,\\fR): \\mu \\mapsto A(\\mu) = DF_{\\mu}(0)$ is the linear part of the vector field at $x=0$. We are interested in the case that the phase space is $4$-dimensional and the linear part $A(\\mu)$ has a double semi-simple pair of complex conjugate imaginary eigenvalues $(i\\beta,-i\\beta,i\\beta,-i\\beta)$, $\\beta \\neq 0$, at $\\mu = 0$. A system with such a linear part is also said to be in 1:1-\\emph{resonance}. With $I$ denoting a two by two identity matrix and $\\beta=1$ we have \\begin{equation}\\label{eq:l} A(0) = L = \\begin{pmatrix} 0 & I\\\\-I & 0 \\end{pmatrix}. \\end{equation} Our aim is to give a description of a small neighborhood of $L$ in $\\gl(4,\\fR)$ and in particular find the stability domain and its boundary. On the latter we expect bifurcations of the stationary point of the original differential equation \\eqref{eq:orig}. Thus this study fits in a much larger study of local bifurcations of systems having a stationary point with zero or imaginary eigenvalues in the linear part. We give a short list of generic bifurcations with increasing number of zero or imaginary eigenvalues in the linear part. We do not specify non-degeneracy conditions, but instead refer to the literature. In dimension one, only one eigenvalue can be zero and we have \\emph{transcritical} (TC) and \\emph{saddle-node} (SN) bifurcations. However in the presence of symmetry the \\emph{pitchfork} (PF) bifurcation occurs. In dimension two there can be two non semi-simple zero eigenvalues denoted by $0^2$, then the system has a \\emph{Bogdanov-Takens} (BT) bifurcation. But there can also be a pair of imaginary eigenvalues, then the system undergoes a \\emph{Hopf} bifurcation. If the two dimensional system is Hamiltonian and has two zero eigenvalues $0^2$, then there is in general a \\emph{Hamiltonian saddle-node} (SSN) bifurcation. When the dimension gets larger, results get sparser because the codimensions of the bifurcations readily increase. In dimension three, the simplest case is the \\emph{Hopf-saddle-node} (HSN) bifurcation when there is one zero eigenvalue and an imaginary pair, denoted $0\\beta$. Other cases are systems with linearizations having eigenvalues $0^3$ or $00^2$. The bifurcations for these cases are not yet well studied. The simplest bifurcation for general systems in dimension 4 is the \\emph{Hopf-Hopf} (HH) bifurcation where the linearization at the stationary point has two pairs of \\emph{non-resonant} imaginary eigenvalues $\\beta_1\\beta_2$. Other results for $4$-dimensional systems are known for special cases only. For Hamiltonian systems there is a bifurcation at $k$:$l$-resonance of imaginary eigenvalues $\\beta_1\\beta_2$. A special place is taken by the 1:$\\pm 1$-resonances, where the sign in 1:$\\pm 1$ designates \\emph{symplectic signature} of eigenvalues. At $1$:$-1$-resonance there is a so called \\emph{Hamiltonian-Hopf} (SH) bifurcation. But at $1$:$1$-resonance there is another bifurcation that has not been analyzed completely. The $1$:$1$-resonance in reversible systems is very similar to the $1$:$-1$-resonance in Hamiltonian systems, note that there is no signature for imaginary eigenvalues in reversible systems. However, there is in reversible systems a \\emph{reversible sign} for real and zero eigenvalues, see \\cite{hvn96}. The case $0_+^4$ has been studied, but the case $0_-^4$ has not, as far as we know. We summarize the local bifurcations (with lowest codimension) occurring in systems up to dimension 4 in table \\ref{tab:bifs}. \\begin{table}[htbp] \\begin{center} \\begin{tabular}{|c|c|l|c|l|l|}\\hline dim & evc & bif & codim & comments & references\\\\\\hline 1 & $0$ & SN & 1 & & \\cite{gh, kuz}\\\\ & & TC & 1 & ``$x=0$ stationary for all $\\mu$'' & \\cite{gh, kuz}\\\\ & & PF & 2 & & \\cite{gh, kuz}\\\\ & & PF & 1 & $\\fZ_2$-symmetric & \\cite{gh, kuz}\\\\\\hline 2 & $0^2$ & BT & 2 & & \\cite{drsz}\\\\ & $\\beta$ & H & 1 & & \\cite{gh, kuz}\\\\ & $0^2$ & SSN & 1 & symplectic & \\cite{wig}\\\\\\hline 3 & $0\\beta$ & HSN & 2 & & \\cite{bv, gh, gpd}\\\\\\hline 4 & $\\beta_1\\beta_2$ & HH & 2 & $\\beta_1 : \\beta_2$ irrational & \\cite{gh, kuz}\\\\ & $\\beta^2$ & SH & 1 & symplectic 1:$-1$-resonance & \\cite{mee}\\\\ & $\\beta_1\\beta_2$ & nn & 1 & symplectic $1$:$2$-resonance & \\cite{dui}\\\\ & $\\beta_1\\beta_2$ & nn & 3 & symplectic $1$:$3$-resonance & \\cite{dui}\\\\ & $\\beta_1\\beta_2$ & nn & 2 & symplectic $k$:$l$-resonance, $k$:$l \\neq 1$:$2$, $k$:$l \\neq 1$:$3$ & \\cite{dui}\\\\ & $\\beta^2$ & nn & 1 & reversible 1:1-resonance & \\cite{msv}\\\\ & $\\beta_+\\beta_+$ & nn & nk & symplectic 1:1-resonance & \\cite{hh}\\\\ & $\\beta\\beta$ & nn & nk & 1:1-resonance &\\\\ & $0_+^4$ & nn & 2 & reversible, reversible sign +1 & \\cite{ioo}\\\\\\hline \\end{tabular} \\end{center} \\caption{\\textit{Low codimension bifurcations of stationary points in systems up to dimension 4. We do not list the non-degeneracy conditions. When such conditions are violated but higher order conditions are met, we generally get a higher codimension bifurcation for the same eigenvalue configuration. The abbreviations have the following meaning dim: dimension of phase space; evc: eigenvalue configuration (see section \\ref{sec:stabdombkar}); bif: name of bifurcation; codim: codimension of bifurcation; SN: saddle-node; TC: transcritical; PF: pitchfork; BT: Bogdanov-Takens; SSN: Hamiltonian saddle-node; HSN: Hopf-saddle-node; HH: Hopf-Hopf; SH: Hamiltonian Hopf; nn: no name; nk: not known.}\\label{tab:bifs}} \\end{table} Here we do not even attempt to describe the bifurcation of system \\eqref{eq:orig}, whose linear part at $x=0$ and $\\mu=0$ is in 1:1-resonance. Instead we focus on the linearized system $\\dot{x}=A(\\mu)x$ near $\\mu=0$. As mentioned before we will describe the neighborhood of $A(0)=L$ in $\\gl(4,\\fR)$, in particular the stability domain and the singularities on its boundary. Moreover it is of both theoretical and practical interest to know how linear Hamiltonian, reversible and equivariant subsystems appear as subspaces in $\\gl(4,\\fR)$ and how they intersect the stability domain and its boundary near $L$. This last issue will be treated in more detail in \\cite{hkb}. The \\emph{stability domain} in $\\gl(4,\\fR)$ is defined as follows. \\begin{equation}\\label{eq:stabdom} \\stabdom = \\{A \\in \\gl(4,\\fR)\\;|\\; \\text{if $\\lambda$ is an eigenvalue of $A$ then $\\Re(\\lambda) < 0$}\\} \\end{equation} Then the boundary of the stability domain $\\dstabdom$, is characterized by vanishing real parts of one or more eigenvalues of $A \\in \\gl(4,\\fR)$. By the implicit function theorem, $\\dstabdom$ as a hyper surface in $\\gl(4,\\fR)$, is smooth in points where the corresponding matrix has a simple zero eigenvalue or a simple complex conjugate pair with vanishing real part. At points where multiple eigenvalues have vanishing real parts we may expect singularities. A simple example being two pairs of purely imaginary eigenvalues $\\pm i \\beta_1$ and $\\pm i \\beta_2$ with $\\beta_1 \\neq \\beta_2$. Then $\\dstabdom$ has generically a transverse self-intersection. We will only consider the stability domain and its boundary in a small neighborhood of $L$ and in figure \\ref{fig:eigcfgs} we list the possible \\emph{eigenvalue configurations}. For an informal definition of eigenvalue configuration see section \\ref{sec:statres}, a precise definition will be given in section \\ref{sec:stabdom}. We use the following coding: $\\gamma$ for a pair of complex conjugate eigenvalues; $\\beta$ for a pair of complex conjugate imaginary eigenvalues; $\\alpha$ for a real eigenvalue; $0$ for a zero eigenvalue; $\\gamma \\gamma$ for semi-simple double eigenvalues and $\\gamma^2$ for double eigenvalues with nilpotent part of height 2. A subscript $\\pm$ denotes the sign of the real part and an optional index is used to denote different eigenvalues. \\begin{figure}[htbp] \\setlength{\\unitlength}{1mm} \\begin{picture}(100,25)(0,0) \\put(10, 15){\\includegraphics{eigcfgs.ps}} \\put(2, 1){$\\beta\\beta$} \\put(24, 1){$\\beta^2$} \\put(42, 1){$\\beta_1\\beta_2$} \\put(60, 1){$\\beta\\gamma_-$} \\put(82, 1){$\\gamma_{-1}\\gamma_{-2}$} \\put(104, 1){$\\gamma_-\\gamma_+$} \\put(124, 1){$\\beta\\gamma_+$} \\put(142, 1){$\\gamma_{+1}\\gamma_{+2}$} \\end{picture} \\caption{\\textit{Eigenvalue configurations near $L$. $\\beta\\beta$ is the eigenvalue configuration of $L$. On $\\dstabdom$ we have $\\beta^2$, $\\beta_1\\beta_2$ or $\\beta\\gamma_-$, on $\\stabdom$ we only have $\\gamma_{-1}\\gamma_{-2}$ and elsewhere we have $\\gamma_-\\gamma_+$, $\\beta\\gamma_+$ or $\\gamma_{+1}\\gamma_{+2}$.}\\label{fig:eigcfgs}} \\end{figure} \\begin{remark}\\label{rem:randcfgs} There are many more eigenvalue configurations on $\\dstabdom$, including zero or real eigenvalues, for example $\\alpha_{-1}\\alpha_{-2}\\beta$, $0\\alpha_-\\beta$ and $0^4$. In points of $\\dstabdom$ where the corresponding matrix has eigenvalue configuration $\\alpha_{-1}\\alpha_{-2}\\beta$, $\\dstabdom$ is smooth, for $0\\alpha_-\\beta$ it has a self-intersection and for $0^4$, $\\dstabdom$ will be more singular. In a study of $0^4$ the present study of $L$ will appear as a sub case. However, these eigenvalue configurations do not occur on an arbitrary small neighborhood of $L$. \\end{remark} The question of singularities on the boundary of the stability domain has been taken up earlier, see for example a discussion of the \\emph{decrement diagram} in \\cite{arn1}. There a classification is guided by codimension that is by the number of parameters in a family of matrices $A(\\mu)$ where $A(0)$ is the central singularity. Also see \\cite{le82,ms99} for an elaboration on this idea with examples. Here we wish to view a study of $L$ in a classification guided by dimension of the phase space. This will lead to high codimensions. Indeed, in studying $L$ we have to consider an eight parameter family. See section \\ref{sec:methrcu} how we reduce such families. In the end it turns out that we have to study a three parameter family. In this family we do find most of the singularities listed in \\cite{arn1} for generic three parameter families. \\subsection{Motivation and main questions}\\label{sec:intromoq} The main motivation for this study comes from a wide variety of applications where the question of stability of a system near 1:1-resonance turns up in various forms. For example, double semi-simple imaginary eigenvalues are natural in the spectra of rotationally and spherically symmetrical models of solids and fluids \\cite{ki09}, ranging from car brakes \\cite{ki08}, rotating shafts \\cite{nn98}, and computer hard discs \\cite{cb92} to rotating elastic Earth \\cite{rv09}, and from vortex tubes \\cite{hf08} to magneto-hydrodynamics \\cite{kgs09}. Double semi-simple eigenvalues are also characteristic of optimal structures and are responsible for high sensitivity of the latter to small imperfections \\cite{slo94}. Many \\emph{dissipation-induced instabilities} in water wave models can be traced back to the occurrence of double imaginary eigenvalues, see \\cite{br97} and references therein, but also \\cite{mk91,bkmr94,mo95,km07,ki07,bmr08}. An early observation of friction induced instability can be found in \\cite{zie}, which has been related to a singularity on the boundary of the stability domain by \\cite{bot}. For more applications and references see \\cite{hkb}. The questions in the applications mentioned above in many instances boil down to questions about the stability domain and its boundary in $\\gl(4,\\fR)$ near a matrix with double semi-simple imaginary eigenvalues. The main questions we wish to address here are the following. \\begin{enumerate}\\itemsep 0pt \\item What are the open domains in parameter space with constant eigenvalue configuration? \\item What are the singularities on the boundaries of these domains? \\item In particular, we address the above two questions for the stability domain. \\commentaar{volgende naar artikel B} \\end{enumerate} We will consider these questions for a general system. But in many applications such a system can be considered as small, dissipative perturbations of a Hamiltonian system in $1:\\pm1$-resonance. However, also reversibility and equivariance with respect to a circle group play a prominent role. In all our constructions to study the neighborhood of a matrix with double semi-simple imaginary eigenvalues, we take care to carry them out in such a way that Hamiltonian, reversible and equivariant subsystems can be recognized easily. \\subsection{Organization}\\label{sec:introorg} In section \\ref{sec:statres} we give an overview of the results, that is a description of the domains with constant eigenvalue configuration near $L$ in $\\gl(4,\\fR)$. In particular we present the stability domain and its singularities. The methods we use will be outlined in section \\ref{sec:methods}. We give a short overview of the centralizer unfolding of a matrix $A$ in $\\gl(n,\\fR)$, a family describing a neighborhood of $A$. We also present a method to reduce the number of parameters in this family leading to a reduced centralizer unfolding with the same properties but easier to analyze. This will be applied in section \\ref{sec:unfol} to the matrix $L$ introduced in section \\ref{sec:introset}. The resulting reduced centralizer unfolding of $L$ is further analyzed in section \\ref{sec:stabdom} where we characterize the stability domain and its singularities. Finally in section \\ref{sec:evcnbhdl} we give a description of a small neighborhood of $L$ in $\\gl(4,\\fR)$. ", "conclusions": "" }, "0911/0911.3221_arXiv.txt": { "abstract": "{Information on physical characteristics of astrometric radio sources, magnitude and redshift in the first place, is of great importance for many astronomical studies. However, data usually used in radio astrometry is incomplete and often outdated.} {Our purpose is to study the optical characteristics of more than 4000 radio sources observed by the astrometric VLBI technique since 1979. Also we studied an effect of the asymmetry in the distribution of the reference radio sources on the correlation matrices between vector spherical harmonics of the first and second degrees.} {The radio source characteristics were mainly taken from the NASA/IPAC Extragalactic Database (NED). Characteristics of the gravitational lenses were checked with the CfA-Arizona Space Telescope LEns Survey. SIMBAD and HyperLeda databases was also used to clarify the characteristics of some objects. Also we simulated and investigated a list of 4000 radio sources evenly distributed around the celestial sphere. We estimated the correlation matrices between the vector spherical harmonics using the real as well as modelled distribution of the radio sources.} {A new list of physical characteristics of 4261 astrometric radio sources, including all 717 ICRF-Ext.2 sources has been compiled. Comparison of our data of optical characteristics with the official International Earth Rotation and Reference Systems Service (IERS) list showed significant discrepancies for about half of 667 common sources. Finally, we found that asymmetry in the radio sources distribution between hemispheres could cause significant correlation between the vector spherical harmonics, especially if the case of sparse distribution of the sources with high redshift. We also identified radio sources having many-year observation history and lack redshift. This sources should be urgently observed at large optical telescopes.} {The list of optical characteristics created in this paper is recommended for use as a supplement material for the next International Celestial Reference Frame (ICRF) realization. It can be also effectively used for cosmological studies and planning of observing programs both in radio and optics.} ", "introduction": "Information on physical characteristics of the astrometric radio sources is important for planning of VLBI experiments and analysis of VLBI data to do a research in cosmology, kinematics of the Universe, etc. In particular, the primary mainspring to this work was a support of the investigation of the systematic effects in apparent motion of the astrometric radio sources observed by VLBI (Gwinn et al. 1997, MacMillan 2005, Titov 2008a, Titov 2008b). The official list of the physical characteristics of the ICRF radio sources is supported by the IERS (International Earth Rotation and Reference Systems Service) ICRS (International Celestial Reference System) Product Center (Archinal et al. 1997). The latest version of the IERS list is available in the Internet\\footnote{http://hpiers.obspm.fr/icrs-pc/info/car\\_physique\\_ext1}. However this list has some deficiencies: \\begin{itemize} \\item Not all the sources observed in the framework of geodetic and astrometric experiments are included in the IERS list. \\item The characteristics of some sources in the IERS list are outdated or doubtful. \\end{itemize} To overcome this problems, we performed a compilation of new list of the physical characteristics of astrometric radio sources using the latest information. Hereafter this list is referred to as OCARS (Optical Characteristics of Astrometric Radio Sources). The list of radio sources with their positions was originally taken from the Goddard VLBI astrometric catalogues\\footnote{http://vlbi.gsfc.nasa.gov/solutions/astro}, version 2007c, with addition of two absent ICRF-Ext.2 (Fey et al. 2004) sources 1039-474 and 1329-665 {(\\bf hereafter we will use 8-char IERS designation HHMMsDDd which is an abridged version of the IAU-compliant name 'IERS BHHMMsDDd')}. In the last version, the source list was updated using the 2009a\\_astro.cat catalogue computed by Leonid Petrov\\footnote{http://astrogeo.org/vlbi/solutions}. It gives 4261 radio sources in total. At this stage mainly the NASA/IPAC Extragalactic Database\\footnote{http://nedwww.ipac.caltech.edu/} (NED) was scoured. Characteristics of the gravitational lenses were checked with the CfA-Arizona Space Telescope LEns Survey\\footnote{http://cfa-www.harvard.edu/glensdata/} (CASTLES). Several sources were checked with SIMBAD\\footnote{http://simbad.u-strasbg.fr/} the HyperLeda\\footnote{http://leda.univ-lyon1.fr/} databases. In the OCARS list we have included only the optical characteristics of astrometric radio sources: source type, redshift and visual magnitude. The flux parameters are not included in our list because they are available from other centers directly working on correlation and primary processing of the VLBI observations. The OCARS was preliminary presented in (Malkin \\& Titov 2008). In this paper we investigated statistical properties of the list in more detail and study their impact on the kinematic analysis of radio source motions. Analysis of the radio source apparent motion revealed some statistically significant systematics described by the vector spherical harmonics of the first and second orders (dipole and quadrupole effects, respectively) (Titov 2008a, Titov 2008b). The dipole effect could be caused by the Galactocentric acceleration of the Solar system (Gwinn et al. 1997; Sovers et al. 1998; Kovalevsky 2003; Klioner 2003; Kopeikin \\& Makarov 2006) or a hypothetic acceleration of the Galaxy relative to the reference quasars. The quadrupole harmonic, considered in details by Kristian \\& Sachs (1966), could be caused either of the primordial gravitational waves or anisotropic expansion of the Universe. This result was confirmed by Ellis et al. (1985) although they stated ``the major problem is that neither the distortion nor the proper motions are likely to be measurable in practice in the foreseeable future''. In this case the quadrupole effect should be redshift-dependent, and the apparent proper motion will increase with redshift. However, Pyne et al. (1996) and Gwinn et al. (1997) also discussed the gravitational waves with the wavelength shorter than the Hubble length. Thus, the proper motion, induced by the short-wavelength gravitational waves, also might be constant over all redshifts. Due to asymmetry of the astrometric radio source distribution around the sky, the correlation between the vector spherical harmonic components is not zero. Therefore, we studied the effect of the asymmetry using the real uneven and simulated even distribution of the sources. The OCARS list can be used as a supplement material for the second realization of the International Celestial Reference Frame (ICRF2), as well as a database for kinematic studies of the Universe and other related works, including scheduling of dedicated IVS (International VLBI Service for Geodesy and Astrometry, Schl\\\"ueter \\& Behrend 2007) programs. ", "conclusions": "To conclude, it is necessary to note that this `historic' deficit of the radio sources (and, additionally, the radio sources with measured redshift) might cause problems with further investigation of the hardly detectable systematic effects in the proper motion of the reference radio sources. Therefore, large observational projects for spectroscopy of the astrometric radio sources in the Southern hemisphere are very important. Nonetheless, some observations in the North hemisphere also need to be undertaken. Independently, the MASIV scintillation survey (Lovell et al. 2003, 2009) also demonstrates a highly significant dramatic decrease in the numbers of scintillators for redshifts in excess of z = 2. The lack of scintillation at high redshifts is clear evidence for an increase in the source angular sizes with increasing redshift. Such an increase may be cosmological in origin or may be a propagation effect of inter-galactic scattering (Lovell et al. 2009). To observe and study radio source in both frequency range (optical and radio) characteristics such as visual magnitude, redshift in optic and flux density in several radio bands have to be measured. We also need to be sure that the same source is observed by the optical and radio instruments. Due to possible misalignment between optical and radio positions the physical characteristics might help to solve the problem of identification. To chase this aim, a new list of optical characteristics of 4261 astrometric radio sources, OCARS, including all 717 ICRF-Ext.2 sources has been compiled. The OCARS list includes source type, redshift and visual magnitude (when available). Detailed comments are provided when necessary, which is especially useful in understanding of incomplete, contradictory and controversial astrophysical data. The OCARS may serve to various VLBI tasks, for instance: \\begin{itemize} \\item As a supplement material for the second ICRF realization ICRF2 (Ma 2008). \\item As a database for VLBI data analysis. \\item For planning of IVS and other observing programs, in order to enrich the observational history of the sources with reliable determined redshift. \\item For future link between optical (GAIA) and radio (ICRF) celestial reference frames, once the GAIA optical position of about 100,000 quasars will be available. \\end{itemize} We performed a detailed comparison of the OCARS list with the official IERS list, and found many discrepancies for about a half of common sources. This comparison showed that the IERS list seems to be outdated and should be used with care. Besides, the IERS list, being intended to provide physical characteristics of the IERS sources only, contains only a small fraction of the whole set of astrometric radio sources used nowadays. We also compared the OCARS with the newest Large Quasar Astrometric Catalogue LQAC, and found discrepancies which worth further investigating. Most of discrepancies seems to be a result of different object identification. This is only the first stage of our work. We are planning the following steps: \\begin{itemize} \\item To continue searching for the missing and checking out the ambiguous characteristics through literature and astronomical databases. \\item To organize photometric and spectroscopy observations of astrometric radio sources with large optical telescopes. In particular, such an observational program started at Pulkovo Observatory in 2008. Observations are being made on the 6-m BTA telescope of the Special Astrophysical Observatory in North Caucasus. \\end{itemize} The list of optical characteristics of astrometric radio sources presented in this paper is publicly available at \\mbox{\\tt www.gao.spb.ru/english/as/ac\\_vlbi/sou\\_car.dat} and is updated once in several months." }, "0911/0911.3198_arXiv.txt": { "abstract": "We propose to use alternative cosmic tracers to measure the dark energy equation of state and the matter content of the Universe [w$(z)$ \\& $\\Omega_m$]. Our proposed method consists of two components: (a) tracing the Hubble relation using HII galaxies which can be detected up to very large redshifts, $z\\sim 4$, as an alternative to supernovae type Ia, and (b) measuring the clustering pattern of X-ray selected AGN at a median redshift of $\\sim 1$. Each component of the method can in itself provide interesting constraints on the cosmological parameters, especially under our anticipation that we will reduce the corresponding random and systematic errors significantly. However, by joining their likelihood functions we will be able to put stringent cosmological constraints and break the known degeneracies between the {\\em dark energy} equation of state (whether it is constant or variable) and the matter content of the universe and provide a powerful and alternative route to measure the contribution to the global dynamics and the equation of state of {\\em dark energy}. A preliminary joint analysis of X-ray selected AGN clustering (based on the largest to-date XMM survey; the 2XMM) and the currently largest SNIa sample, the {\\em Constitution} set (Hicken et al.), using as priors a flat universe and the WMAP5 normalization of the power-spectrum, provides: $\\Omega_{\\rm m}=0.27\\pm 0.02$ and w$=-0.96\\pm 0.07$. Equivalent and consistent results are provided by the joint analysis of X-ray selected AGN clustering and the latest Baryonic Acoustic Oscillation measures, providing: $\\Omega_{\\rm m}=0.27\\pm 0.02$ and w$=-0.97\\pm 0.04$. ", "introduction": "We live in a very exciting period for our understanding of the Cosmos. Over the past decade the accumulation and detailed analyses of high quality cosmological data (eg., supernovae type Ia, CMB temperature fluctuations, galaxy clustering, high-z clusters of galaxies, etc.) have strongly suggested that we live in a flat and accelerating universe, which contains at least some sort of cold dark matter to explain the clustering of extragalactic sources, and an extra component which acts as having a negative pressure, as for example the energy of the vacuum (or in a more general setting the so called {\\em dark energy}), to explain the observed accelerated cosmic expansion (eg. Riess, et al. 1998; 2004; 2007, Perlmutter et al. 1999; Spergel et al. 2003, 2007, Tonry et al. 2003; Schuecker et al. 2003; Tegmark et al. 2004; Seljak et al. 2004; Allen et al. 2004; Basilakos \\& Plionis 2005; 2006; 2009; Blake et al. 2007; Wood-Vasey et al. 2007, Davis et al. 2007; Kowalski et al. 2008, Komatsu et al. 2008; Hicken et al. 2009, etc). Due to the absence of a well-motivated fundamental theory, there have been many theoretical speculations regarding the nature of the exotic {\\em dark energy}, on whether it is a cosmological constant, a scalar or vector fields which provide a time varying dark-energy equation of state, usually parametrized by: \\begin{equation} p_Q= {\\rm w}(z) \\rho_Q\\;, \\end{equation} with $p_Q$ and $\\rho_Q$ the pressure and density of the exotic dark energy fluid and \\be {\\rm w}(z)={\\rm w}_0 + {\\rm w}_1 f(z) \\;, \\label{eqstatez} \\ee with w$_0=$w$(0)$ and $f(z)$ an increasing function of redshift [ eg., $f(z)=z/(1+z)$] (see Peebles \\& Ratra 2003 and references therein, Chevalier \\& Polarski 2001, Linder 2003, Dicus \\& Repko 2004; Wang \\& Mukherjee 2006). Of course, the equation of state could be such that w does not evolve cosmologically. Two very extensive recent reports have identified {\\em dark energy} as a top priority for future research: \"Report of the Dark Energy Task Force (advising DOE, NASA and NSF) by Albrecht et al. (2006), and ``Report of the ESA/ESO Working Group on Fundamental Cosmology'', by Peacock et al. (2006). It is clear that one of the most important questions in Cosmology and cosmic structure formation is related to the nature of {\\em dark energy} (as well as whether it is the sole interpretation of the observed accelerated expansion of the Universe) and its interpretation within a fundamental physical theory. To this end a large number of very expensive experiments are planned and are at various stages of development, among which the {\\em Dark Energy Survey} (DES: {\\tt http://www.darkenergysurvey.org/}), the {\\em Joint Dark Energy Mission} (JDEM: {\\tt http://jdem.gsfc.nasa.gov/}), {\\em HETDEX} ({\\tt http://www.as.utexas.edu/hetdex/}), Pan-STARRS: {\\tt http://pan-starrs.ifa.hawaii.edu}, etc. Therefore, the paramount importance of the detection and quantification of {\\em dark energy} for our understanding of the cosmos and for fundamental theories implies that the results of the different experiments should not only be scrutinized, but alternative, even higher-risk, methods to measure {\\em dark energy} should be developed and applied as well. It is within this paradigm that our current work falls. Indeed, we wish to constrain the {\\em dark energy} equation of state using, individually and in combination, the Hubble relation and large-scale structure (clustering) methods, but utilizing alternative cosmic tracers for both of these components. From one side we wish to trace the Hubble function using HII galaxies, which can be observed at higher redshifts than those sampled by current SNIa surveys and thus at distances where the Hubble function is more sensitive to the cosmological parameters. The HII galaxies can be used as standard candles (Melnick, Terlevich \\& Terlevich 2000, Melnick 2003; Siegel et al. 2005; Plionis et al. 2009) due to the correlation between their velocity dispersion, metallicity and $H_{\\beta}$ luminosity (Melnick 1978, Terlevich \\& Melnick 1981, Melnick, Terlevich \\& Moles 1988), once we reduce significantly their distance modulus uncertainties, which at present are unacceptably large for precision cosmology ($\\sigma_\\mu\\simeq 0.52$ mag; Melnick, Terlevich \\& Terlevich 2000). Furthermore, the use of such an alternative high-$z$ tracer will enable us to check the SNIa based results and lift any doubts that arise from the fact that they are the only tracers of the Hubble relation used to-date (for possible usage of GRBs see for example, Ghirlanda et al. 2006; Basilakos \\& Perivolaropoulos 2008)\\footnote{ GRBs appear to be anything but standard candles, having a very wide range of isotropic equivalent luminosities and energy outputs. Nevertheless, correlations between various properties of the prompt emission and in some cases also the afterglow emission have been used to determine their distances. A serious problem that hampers a straight forward use of GRBs as Cosmological probes is the intrinsic faintness of the nearby events, a fact which introduces a bias towards low (or high) values of GRB observables and therefore the extrapolation of their correlations to low-$z$ events is faced with serious problems. One might also expect a significant evolution of the intrinsic properties of GRBs with redshift (also between intermediate and high redshifts) which can be hard to disentangle from cosmological effects. Finally, even if a reliable scaling relation can be identified and used, the scatter in the resulting luminosity and thus distance modulus is still fairly large.}. From the other side we wish to use X-ray selected AGN at a median redshift of $\\sim 1$, which is roughly the peak of their redshift distribution (see Basilakos et al. 2004; 2005, Miyaji et al. 2007), in order to determine their clustering pattern and compare it with that predicted by different cosmological models. Although each of the previously discussed components of our project (Hubble relation using HII galaxies and angular/spatial clustering of X-ray AGN) will provide interesting and relatively stringent constraints on the cosmological parameters, especially under our anticipation that we will reduce significantly the corresponding random and systematic errors, it is the combined likelihood of these two type of analyses that enables us to break the known degeneracies between cosmological parameters and determine with great accuracy the {\\em dark energy} equation of state (see Basilakos \\& Plionis 2005; 2006; 2009). Below we present the basic methodology of each of the two main components of our proposal, necessary in order to constrain the {\\em dark energy} equation of state. ", "conclusions": "" }, "0911/0911.3151_arXiv.txt": { "abstract": "Stellar population synthesis (SPS) provides the link between the stellar and dust content of galaxies and their observed spectral energy distributions. In the present work we perform a comprehensive calibration of our own flexible SPS (FSPS) model against a suite of data. These data include ultraviolet, optical, and near--IR photometry, surface brightness fluctuations, and integrated spectra of star clusters in the Magellanic Clouds (MCs), M87, M31, and the Milky Way (MW), and photometry and spectral indices of both quiescent and post--starburst galaxies at $z\\sim0$. Several public SPS models are intercompared, including the models of Bruzual \\& Charlot (BC03), Maraston (M05) and FSPS. The relative strengths and weaknesses of these models are evaluated, with the", "introduction": "\\label{s:intro} The spectral energy distribution (SED) of a galaxy contains a wealth of information regarding its star formation history, dust content, and chemical abundance pattern. These properties provide essential clues to the physical processes governing the formation and evolution of galaxies from high redshift to the present. It is therefore highly desirable to have a robust method for extracting the physical properties of galaxies from their SEDs. The process of translating observed SEDs into physical properties is, unfortunately, very challenging because it requires 1) an accurate understanding of all phases of stellar evolution, 2) well--calibrated stellar spectral libraries for converting stellar evolution calculations into measurable fluxes, 3) an initial mass function (IMF), specifying the weight given to each stellar mass, 4) detailed knowledge of the star--dust geometry in conjunction with an appropriate extinction curve; i.e., knowledge of the physical conditions of the interstellar medium (ISM). Each of these requirements depend on chemical composition, further compounding the problem. Combining these ingredients in order to predict the spectrum of a galaxy is known as stellar population synthesis (SPS), and has an extensive history \\citep[e.g.,][]{Tinsley76, Tinsley80, Bruzual83, Renzini86, Buzzoni89, Bruzual93, Worthey94, Maraston98, Leitherer99, Fioc97, Vazdekis99, Yi03, Bruzual03, Jimenez04, LeBorgne04, Maraston05, Schiavon07, Coelho07, Conroy09a, Molla09, Kotulla09}. In the past two decades enormous progress has been made on each of the requirements mentioned above. Yet, substantial uncertainties remain. A non--exhaustive list includes 1) the treatment of the core convective boundary in main sequence stars (i.e., convective core overshooting); 2) metallicity--dependent mass--loss along the red giant branch (RGB) and, relatedly, the morphology of the horizontal branch (HB); 3) the treatment of the thermally--pulsating asymptotic giant branch (TP--AGB) phase; 4) the spectral libraries, especially for M giants, TP--AGB stars, and stars at non--solar metallicities \\citep[e.g.,][]{Martins07}; 4) blue straggler (BS) stars and their ubiquity \\citep[e.g.,][]{Preston00, Li08a}; 5) the effects of binarism, which might be especially relevant for massive star evolution \\citep[e.g.,][]{Eldridge08} and BS stars; 6) the importance of rotation on massive star evolution \\citep[e.g.,][]{Meynet00}; 7) non--solar abundance ratios, which effect not only the stellar evolution calculations but also the stellar spectra \\citep[e.g.,][]{Coelho07}; 8) the unknown dependence of the IMF on ISM properties such as metallicity and pressure. These uncertainties can in many cases dramatically impact the ability to convert observables into physical properties and vice--versa \\citep[e.g.,][]{Charlot96a, Charlot96b, Yi03a, Gallart05, Maraston06, LeeHC07, Conroy09a, Muzzin09, Conroy09c}. Thankfully, there exists a wide array of data capable of constraining SPS models. By far the most common type of object used for comparison is the star cluster\\footnote{Herein both open and globular clusters will be referred to as star clusters.}. The approximately uniform age and metallicity of the stars within star clusters and the lack of internal reddening (except perhaps for very young clusters) affords direct comparison between them and the most basic ingredients in SPS --- the stellar evolution calculations and stellar spectral libraries. Comparisons to star clusters are either made in CMD space or in an integrated sense (i.e., the light from all the cluster stars are added together). The former technique provides a sensitive probe of the main--sequence including the turn-off point, sub--giant and red giant branches \\citep[e.g.,][]{Worthey94, Bruzual03, An09}, while the latter technique provides a more robust measure of the rarer brightest stars that dominate the integrated light from the cluster. It is the latter method of comparison that is more relevant if the goal of SPS modeling is to understand the integrated light from galaxies. For this reason, the integrated colors, surface brightness fluctuations (SBFs), and spectra of star clusters have been used extensively to compare and calibrate SPS models \\citep[e.g.,][]{Bruzual03, Gonzalez04, Maraston05, Cohen07, Cordier07, Pessev08, Marigo08, Koleva08, LeeHC09b}. There are two significant sources of concern when attempting to calibrate SPS models with star cluster data. The first concern is that the brightest stars, which dominate the integrated light, are rare, and thus stochastic effects must be carefully modeled. Recently, this issue has been addressed observationally by stacking clusters in bins of age in order to synthesize `superclusters' that are relatively unaffected by stochastic effects \\citep{Pessev06, Gonzalez04}. The second concern is less tractable, even in principle, and arises from the fact that the primary sources of star clusters for which ages and metallicities can be reliably estimated are the Milky Way (MW), M31, and the Large and Small Magellanic Clouds (LMC and SMC, respectively). This fact imposes severe restrictions on the region of the age--metallicity parameter space that can be constrained with star cluster data. For example, one of the most important regions of this parameter space for studying galaxies --- old and metal rich --- contains very few star clusters (except in the bulge of the MW, although high extinction diminishes the utility of star clusters there). In Local Group star clusters, a bin in age contains clusters of a relatively narrow range in metallicity owing to the simple fact that the metallicity of galaxies tends to increase with age. Because of these issues, entire galaxies are often also used to assess the accuracy of SPS models \\citep[e.g.,][]{Worthey94, Bruzual03, Maraston06, Eminian08}. Galaxies are not subject to the two concerns mentioned above, but using them to constrain SPS models can be very difficult because 1) galaxies contain stars of a range of ages and metallicities; and 2) starlight in galaxies is attenuated by interstellar dust. Nevertheless, if subsamples of galaxies are carefully constructed with known and regular properties, then galaxies can be used to assess the reliability of SPS models. The present work seeks to provide a comprehensive comparison between SPS models and observed star clusters and galaxies. Such a comparison is timely because of an abundance of new, high--quality data, including near--IR photometry and SBFs of star clusters in the LMC and SMC, UV photometry of clusters in M31 and M87, CMDs and integrated spectra of clusters in the MW, and optical and near--IR photometry and spectral indices for large samples of low--redshift galaxies. We will consider not only our own SPS model \\citep{Conroy09a} but also the commonly used models of \\citet[][BC03]{Bruzual03} and \\citet[][M05]{Maraston05}. Such a detailed comparison between these popular models has not been undertaken until now, and will thus provide a much--needed evaluation of the relative strengths and weaknesses of these models. In addition, the flexible nature of our own model will be exploited to explore the extent to which various uncertain phases of stellar evolution, including thermally--pulsating asymptotic giant branch stars (TP--AGB), blue HB stars, and post--AGB stars, can be constrained by the data. This task is critical if we are to have confidence in the derived physical properties of galaxies, such as stellar masses and star formation rates. We proceed as follows. $\\S$\\ref{s:data} contains a detailed description of the star cluster and galaxy data used to constrain the models, which are themselves described in $\\S$\\ref{s:model}. In $\\S$\\ref{s:calib} we present an extensive comparison between models and data. A discussion and summary are provided in $\\S$\\ref{s:disc} and $\\S$\\ref{s:sum}, respectively. The zero points of the star cluster UBVRIJHK magnitudes are in the Vega system; all other magnitudes are quoted in the $AB$ system \\citep{Oke83}. Where necessary, we adopt a flat $\\Lambda$CDM cosmology with $(\\Omega_m, \\Omega_\\Lambda,h)=(0.26,0.74,0.72)$. ", "conclusions": "\\label{s:sum} We now summarize our principle results. \\begin{itemize} \\item The latest stellar models from the Padova group cannot fit the near--IR photometry and surface brightness fluctuations of star clusters in the MCs. With our flexible SPS (FSPS) code we have modified the TP--AGB phase in the Padova isochrones to produce better fits to the data. The M05 models also fail to reproduce the data, both because the colors are too red and the age--dependence is incorrect. The BC03 and the FSPS model using the BaSTI isochrones fare well, although both models are somewhat too blue in the near--IR. A significant limiting factor in more accurate calibrations is the difficulty in constructing reliable photometry of MC clusters, owing to the fact that TP--AGB stars are luminous and rare. \\item At low metallicities (log$(Z/Z_\\Sol)<-0.5$) and old ages, all models predict similar $UBVRIJHK$ colors (the typical variation between models is $\\approx0.05$ mags). When compared to MW and M31 star clusters, the models are 0.2 mags too blue in $V-K$ at log$(Z/Z_\\Sol)>-1.0$, and 0.1 mags too red in $J-K$ at all $Z$. The origin of the discrepancies in the near--IR is unclear. Decreasing the temperature of the RGB by $\\sim200$K alleviates much of the disagreement. If this modification is isolated to the brightest stars, then such a modification could improve agreement with integrated photometry without compromising agreement in CMD--space. The FSPS model with the BaSTI isochrones performs worst at low metallicities and old ages of the models considered (although the disagreement is not dramatic), while the FSPS model with the Padova isochrones performs well. Multi--color CMDs of two metal--rich clusters, NGC 6791 and NGC 188, are well--fit by FSPS. Spectral indices of MW star clusters are generally well--fit by both FSPS and BC03, although the models predict D$_n4000$ strengths too large and H$\\delta_A$ strengths too weak compared to two log$(Z/Z_\\Sol)\\approx0.0$ clusters. \\item The ultraviolet photometry of the MW, M87, and M31 star clusters can be well--fit by FSPS because FSPS contains flexible treatments of the post--AGB and horizontal branch evolutionary phases. This flexibility is essential given the substantial scatter in the ultraviolet data at fixed metallicity, and given our inadequate theoretical understanding of these phases. The M05 and BC03 models perform less well because of their incomplete treatment of these advanced evolutionary phases. \\item SPS models are compared to $ugrzYJHK$ photometry of massive red sequence galaxies. The FSPS, BC03, and M05 models fare well in most respects, with a few exceptions: all models are too blue in $J-H$; the M05 model is somewhat too blue in $Y-K$; all models are far too red in $g-r$ and $u-g$. These conclusions hold under the assumption that red sequence galaxies are composed exclusively of old metal--rich stars with canonical evolutionary phases. The disagreement in the $ugr$ colors can be alleviated if some combination of young ($\\sim1$ Gyr) stars, metal--poor stars, and blue straggler stars are added in small amounts. \\item Optical spectral indices of massive galaxies are generally well--fit by the FSPS and BC03 models, with the important exceptions that both models predict D$_n4000$ in excess of observations and neither fit the observed locus of galaxies in the D$_n4000-$H$\\delta_A$ plane. These conclusions hold for the same assumptions noted above, and, as above, small additions of some combination of young stars, metal--poor stars, blue straggler and horizontal branch stars can remedy this disagreement. \\item The FSPS and BC03 models adequately describe the $grizYK$ colors of K+A, or `post--starburst' galaxies, while the M05 model performs less well. Such galaxies contain a large proportion of intermediate age ($0.52$ radio galaxies. Here we present Gemini-North NIFS Intregral Field Unit (IFU) observations of the [O~{\\sc iii}]$\\lambda$5007 emission from a $z\\approx$~2 ultraluminous infrared galaxy ($L_{\\rm IR}>10^{12}$~$L_{\\odot}$) with an optically identified Active Galactic Nucleus (AGN). The spatial extent ($\\approx$~4--8~kpc) of the high velocity and broad [O~{\\sc iii}] emission are consistent with that found in $z>2$ radio galaxies, indicating the presence of a large-scale energetic outflow in a galaxy population potentially orders of magnitude more common than distant radio galaxies. The low radio luminosity of this system indicates that radio-bright jets are unlikely to be responsible for driving the outflow. However, the estimated energy input required to produce the large-scale outflow signatures (of order $\\approx10^{59}$~ergs over $\\approx$~30~Myrs) could be delivered by a wind radiatively driven by the AGN and/or supernovae winds from intense star formation. The energy injection required to drive the outflow is comparable to the estimated binding energy of the galaxy spheroid, suggesting that it can have a significant impact on the evolution of the galaxy. We argue that the outflow observed in this system is likely to be comparatively typical of the high-redshift ULIRG population and discuss the implications of these observations for galaxy formation models. ", "introduction": "The most successful current models of galaxy formation invoke large-scale energetic outflows to explain many of the properties of local massive galaxies and the intragalactic medium (IGM; i.e.,\\ red optical colours of massive galaxies, steep optical galaxy luminosity function, the black-hole--spheroid mass relationship, metal-enrichment of the IGM; e.g.,\\ Silk \\& Rees 1998; Fabian 1999; King 2003; Benson et~al. 2003; Granato et~al. 2004; Di Matteo et~al. 2005; Springel et~al. 2005; Bower et~al. 2006; Croton et~al. 2006; Hopkins et~al. 2006). A key attribute of these simulated outflows is the injection of significant amounts of kinetic energy into the interstellar medium (ISM), which can inhibit and terminate star formation by either heating the ISM or ejecting the gas out of the gravitational potential of the host galaxy. Large-scale outflows can be powered by star formation and/or Active Galactic Nuclei (AGN) activity (e.g.,\\ Heckman et~al. 1990; Crenshaw et~al. 2003; Veilleux et~al. 2005), although most galaxy formation models predict that only AGN-driven outflows will be sufficiently energetic to have significant impact on the formation and evolution of massive galaxies. These large-scale outflows are expected to be particularly effective at $z\\approx$~2, when star formation in the most massive galaxies was in significant decline (e.g.,\\ Juneau et~al. 2005; Panter et~al. 2007; P{\\'e}rez-Gonz{\\'a}lez et~al. 2008; Damen et~al. 2009). Broadly speaking, AGN-driven outflows are kinematically energetic ``winds'' or ``jets'' that are initially launched close to the central black hole. The two main catalysts that have been proposed to power an AGN-driven outflow are (1) a radio jet or lobe, and (2) a radiatively driven wind. A radio-jet/lobe driven outflow will be most prevalent in radio-loud AGN but might only occur in a minority of the AGN population.\\footnote{The typical definition of a radio-loud AGN is ${\\nu}{L_{\\rm 5GHz}}$/${\\nu}{L_{\\rm 440nm}}\\simgt$~10 (e.g.,\\ Kellerman et al. 1989).} For example, $\\approx$~4--24\\% of the optically bright (radiatively strong) AGN population are radio loud, which appears to be a function of both redshift and AGN luminosity (e.g.,\\ Hooper et~al. 1995; Jiang et~al. 2007), although this may indicate the ``duty cycle'' of radio AGN activity. The fraction of optically faint (radiatively weak) radio-loud AGN fraction is a strong function of stellar mass (e.g.,\\ Best et~al. 2005; Smol{\\v c}i{\\'c} et~al. 2009). By comparison, a radiatively driven wind would be most effective in luminous AGNs with high mass-accretion rates, since the wind is directly driven by radiation pressure from the accretion disk. High signal-to-noise ratio X-ray and ultra-violet absorption-line spectroscopy have shown that high-velocity outflows are present in many AGNs, and may be a ubiqitious property of the AGN population at both low and high redshift (e.g.,\\ Crenshaw et~al. 1999: Chartas et~al. 2002, 2007a,b; Laor \\& Brandt 2002; Pounds et~al. 2003; Reeves et~al. 2003; Porquet et~al. 2004; Ganguly \\& Brotherton 2008; Gibson et~al. 2009). For example, at least $\\approx$~60\\% of unobscured AGNs show evidence for high-velocity outflows, and the maximum measured outflow velocity is a strong function of the AGN luminosity (e.g.,\\ Crenshaw et~al. 1999; Ganguly \\& Brotherton 2008). The estimated energy injection required to produce the X-ray absorption features (which sometimes suggest outflow velocities exceeding $v\\approx$~0.1$c$) is typically $\\approx$~0.1--1 of the AGN bolometric luminosity (e.g.,\\ Pounds et~al. 2003; Reeves et~al. 2003; Chartas et~al. 2007a; Pounds \\& Reeves 2009). However, X-ray variability studies and consideration of the required energy input to produce the high-velocity features, indicate that these outflow signatures must be produced in the vicinity of the accretion disc on $<$~1~pc scales (e.g.,\\ Crenshaw et~al. 2003; King \\& Pounds 2003). Therefore, while it is clear that energetic outflows are present close to the accreting black hole of AGNs, it is far from clear what impact these outflows will have on the gas and star formation in the host galaxy on $\\approx$~1--10~kpc scales. To directly test the impact that outflows have on the formation and evolution of galaxies, it is necessary to identify large-scale energetic outflows, particularly in high-redshift ($z\\simgt2$) massive galaxies where they are predicted to have been most prevalent. Optical spectroscopy of distant galaxies have revealed blueshifted absorption/emission lines in many systems, a signature of outflowing gas (e.g.,\\ Pettini et~al. 2001; Shapley et~al. 2003; Tremonti et~al. 2007). However, the most direct insight into the identification and interpretation of large-scale outflows is provided by spatially resolved spectroscopy (integral-field unit; IFU observations), which provide direct information about both the velocity {\\it and} extent of any outflowing gas (e.g.,\\ Swinbank et~al. 2005, 2006; Nesvadba et~al. 2006). Using rest-frame optical IFU observations of distant radio-loud AGNs (i.e.,\\ high-redshift radio galaxies, hereafter referred to as HzRGs), Nesvadba et~al. (2006, 2007a, 2008) showed that large-scale energetic outflows are present in at least a fraction of the high-redshift galaxy population. The key diagnostic that revealed an AGN-driven outflow in these systems was the presence of kinematically complex and extended [O~{\\sc iii}]$\\lambda$5007 emission;\\footnote{Emission from [O~{\\sc iii}] can be produced by a number of processes, including photo-ionisation and shocks (Osterbrock 1989) but it cannot be produced in dense environments, such as the broad-line region of AGN, without being collisionally de-excited (e.g.,\\ Osterbrock 1989; Robson 1996). The detection of broad (FWHM$\\simgt$~500--1000~km$^{-1}$) [O~{\\sc iii}] emission therefore indicates the presence of off-nuclear kinematic components.} see Holt et~al. (2008) for similar constraints on $z\\simlt0.6$ radio galaxies. Nesvadba et~al. (2008) showed that the extent of the broad [O~{\\sc iii}] emission is similar to that of the radio emission, providing evidence for a causal connection between the radio emission and a large-scale outflow. The estimated kinetic energy required to drive the outflows in these HzRGs is $\\approx$~1--40\\% of that potentially provided by the radio jet, indicating that they could be powered by mechanical energy from the radio jet. However, HzRGs are rare systems (co-moving space densities at $z\\approx$~2 of $\\Phi\\approx10^{-8}$--$10^{-7}$~Mpc$^{-3}$; e.g.,\\ Willott et~al. 1998), possibly representing the most massive galaxies in the most biased regions of the distant Universe (e.g.,\\ Stevens et~al. 2003; Seymour et~al. 2007). To fully assess the global impact of large-scale energetic outflows on the formation and evolution of distant galaxies, and to distinguish between the different mechanisms proposed to drive outflows, it is therefore necessary to search for large-scale energetic outflows in the more typical radio-quiet AGN population. In this paper we present Gemini-North NIFS IFU observations of the [O~{\\sc iii}] emission from a $z\\approx$~2 quasar (SMM J123716.01+620323.3, hereafter referred to as SMM~J1237+6203) that is hosted in an Ultraluminous Infrared (IR) Galaxy (ULIRG; $L_{\\rm IR}>10^{12}$~$L_{\\odot}$; e.g.,\\ Sanders \\& Mirabel 1996). ULIRGs (both at low and high redshift) are rapidly evolving galaxies undergoing intense dust-obscured star formation and AGN activity (e.g.,\\ Sanders \\& Mirabel 1996; Genzel et~al. 1998; Page et~al. 2001; Alexander et~al. 2005b; Stevens et~al. 2005; Schweitzer et~al. 2006; Coppin et~al. 2008; Lutz et~al. 2008; Pope et~al. 2008). Ultra-luminous infrared quasars are often believed to be a key phase in the evolution of a ULIRG, where energetic outflows are starting to shut down star formation in the host galaxy, before revealing an unobscured quasar (e.g.,\\ Sanders et~al. 1988; Canalizo \\& Stockton 2001; Granato et~al. 2004; Hopkins et~al. 2005a; Kawakatu et~al. 2006). It is possible that every massive galaxy in the local Universe underwent a ULIRG phase at some time over the past $\\approx$~13~Gyrs during which the galaxy spheroid and black hole were predominantly grown (e.g.,\\ Swinbank et~al. 2006; Alexander et~al. 2008; Tacconi et~al. 2008). The [O~{\\sc iii}] luminosity of SMM~J1237+6203 is comparable to the HzRGs studied by Nesvadba et~al. (2006, 2007a, 2008), suggesting similar intrinsic AGN luminosities. However, importantly, SMM~J1237+6203 is $\\approx$~3--4 orders of magnitude fainter at radio luminosities than the HzRGs. We identify the signatures of a large-scale energetic outflow in this ultraluminous infrared quasar, showing that these features are not restricted to the distant radio-loud AGN population. We estimate the energy input required to produce these features, assess what could be responsible for driving this large-scale outflow, and discuss the implications for galaxy formation models. We have adopted $H_{0}=71\\kms$, $\\Omega_{M}=0.27$ and $\\Omega_{\\Lambda}=0.73$; in this cosmology 1$''$ corresponds to 8.5\\,kpc at $z=2.0$. ", "conclusions": "We have presented Gemini-North NIFS IFU observations of the [O~{\\sc iii}]$\\lambda$5007 emission in a $z\\approx$~2 ultraluminous infrared galaxy hosting an optically identified AGN (SMM~J1237+6203). Our main findings are the following: \\begin{itemize} \\smallskip \\item The spatial extent ($\\approx$~4--8~kpc) of the high velocity (up-to $v\\approx$~300~km~s$^{-1}$) and broad (up-to $\\approx$~900~km~s$^{-1}$) [O~{\\sc iii}] emission is consistent with that found in $z>2$ radio galaxies, suggesting the presence of a large-scale energetic outflow in a galaxy population potentially orders of magnitude more common than $z>2$ radio galaxies. See \\S3, \\S4.1, \\& \\S4.3. \\smallskip \\item The estimated energy required to produce these large-scale outflow features ($\\approx$~[0.6--3]~$\\times10^{44}$~erg~s$^{-1}$) could be provided by a wind radiatively driven by the quasar ($\\approx$~[0.3--3]~$\\times10^{45}$~erg~s$^{-1}$) and/or supernovae winds from intense star formation ($\\approx3\\times10^{44}$~erg~s$^{-1}$), so long as the wind-gas coupling efficiencies are comparatively high (of order $\\approx$~10--100\\%). However, the low radio luminosity of this system indicates that radio-bright jets are unlikely to be responsible for driving the outflow, in contrast to $z>2$ radio galaxies. See \\S4.1 and \\S4.2. \\smallskip \\item The estimated energy injection required to drive the large-scale outflow is comparable to the estimated binding energy of the galaxy spheroid in SMM~J1237+6203, suggesting that the outflow can have a significant impact on the evolution of the galaxy. However, on the basis of the maximum energy input into the outflow and the measured outflow velocities, it seems unlikely that the majority of the ISM will be expelled from the host galaxy. This may mean that another process (or longer lived AGN activity) is required to keep the gas unbound and prevent future star formation from occuring. See \\S4.3. \\end{itemize} Finally, we conclude that SMM~J1237+6203 is the first high-redshift ULIRG with spatially extended broad [O~{\\sc III}] emission to be mapped with IFU observations, revealing the signatures expected for a large-scale energetic outflow. However, since $\\approx$~50\\% of nearby ULIRGs hosting optical AGN activity also have broad [O~{\\sc iii}] emission, we expect SMM~J1237+6203 to be comparatively typical of the high-redshift ULIRG population. High signal-to-noise ratio (and high spatial resolution) IFU observations of further high-redshift ULIRGs will provide more definitive constraints." }, "0911/0911.2011_arXiv.txt": { "abstract": "We describe the first axisymmetric numerical code based on the generalized harmonic formulation of the Einstein equations, which is regular at the axis. We test the code by investigating gravitational collapse of distributions of complex scalar field in a Kaluza-Klein spacetime. One of the key issues of the harmonic formulation is the choice of the gauge source functions, and we conclude that a damped wave gauge is remarkably robust in this case. Our preliminary study indicates that evolution of regular initial data leads to formation both of black holes with spherical and cylindrical horizon topologies. Intriguingly, we find evidence that near threshold for black hole formation the number of outcomes proliferates. Specifically, the collapsing matter splits into individual pulses, two of which travel in the opposite directions along the compact dimension and one which is ejected radially from the axis. Depending on the initial conditions, a curvature singularity develops inside the pulses. ", "introduction": "\\label{sec_intro} In general, a detailed investigation of fully nonlinear gravitational dynamics is impossible by other than numerical means. Luckily, the numerical methods have recently reached the level of maturity that finally allows addressing many long-standing puzzles. Perhaps the most remarkable is the progress achieved in solving a general-relativistic two-body problem---the coalescence of black holes \\cite{FP2,FP1,FP3,RIT,nasa,AEI,CaltechCornell}. Driven by the gravitational wave detection prospects, the problem of the collision of two black holes or neutron stars continues to be the central front where an overwhelming majority of numerical relativity research is done. Fortunately, the computational methods employed there are portable and---as demonstrated below---can readily be applied on other problems of interest. The success of the numerical simulations was backed up by a parallel development of the software and the hardware, which provided the necessary computational resources. The rapid hardware evolution combined with the persisting regularity problems in axial symmetry eventually led to a direct transition from highly symmetrical spherical configurations to fully general 3D situations without any symmetries at all, essentially bypassing the intermediate axisymmetric case. However, here we argue that important theoretical and practical reasons exist to explore axisymmetry better, and we describe a new regular numerical code that, we believe, will be capable of achieving this.~\\footnote{See \\cite{Garfinkle_axi,Choptuik_axi,Rinne_axi_1,Rinne_axi_2} for alternative approaches.} Before describing any concrete setup we would point out one important possible use of an axisymmetric code, specifically, that it can be regarded as an efficient ``calibration tool'' for more general 3D codes. \\footnote{and as a probe of reliability of the cartoon methods \\cite{cartoon_method,FP1} used to effectively simulate axisymmetric spacetimes in 3D.} Indeed, we expect that an intrinsically axisymmetric code applied to, say, the head-on collision of two black holes would be capable of following the evolution and the resulting gravitational radiation more accurately compared to Cartesian 3D numerical implementations, both because of explicit use of the symmetry and since higher numerical resolution can be employed for given hardware resources. Several interesting open problems arise in axisymmetric gravitational collapse situations. In particular, it remains unclear whether or not the weak cosmic censorship is violated in collapse of prolate Brill waves \\cite{BrillWaves,Rinne_axi_1}. An independent observation of universality in critical collapse of gravitational waves \\cite{AbrahamsEvans} is pending, as well as further investigation of the non-spherical unstable mode that apparently shows up at threshold for black hole formation in axisymmetric collapse of a scalar field \\cite{crit_collapse_axi}. A basic problem of mathematical relativity concerning the stability of black holes with respect to nonlinear axisymmetric perturbations, can be equivalently addressed in a collapse situation by computing how fast a newly formed black hole radiates away higher multipole moments. In addition, we shall mention other, rarely cited in the context of numerical relativity, axisymmetric systems which are of great interest in theoretical work on higher-dimensional gravity. One of the most basic motivations for studying higher-dimensional spacetimes relies on the observation that the Einstein equations, describing classical General Relativity (GR), are independent of the spacetime dimension. Nevertheless, certain properties of the solutions to these equations vary dramatically with the dimension. A striking example is that axisymmetric black holes in dimensions greater than four do not necessarily have spherical horizons, but also admit horizons of toroidal topology \\cite{BlackRing}. Moreover, and in sharp contrast to their four-dimensional counterparts, higher dimensional black holes do not respect the Kerr limit \\cite{MP}, and they are unstable in certain range of parameters \\cite{GL}. One of the fundamental unresolved puzzles \\cite{HM,Choptuik_BS} is whether or not the instability leads to fragmentation of the horizon and exposition of the inner singularity, hence violating the cosmic censorship hypothesis. Since it is unlikely that any of these these problems can be addressed analytically in a systematic manner, we turn to computational methods. However, solving the Einstein equations numerically is notoriously difficult and depends crucially on the way these equations are formulated and evolved. In this paper we focus on the generalized harmonic (GH) formulation \\cite{Friedrich85,Friedrich96,Garfinkle2001} that has recently gained popularity because of its great success in the simulations of black hole binaries \\cite{FP2,FP3,CaltechCornell,FP1,Lindblom_etal}. In a nutshell, the GH approach is a way to write the field equations such that the resulting system is manifestly hyperbolic, taking the form of a set of quasilinear wave equations for the metric components. As the name suggests, the GH method generalizes the harmonic approach, which achieves strong hyperbolicity by choosing the harmonic spacetime coordinates. In the GH approach much of the coordinate freedom is regained through the introduction of certain gauge source functions, which also maintains the desirable property of strong hyperbolicity of the field equations. In fact, the source functions can be thought of as representing the coordinate freedom of the Einstein equations, and when constructing solutions of the equations, via an initial value approach, for example, they must be completely specified in some fashion. Choosing the source functions in a controlled way is a key issue of the GH technique and after testing several recent prescription, we conclude that a damped wave gauge \\cite{LS} is remarkably robust in collapse situations. Applications of the GH approach in spherically-symmetric situations were studied in \\cite{SorkinChoptuik} and here we employ the method in axisymmetry. In both cases it is natural to use coordinates in which the symmetries of the spacetime are explicit. However, these coordinates, are formally singular: at the origin in spherical symmetry and on the axis in axial symmetry. Thus, the field equations have to be regularized in numerical implementations; here we describe a regularization procedure that is compatible with the GH formulation. Our numerical implementation of the GH system is a free evolution code that advances initial data by solving a set of wave equations. In addition, there are also constraint equations that must be satisfied during the evolution. Although the constraints are consistently preserved in the GH approach in the continuum limit, in numerical computations at finite resolution, constraint violations generically develop. In order to maintain stability these deviations must be damped and we discuss an effective method that achieves this. Having in mind implications for higher-dimensional GR we test our new code by studying gravitational collapse of a complex scalar field in a $D$-dimensional Kaluza-Klein (KK) spacetime. This background, which has a single compact extra dimension curled into a circle, is a classical example of a higher-dimensional compactified spacetime that in certain limits can appear four dimensional, for example when the size of the compact dimension is small.\\footnote{In this paper, however, the size of the KK circle is arbitrary.} Assuming spherical symmetry in the infinite $(D-2)$-dimensional portion of the space makes the problem 2+1 that depends on time and two spatial coordinates: one in the radial direction, and one along the KK circle. We perform a series of numerical simulations where the initial distribution of the scalar matter is freely specified and the outcome of the evolution depends on the ``strength'' of the initial data as well as on its topology. The weak data correspond to the dispersion of relatively dilute pulses, while a typical strong data configuration leads to black hole formation or nearly does so. Our preliminary results indicate a wide range of the black hole forming scenarios, including how many holes form and of what topology. For instance, a static distribution of matter centered at the axis and localized in the KK direction with the energy-density above certain threshold collapses to form a black hole with a quasi-cylindrical horizon, smeared along the extra-dimension. Data with the energy-density below this threshold evolve to form a quasi-spherical horizon centered around the initial matter distribution. By further reducing the density we find that resulting black holes become progressively smaller, and at some critical density a pulse of matter is emitted radially away from the axis and a curvature singularity develops inside it. We find that for slightly lower initial densities the evolving matter splits into several pulses and two of them individually collapse to form the singularities moving apart along the KK dimension. Finally, when the initial matter distribution becomes dilute below a certain limit, no black holes or curvature singularities are created. In the next section we describe the class of effectively 2+1 dimensional models where our code can currently be applied. In Sec. \\ref{sec_GH} we present the basic formulas of the GH formulation and discuss the constraints and a method to damp their violations. We describe an axis regularization in Sec. \\ref{sec_axis} and boundary conditions in Sec. \\ref{sec_eqs}. Although, in this work we integrate full $D$-dimensional equations, in the Appendix we describe an alternative approach---one that uses a dimensional reduction on the symmetry and integrates the reduced 2+1 equations---and compare its performance with ours. Coordinate conditions and the initial data problem are formulated in Secs. \\ref{sec_coord_condition} and \\ref{sec_id}, respectively. We use several diagnostics to probe the spacetimes that we construct, including computation of asymptotic measurables and apparent horizons, and describe that in Sec. \\ref{sec_diagnostics}. After elaborating on our numerical algorithm in Sec. \\ref{sec_numercal_approach} we test its performance in Sec. \\ref{sec_num_experements}, giving detailed accounts of various aspects, such as specific coordinate choices, constraint damping, numerical dissipation, and convergence. We conclude in Sec. \\ref{sec_conclusion}, outline possibilities of improvement, and discuss some future prospects. ", "conclusions": "" }, "0911/0911.1077_arXiv.txt": { "abstract": "We discuss the use of SPICA to study the cosmic history of star formation and accretion by supermassive black holes. The cooling lines, in particular the high-$J$ rotational lines of CO, provide a clear-cut and unique diagnostic for separating the contributions of star formation and AGN accretion to the total infrared luminosity of active, gas-rich galaxies. We briefly review existing efforts for studying high-$J$ CO emission from galaxies at low and high redshift. We finally comment on the detectability of cooling radiation from primordial (very low metallicity) galaxies containing an accreting supermassive black hole with SPICA/SAFARI. ", "introduction": "For several decades, our only direct probes of the high-redshift universe were QSOs and radio galaxies. While significant insight into the properties and evolution of these populations was obtained, almost nothing was known about the population of normal galaxies at high $z$. This has changed radically over the last decade. Deep imaging and spectroscopy from the ground and from space have revealed a rich and diverse population of high-$z$ galaxies, and enormous progress has been made in establishing a complete inventory of the high-$z$ universe, in terms of mass and metallicity budget, energy output and evolution over cosmic time. In addition, first investigations of the genesis of the fundamental scaling relations of galaxies (e.g., the fundamental plane and the Tully-Fisher relation) are being undertaken, and will provide insight into the origin of the galaxy population itself. One of the most remarkable scaling relations is the $M_\\bullet-\\sigma$ relation, i.e., the relation between the mass of a nuclear supermassive black hole (SMBH) and the velocity dispersion of the spheroid (elliptical or bulge) hosting the black hole \\citep{Magorrianetal98,FerrareseMerritt00}. The fact that the black hole mass and total galaxy mass are closely related is truly remarkable, given that there is a factor of $\\sim10^8$ between the AU-size Schwarzschild radius of the black hole and the kpc-size dimension of the galaxy. This is generally interpreted as evidence that the formation and growth of the SMBH is directly related to the formation process of the stellar population, e.g., in a violent burst of star formation. The picture of the simultaneous build-up of a SMBH with an extreme burst of star formation provides a new context for the luminous and ultraluminous infrared galaxies (LIRGs and ULIRGs), long suggested to be the birthplaces of SMBHs powering active galactic nuclei (AGNs). Studies of local ULIRGs demonstrate that they plausibly evolve into moderate mass field ellipticals \\citep{Genzeletal01}. The giant ellipticals in the local universe would then be the result of similar but much more extreme events at higher redshifts. Indeed, the population of submillimetre galaxies (SMGs) provides the more luminous high redshift counterpart of the local LIRGs and ULIRGs and may be responsible for the formation of local massive spheroids and for generating QSO activity at high redshift \\citep{Blainetal99}. \\begin{figure}[th] \\begin{center} \\includegraphics[width=8cm]{vanderWerfP_fig1.eps} \\end{center} \\caption{CO emergent intensity distributions for a PDR and an XDR with {\\it identical\\/} impinging energy density, as described in Section~\\ref{vanderWerfP.SecDiag}. Note that high-$J$ CO emission is expected in both cases, but vastly more prominently in the XDR \\citep{SpaansMeijerink08}.} \\label{vanderWerfP_fig1} \\end{figure} ", "conclusions": "" }, "0911/0911.1241_arXiv.txt": { "abstract": "We provide an analytical expression for the trispectrum of the Cosmic Microwave Background (CMB) temperature anisotropies induced by cosmic strings. Our result is derived for the small angular scales under the assumption that the temperature anisotropy is induced by the Gott--Kaiser--Stebbins effect. The trispectrum is predicted to decay with a non-integer power-law exponent $\\ell^{-\\rho}$ with $6<\\rho<7$, depending on the string microstructure, and thus on the string model. For Nambu--Goto strings, this exponent is related to the string mean square velocity and the loop distribution function. We then explore two classes of wavenumber configuration in Fourier space, the kite and trapezium quadrilaterals. The trispectrum can be of any sign and appears to be strongly enhanced for all squeezed quadrilaterals. ", "introduction": "Although cosmic strings may be of various early universe origins~\\cite{Kibble:1976sj,Dabholkar:1990yf, Hindmarsh:1994re, VilShe94, Kofman:1994rk,Yokoyama:1989pa, Sakellariadou:2006qs, Copeland:2003bj, Sarangi:2002yt, Dvali:2003zj}, being line-like gravitational objects, they induce temperature discontinuities in the CMB through the Gott--Kaiser--Stebbins (GKS) effect \\cite{Gott:1984ef,Kaiser:1984iv}. Direct searches for such discontinuities have been performed without success but do provide upper limits to the string tension $U$~\\cite{Jeong:2004ut, Jeong:2007, Christiansen:2008vi}. On the other hand, if cosmic strings are added to the standard power-law $\\Lambda$CDM model~\\cite{Bouchet:2000hd}, it has been shown in Refs.~\\cite{Battye:2006pk,Bevis:2007gh} that the CMB data are fitted even better if the fraction of the temperature power spectrum due to strings is about $10\\%$ (at $\\ell=10$). Such a fraction of string would even dominate the primary anisotropies of inflationary origin for $\\ell \\gtrsim 3000$~\\cite{Fraisse:2007nu}. With the advent of the arc-minute resolution CMB experiments and the soon incoming Planck satellite data, it is therefore crucial to develop reliable tests for strings~\\cite{Seljak:2006}, as to understand the non-Gaussian signals. The probability distribution of the fluctuations due to the GKS effect is known to be skewed and has a less steep decay than Gaussian~\\cite{Fraisse:2007nu}, a feature which can be explained in a simple model of kinked string~\\cite{Takahashi:2008ui}. In Ref.~\\cite{Hindmarsh:2009qk}, we have studied the temperature bispectrum induced by cosmic strings both analytically and numerically by using Nambu--Goto string simulations. We found good agreement between the analytical and numerical bispectrum for both the overall amplitude and the geometrical factor associated with various triangle configurations of the wavevectors. This agreement suggests that our analytical assumptions are capturing the relevant non-Gaussian features of a string network and could be used to derive other statistical properties. In this paper, we present new results concerning the trispectrum, i.e. the four-point function of the temperature anisotropy~\\cite{Hu:2001fa}. As pointed out in Ref.~\\cite{Hindmarsh:2009qk}, the bispectrum is generated only when the background spacetime breaks the time reversal symmetry. Because our universe is expanding, the time reversal symmetry is indeed broken and we get a non-vanishing string bispectrum. On the other hand, the trispectrum can be generated even in Minkowski spacetime and one may naively expect a stronger non-Gaussian signal than for the bispectrum. Motivated by this observation, we provide in this paper an analytical derivation of the string trispectrum and study its dependency for various quadrilateral configurations in Fourier space. Given the fact that analytical predictions and numerical results exhibit good agreement both for the power spectrum~\\cite{Hindmarsh:1993pu, Fraisse:2007nu} and the bispectrum~\\cite{Hindmarsh:2009qk}, we expect our result here to agree as well with the numerics. Performing such a comparison would however require a significant amount of computing resources which motivated us to leave it for a future work. Our main result can be summarized by Eq.~(\\ref{eq:finaltri}). Interestingly, the power-law behaviour of the trispectrum exhibits a non-integer exponent which can be related to the small scale behaviour of the string tangent vector correlator. In the framework of Nambu--Goto strings, this exponent is related to the mean square string velocity~\\cite{Polchinski:2006ee, Copeland:2009dk} and to the scaling loop distribution function~\\cite{Ringeval:2005kr, Rocha:2007ni}. The paper is organised as follows. In the next section we briefly recall the assumptions at the basis of our analytical approach and derive the trispectrum in Sec.~\\ref{sec:trispectrum}. As an illustration, we apply our result to some specific quadrilaterals in Sec.~\\ref{sec:quad} and exhibit some configurations that leads to a divergent trispectrum. They should provide the cleanest way to look for a non-Gaussian string signal. ", "conclusions": "\\label{sec:concl} In this paper, we have analytically derived the CMB temperature trispectrum induced by cosmic strings using the string correlation functions in the Gaussian approximation. The trispectrum generically decays with a non-integer power-law behaviour at small angular scales which depends on the string microstructure through the behaviour of the tangent vector correlator on small distances. Its eventual detection and measurement may therefore help to distinguish between different string models. We have also found that the trispectrum diverges, in the framework of our approximations, on all squeezed configurations whose measurements remain however limited by the finite experimental resolution. In fact, such a non-integer power-law is linked to the existence of a ``flat direction'' at leading order and the four-point function ends up being sensitive to the next-to-leading order string tangent vector correlator. This situation is also present in the n-point function and we do expect all of the higher n-point function to exhibit non-integer power-law behaviours. Since this situation was not encountered for the two- and three-point functions, the next step will be to compare our results here with the trispectrum computed from CMB maps obtained by string network simulations. Finally, let us notice that we have not attempted to make any comparison with a CMB trispectrum produced by primordial non-Gaussianities of inflationary origin. The situation is nearly the same as it is for the string bispectrum~\\cite{Hindmarsh:2009qk}. The so-called $\\tauNL$ and $\\gNL$ parameters quantify the amplitude of the primordial four-point function of the curvature perturbation on super-Hubble scales. As a result, the induced trispectrum of the CMB temperature fluctuations strongly depends on the CMB transfer functions and exhibits damped oscillations with respect to the multipole moments. Here, we have direcly derived the CMB temperature trispectrum produced by the strings and it would therefore make no sense to find an associated $\\tauNL$ and $\\gNL$. An alternative approach might be to estimate what values $\\tauNL$ and $\\gNL$ would assume in a primordial-type oriented data analysis if the non-Gaussianities were actually due to strings. This could be done with a Fisher matrix analysis for a given experiment but we leave this question for a forthcoming work." }, "0911/0911.1131_arXiv.txt": { "abstract": "We use a novel method to predict the contribution of normal star-forming galaxies, merger-induced bursts, and obscured AGN, to IR luminosity functions (LFs) and global SFR densities. We use empirical halo occupation constraints to populate halos with galaxies and determine the distribution of normal and merging galaxies. Each system can then be associated with high-resolution hydrodynamic simulations. We predict the distribution of observed luminosities and SFRs, from different galaxy classes, as a function of redshift from $z=0-6$. We provide fitting functions for the predicted LFs, quantify the uncertainties, and compare with observations. At all redshifts, `normal' galaxies dominate the LF at moderate luminosities $\\sim L_{\\ast}$ (the `knee'). Merger-induced bursts increasingly dominate at $L\\gg L_{\\ast}$; at the most extreme luminosities, AGN are important. However, all populations increase in luminosity at higher redshifts, owing to increasing gas fractions. Thus the `transition luminosity' between normal and merger-dominated sources increases from the LIRG-ULIRG threshold at $z\\sim0$ to bright Hyper-LIRG thresholds at $z\\sim2$. The transition to dominance by obscured AGN evolves similarly, at factor of several higher $L_{\\rm IR}$. At all redshifts, non-merging systems dominate the total luminosity/SFR density, with merger-induced bursts constituting $\\sim5-10\\%$ and AGN $\\sim1-5\\%$. Bursts contribute little to scatter in the SFR-stellar mass relation. In fact, many systems identified as `ongoing' mergers will be forming stars in their `normal' (non-burst) mode. Counting this as `merger-induced' star formation leads to a stronger apparent redshift evolution in the contribution of mergers to the SFR density. We quantify how the evolution in LFs depends on evolution in galaxy gas fractions, merger rates, and possible evolution in the Schmidt-Kennicutt relation. We discuss areas where more detailed study, with full radiative transfer treatment of complex three-dimensional clumpy geometries in mixed AGN-star forming systems, is necessary. ", "introduction": "\\label{sec:intro} Understanding the global star-formation history of the Universe remains an important unresolved goal in cosmology. In recent years, observations of the properties of galaxies in the infrared, at redshifts $z=0-3$, have begun to shed light on this history, but have also revealed a number of intriguing questions. Of particular interest are the roles of mergers and AGN in driving star formation and/or the infrared luminosities of massive systems. A wide range of observed phenomena support the view that gas-rich\\footnote{ By ``gas,'' we refer specifically to cold, star-forming gas in galaxy disks, as opposed to hot, virialized gas.} mergers are important to galaxy evolution; but it is less clear what their role is in the global star formation process and buildup of stellar mass in the Universe. In the local Universe, the population of star-forming galaxies appears to transition from ``quiescent'' (non-disturbed)\\footnote{In this paper, we use the term ``quiescent'' to refer to star-forming systems that are not strongly disturbed in e.g.\\ major mergers and are forming stars in similar fashion to most ``normal'' disks. We do {\\em not} mean non-star forming systems, as the term is used in some literature.} disks -- which dominate the {\\em total} star formation rate/IR luminosity density -- at the luminous infrared galaxy (LIRG) threshold $10^{11}\\,\\lsun$ ($\\dot{M}_{\\ast}\\sim 10-20\\,\\msun\\,{\\rm yr^{-1}}$) to clearly merging, violently disturbed systems at a few times this luminosity. The most intense starbursts at $z=0$, ultraluminous infrared galaxies (ULIRGs; $L_{\\rm IR}>10^{12}\\,\\lsun$), are invariably associated with mergers \\citep[e.g.][]{joseph85,sanders96:ulirgs.mergers, evans:ulirgs.are.mergers}, with dense gas in their centers providing material to feed black hole (BH) growth and to boost the concentration and central phase space density of merging spirals to match those of ellipticals \\citep{hernquist:phasespace,robertson:fp}. Various studies have shown that the mass involved in these starburst events is critical to explain the relations between spirals, mergers, and ellipticals, and has a dramatic impact on the properties of merger remnants \\citep[e.g.,][]{LakeDressler86,Doyon94,ShierFischer98,James99, Genzel01,tacconi:ulirgs.sb.profiles,dasyra:mass.ratio.conditions,dasyra:pg.qso.dynamics, rj:profiles,rothberg.joseph:kinematics,hopkins:cusps.ell,hopkins:cores}. At high redshifts, the role of mergers is less clear. It is clear that LIRGs and ULIRGs increase in relative importance with redshift, with LIRGs dominating the star formation rate/IR luminosity densities at $z\\sim1$ and ULIRGs dominating at $z\\sim2$ \\citep[e.g.][]{lefloch:ir.lfs,perezgonzalez:ir.lfs,caputi:ir.lfs,magnelli:z1.ir.lfs}. This, together with the fact that merger rates are expected and observed to increase with redshift \\citep[by a factor $\\sim10$ from $z=0-2$; see e.g.][and references therein]{hopkins:merger.rates} has led to speculation that the merger rate evolution may in fact drive the observed evolution in the cosmic SFR density, which rises rapidly from $z\\sim0-2$ and then turns over, declining more slowly \\citep[e.g.][and references therein]{hopkinsbeacom:sfh}. However, many LIRGs at $z\\sim1$, and potentially ULIRGs at $z\\sim2$, appear to be ``normal'' galaxies, without dramatic morphological disturbances associated with the local starburst population or large apparent AGN contributions \\citep{yan:z2.sf.seds,sajina:pah.qso.vs.sf, dey:2008.dog.population,melbourne:2008.dog.morph.smooth, dasyra:highz.ulirg.imaging.not.major}. At the same time, even more luminous systems appear, including large populations of Hyper-LIRG (HyLIRG; $L_{\\rm IR}>10^{13}\\,\\lsun$) and bright sub-millimeter galaxies \\citep[e.g.][]{chapman:submm.lfs,younger:highz.smgs, younger:sma.hylirg.obs,casey:highz.ulirg.pops}. These systems exhibit many of the traits more commonly associated with merger-driven starbursts, including morphological disturbances, and may be linked to the emergence of massive, quenched (non star-forming), compact ellipticals at times as early as $z\\sim2-4$ \\citep{papovich:highz.sb.gal.timescales, younger:smg.sizes,tacconi:smg.maximal.sb.sizes, schinnerer:submm.merger.w.compact.mol.gas, chapman:submm.halo.clustering,tacconi:smg.mgr.lifetime.to.quiescent}. But reproducing their abundance and luminosities remains a challenge for current models of galaxy formation \\citep{baugh:sam, swinbank:smg.counts.vs.durham, narayanan:smg.modeling, younger:warm.ulirg.evol}. In a related vein, observations of a tight correlation between the masses of super-massive BHs and their host spheroid properties \\citep{Gebhardt00,FM00,magorrian,novak:scatter,hopkins:bhfp.obs} suggest a tight coupling between BH growth and star formation, perhaps in particular to the mergers believed to drive the formation of the most massive bulges. Considering the energy output required to form the BH population \\citep[e.g.][]{soltan82}, or the observed bolometric quasar energy density as a function of redshift \\citep[see][and references therein]{hopkins:bol.qlf}, it is clear that the {\\em bolometric} output of quasars and AGN is at least roughly comparable to the total infrared luminosity density of the Universe at most redshifts ($z\\sim1-3$) -- although the measurements above suggest it is still a factor $\\sim2-3$ lower. Some recent observations have suggested that the population of very luminous, highly obscured (Compton-thick) quasars may be considerably larger than previous estimates, in which case the heavily-obscured AGN population could represent a large fraction of the total IR luminosity density at high redshifts \\citep{hickox:bootes.obscured.agn,daddi:2007.high.compton.thick.pops, treister:compton.thick.fractions}. This would have dramatic implications not just for BH populations and e.g.\\ the implied radiative efficiencies of BH accretion, but also for the implied total star formation rate density. Some apparent discrepancies between e.g.\\ the total mass density observed in old stars and the implied star formation rate density have been cited as possible evidence of a time-dependent stellar initial mass function (IMF); but a rising contribution from obscured AGN at high redshifts could mimic this effect \\citep{hopkinsbeacom:sfh,dave:imf.evol}. In particular, there are long-standing questions of what powers the most luminous infrared sources, for example, ULIRGs and sub-millimeter galaxies. This debate extends to the discovery of these objects \\citep[see e.g.][]{soifer84b,soifer:iras,scoville86, sargent87,sanders88:warm.ulirgs, solomon.downes:ulirg.ism}, and has persisted despite the addition of millimeter spectroscopy and observations in a large number of independent wavebands \\citep[for a review of the debate, see both][]{sanders:agn.vs.sf.in.ulirgs, joseph:sb.vs.agn.power.ulirgs}. Although some evidence suggests that they are primarily powered by star formation \\citep{farrah:qso.vs.sf.sed.fitting, lutz:pah.qso.vs.sf.local,sajina:pah.qso.vs.sf, pope:2008.pah.agn.dont.dominate.smgs,pope:2008.dog.sfr.properties, watabe:highz.ulirg.sb.vs.agn,nardini:2009.agn.vs.sb.contrib.in.ulirgs}, the constraints and correlations typically invoked have inherent factor $\\sim2$ uncertainties, and thus could easily accommodate comparable power input from star formation and AGN. Moreover, a sufficiently obscured AGN, in a medium with the right optical depth properties, is indistinguishable from star formation by the usual indicators (e.g.\\ PAH strengths, emission region sizes, or any other infrared spectral or morphological criteria). Hence even at $z=0$, debate surrounds the power source of many bright infrared systems, and there exist a number of examples of systems classified as ``star formation dominated'' by all of these metrics that later revealed Compton-thick AGN whose longer-wavelength emission has been fully re-processed, even into ``cool'' dust \\citep[see e.g.][and references therein]{alexander:2008.compton.thick.z2.qsos}. In a bolometric sense, the most luminous galaxies observed (with $L\\gtrsim 10^{14}\\,\\lsun$ or $\\gg 10^{47}\\,{\\rm erg\\,s^{-1}}$) are the most luminous quasars; although the contribution of these systems to the infrared remains highly uncertain. This may be important for resolving the theoretical difficulties in modeling these bright systems. There has been important theoretical progress in modeling these processes in an {\\em a priori} manner \\citep[see e.g.][]{baugh:sam, hopkins:groups.qso,narayanan:smg.modeling,younger:warm.ulirg.evol}. However, two basic limitations remain. In direct cosmological hydrodynamic simulations, as well as semi-analytic models of galaxy formation, it is well known that it remains challenging to accurately reproduce global quantities such as the galaxy mass function and the distribution of sizes, gas fractions, and hence star formation rates, especially the distributions of star-forming gas and their relations to whether or not galaxies are ``quenched'' \\citep[recently, see e.g.][]{weinmann:group.cat.vs.sam, maller:sph.merger.rates,kimm:passive.sats.vs.centrals,fontanot:downsizing.vs.sams}. This makes it difficult to determine whether discrepancies between such models and the observations owe to their treatment of star formation, or to discrepancies in these quantities. Moreover, it is difficult to disentangle the effects of these different properties on the distribution of star formation rates. In addition, for merger-induced starburst and AGN activity, although it may be possible to roughly estimate some global quantities (e.g.\\ the total mass involved in a starburst) from simple analytic motivations or low-resolution cosmological simulations (several $\\sim$kpc typical), it is not straightforward to estimate the chaotic, time-dependent behavior of full lightcurves needed to estimate the distribution of time spent at different, rapidly varying luminosities without high-resolution simulations of individual systems. Since the number density of the most bright systems is exponentially declining, fluctuations and features in the starburst/AGN fueling history on small time and spatial scales ($\\Delta t\\sim10^{7}\\,{\\rm yr}$, $R\\lesssim 100\\,$pc) can be critical for correct estimates of their contributions to bright populations. In this paper, we present theoretical predictions for the distribution of galaxy star formation rates and infrared luminosities, as a function of galaxy mass and redshift, using a novel methodology that can circumvent some of these obstacles. We combine a halo-occupation based approach, in which we take galaxy properties as fixed from observations at each epoch, and then apply rules for the distribution of star formation rates/infrared luminosities in ``quiescent'' systems, merger-induced starbursts, and obscured AGN, calculated from a large suite of high-resolution hydrodynamic simulations of individual galaxies and galaxy mergers. We use this to independently estimate the contributions of ``normal'' galaxies, mergers, and AGN to the luminosity functions. The comparisons we make are approximate -- we do not include full time-dependent radiative transfer in simulations (the subject of future work, in progress), and so focus on integral quantities such as the total IR luminosity and SFR distributions, that are less sensitive to issues of e.g.\\ the exact dust distribution, temperature, and other properties. We also explicitly separate the contributions of AGN and star formation, but stress that, in real systems where the two are comparable, their additive effects are non-linear, and will require further study. We show how adding or removing components of the model taken from observations such as e.g.\\ the distribution of galaxy sizes and gas fractions affects these consequences. We compare to observations of all quantities, where available, and find reasonable agreement but with some interesting apparent discrepancies at high redshifts. We also show how these populations relate to the scatter in star formation rates at fixed galaxy masses, and in a global sense to the total star formation rate density, the star formation rate density in mergers, and the fraction of the inferred star formation rate density which might really be driven by obscured AGN activity. Readers interested primarily in the comparison of predictions and observations may wish to skip directly to \\S~\\ref{sec:lfs}. Throughout, we adopt a $\\Omega_{\\rm M}=0.3$, $\\Omega_{\\Lambda}=0.7$, $h=0.7$ cosmology and a \\citet{chabrier:imf} stellar IMF (discussed further below), but these choices do not affect our conclusions. ", "conclusions": "\\label{sec:discussion} We present a simple model for the distribution of star formation rates and infrared luminosities owing to ``normal'' star-forming disks, merger-induced starbursts, and AGN. Comparing this with observations, we find reasonable agreement at $z\\sim0-3$. At all redshifts, we find that the low-luminosity population is dominated by disks, whereas the high-luminosity population becomes progressively more dominated by merger-induced bursts and then, ultimately, obscured AGN. The threshold for this transition is always at high luminosities and low space densities. At higher redshifts, gas fractions in all systems increase -- hence, specific star formation rates even in ``typical'' systems are much higher at high-$z$. As a consequence, the luminosity threshold for the disk-merger transition increases with redshift, from between LIRGs and ULIRGs in the local Universe, to ULIRG luminosities at $z\\sim1$, to HyLIRG luminosities at $z\\sim2-4$. Similarly, the threshold for AGN dominance, at the bright ULIRG range at low redshifts, rises to HyLIRG luminosities at $z\\sim1-2$ and to the very brightest HyLIRGs at $z>2$. We provide simple fitting functions for each of these quantities, and show how they depend on different parameters in the model. Most critically, it is the evolution in galaxy gas fractions that drives most of the evolution in the IR LFs. Observations have shown that typical gas fractions in massive, star-forming galaxies increase rapidly with redshift \\citep[see e.g.][]{erb:lbg.gasmasses, bouche:z2.kennicutt,puech:tf.evol, mannucci:z3.gal.gfs.tf, forsterschreiber:z2.sf.gal.spectroscopy}. This naturally follows from the facts that cooling rates onto galaxies at high redshift are much higher than at low redshift, and there has simply been less time to process gas into stars (higher cosmic densities may also make stellar and AGN feedback relatively less efficient). A higher gas fraction by a factor of $\\sim3$, as implied by observations of Milky-Way mass disks at $z\\sim2-3$, leads to a factor $\\sim5-6$ higher star formation rate, according to the Kennicutt-Schmidt relation. At high masses/luminosities, i.e.\\ where the number density of systems is falling exponentially, such a systematic increase in luminosity more than offsets the declining space density of massive galaxies with redshift. Other parameters have little effect. Changes in galaxy sizes and/or structural parameters, while potentially very important for e.g.\\ how gas fractions are maintained, make little difference to the SFR distributions given some gas fraction and stellar mass distributions. Allowing for a more steep index in the \\citet{kennicutt98} relation may help to account for the most luminous observed systems such as bright sub-millimeter galaxies at the Hyper-LIRG threshold at $z\\sim2-4$ \\citep{chapman:submm.lfs}. However, the number counts of such objects remains quite uncertain, and recently it has been suggested that the average counts might be much lower with cosmic variance still a concern \\citep{austermann:2009.aztec.submm.source.counts}. Larger samples and better calibration of bolometric corrections -- in particular, real knowledge of the appropriate dust temperatures for conversion to total-IR luminosities, which requires sampling both sides of the cold dust peak \\citep[see e.g.][]{younger:mm.obs.z2.ulirgs} -- will be needed for better understanding of extreme systems. Our simple model succeeds reasonably well at explaining the observed global SFR density. At all redshifts, normal systems dominate the global SFR density. Obscured AGN contribute little, $\\lesssim5\\%$, to the total IR luminosity density. Substantial bias to IR-based SFR density estimates from obscured AGN would require an undiscovered population of heavily, isotropically obscured sources with luminosity densities $\\sim5$ times what is currently suggested (which would be in conflict with relic BH mass densities). The contribution of merger-induced bursts is similarly small, $\\sim5-10\\%$ at most redshifts. This owes both to the physics discussed above -- disks also rapidly increase their SFR density with redshift owing to higher gas fractions -- and also to the fact that increasing disk gas fractions arbitrarily will not continue to increase the merger-induced burst contribution arbitrarily. Rather, as discussed in detail in \\citet{hopkins:disk.survival,hopkins:disk.survival.cosmo}, at high gas fractions angular momentum loss in mergers becomes less efficient -- thus for an otherwise identical merger with a much larger gas fraction, the fraction of gas funneled into the nuclear starburst, relative to the total available, will be less (and the fraction that remains in an extended disk distribution to continue ``normal'' mode star formation will be larger), even if the absolute mass in the burst is larger. Similarly, at a given mass, the distribution of SFRs at all redshifts is dominated by normal-mode star formation -- merger-induced bursts are important only in the high-SFR tail ($\\sim2-3\\,\\sigma$) of the distribution. These trends explain a number of recent observations that similarly indicate a small effect of merger enhancements to star formation rates, and that show that most star formation by number and luminosity density appears to follow a simple trend or ``main sequence'' as a function of galaxy stellar mass and redshift \\citep{blain:ir.lf.synthesis.model, noeske:2007.sfh.part1, noeske:sfh,papovich:ssfr, bell:morphology.vs.sfr, jogee:merger.density.08, robaina:2009.sf.fraction.from.mergers}. Support for these fractions also comes from completely independent sources. Recently, a number of high-resolution studies of spheroid formation via galaxy mergers have shown that properties such as the surface brightness profiles, sizes, concentrations, kinematics, and isophotal shapes of spheroids are very sensitive to the mass fractions formed in such bursts, which produce dense, disky, nuclear mass concentrations, versus the mass in a more extended envelope formed via the violent relaxation of the pre-burst stellar disks \\citep[see e.g.][]{cox:kinematics, naab:gas,robertson:fp,burkert:anisotropy, jesseit:merger.rem.spin.vs.gas,hopkins:cusps.mergers, hopkins:cusps.fp,hopkins:cusps.evol}. In particular, typical $\\sim L_{\\ast}$ early-type galaxies, which dominate the spheroid stellar mass density, have properties that are reproduced accurately by simulations if and only if this burst fraction is $\\sim10\\%$ \\citep[for details, see][]{hopkins:cusps.ell,hopkins:cores}. Independent analysis of their stellar population properties leads to similar conclusions \\citep{mcdermid:sauron.profiles,sanchezblazquez:ssp.gradients, reda:ssp.gradients,foster:metallicity.gradients.prep}. There is an important technical distinction between the total SFR density in ``ongoing'' or recent mergers and that actually {\\em induced} by the merger (we present predictions for both). The former includes systems in their ``normal'' star-forming mode, observable for $\\sim$Gyr as perturbed or in pairs; the latter reflects specifically the $\\sim10^{8}\\,$yr event where gravitational torques drive a nuclear starburst. Under some circumstances, especially in very gas-rich mergers, the sum of this ``normal mode'' star formation over a long $\\sim$Gyr duration yields significantly more total stellar mass formed than in the burst itself. The ``ongoing'' merger SFR density must, of course, rise with the observed merger fraction, reaching $\\sim20\\%$ at $z=2$ and as high as $\\sim20-50\\%$ at $z>4$; however we stress that these high fractions reflect predominantly the ``normal'' modes of star formation simply present in systems that may be on their way to merging. Finally, we caution that the above comparisons are approximate, and intended as a broad means of comparing the primary drivers of star formation and their contributions relative to observed IR luminosity functions and SFR distributions. We have ignored a number of potentially important effects: for example, obscuration is a strong function of time in a merger, and may affect various luminosities and morphological stages differently. Moreover, our simple linear addition of the star formation contribution of mergers to the IR LF and the AGN contribution is only technically correct if one or the other dominates the IR luminosity at a given time in the merger; however, there are clearly times during the final merger stages when the contributions are comparable. Resolving these issues requires detailed, time-dependent radiative transfer solutions through high-resolution simulations that properly sample the merger and quiescent galaxy parameter space at each redshift, and is outside the scope of this work \\citep[although an important subject for future, more detailed study; see, e.g.][]{li:radiative.transfer,narayanan:smg.modeling, narayanan:molecular.gas.in.smgs,younger:warm.ulirg.evol}. It would be a mistake, therefore, to read too much into e.g.\\ the detailed predictions for sub-millimeter galaxies or other extreme populations in Figure~\\ref{fig:ir.lfs} that may have complex dust geometries and/or a non-trivial mix of contributions from all of ``normal'' and ``burst'' mode star formation as well as AGN. However, most of our predicted qualitative trends, including the evolution of the luminosity density (and approximate relative contribution of mergers) and the shift in where quiescent or merger-driven populations dominate the bright IR LF, should be robust." }, "0911/0911.4839_arXiv.txt": { "abstract": "{There is mounting evidence for an extra-planar gas layer around the Milky Way disk, similar to the anomalous \\hi~ gas detected in a few other galaxies. As much as 10\\% of the gas may be in this phase.} {We analyze \\hi~ clouds located in the disk-halo interface outside the solar circle to probe the properties of the extra-planar \\hi~ gas, which is following Galactic rotation.} {We use the Leiden/Argentine/Bonn (LAB) 21-cm line survey to search for \\hi~ clouds which take part in the rotation of the Galactic plane, but are located above the disk layer. Selected regions are mapped with the Effelsberg 100-m telescope. Two of the \\hi~ halo clouds are studied in detail for their small scale structure using the Westerbork Synthesis Radio Telescope (WSRT)\\thanks{The Westerbork Synthesis Radio Telescope is operated by ASTRON (Netherlands Foundation for Research in Astronomy) with support from the Netherlands Foundation for Scientific Research NWO.} and the NRAO Very Large Array (VLA)\\thanks{The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. }.} {Data from the 100m telescope allow for the parameterization of 25 distinct \\hi~ halo clouds at Galactocentric radii 10 kpc $1.3$ cm$^{-3}$, masses up to 24 \\msun, and are on average in pressure equilibrium with the surrounding envelopes. Pressures and densities fall within the expectations from theoretical phase diagrams ($P$ vs $\\langle n_{h} \\rangle$). The \\hi~ cores tend to be unstable if one assumes a thermally bistable medium, but are in better agreement with models that predict thermal fragmentation driven by a turbulent flow. } {} ", "introduction": "\\label{sec_1} The \\hi~ gas is a major constituent of the interstellar medium (ISM), and it is a well known property of this gas that it settles in the Galactic plane. From the very first \\hi~ observations it is also known that the disk gas is co-rotating with the stellar disk. Taking both properties together one may use the gas distribution to describe the morphology of the Galactic disk. As yet, there is no sharp boundary for the disk emission. \\citet{Oort1962dm1} was the first who mentioned this fact. ``Well outside the real disk one still finds neutral hydrogen with an average density of between 5 and 10 per cent of the intensities one observes in the plane''. Oort was referring to observations with the Dwingeloo telescope, and \\citet{Shane1967IAUS} described this gas later as a ``galactic envelope'', a smooth envelope of neutral hydrogen surrounding the spiral structure, following the same Galactic rotation as the gas in the plane. Further discussion of this envelope was given by \\citet{Takakubo1967BAN} and by \\citet{Shane1971AAS}, but there was some concern about a possible contamination by stray radiation from the antenna diagram of the Dwingeloo telescope. The extra-planar gas component was also visible in the \\citet{WW1973AAS} survey, and \\citet{Lockman1984} studied this feature in some more detail. Supplementing observations were made with the NRAO 300-foot telescope and came also from the NRAO 140-foot survey by \\citet{Burton1983AAS}. Lockman argued that his analysis was not affected by stray radiation. He found for Galactocentric radii $4 \\la R \\la 8$ kpc that 13\\% of the \\hi~ gas is located outside the disk, extending to $z$-distances of 1 kpc or more and termed this component an ``\\hi~ halo''. The Bell Labs survey \\citep{Stark1992ApJS} is only little affected by instrumental effects, and \\citet{Lockman1991} used this survey to analyze the nature of the vertical \\hi~ gas distribution in the direction of the Galactic poles. They proposed for the \\hi~ gas a decomposition in several layered structures, corresponding to distinct different isothermal cloud populations. The scale height for each component results from the pressure balance of a cloudy turbulent medium against the gravitational potential of the Milky Way. The concept of a layered structure of the \\hi~ contains essentially three components: a cold neutral medium (CNM), a warm neutral medium (WNM) and an extra-planar component \\citep{Dickey1990ARAAD} which is often called ``Lockman Layer''. The layer concept is based on the average emission from the extra-planar \\hi~ gas layer which is very faint. The clumpy nature of the \\hi~ gas implies then that extra-planar \\hi~ clouds must have a low volume filling factor. The layer concept describes therefore an ensemble of \\hi~ clouds or the probability distribution of such objects. First indications for a population of such clumps were found by \\citet{Simonson1971AA}. These early data came from the Dwingeloo telescope. Almost three decades later the Leiden/Dwingeloo survey (LDS,\\citep{Hartmann1997}) became available and provided a much improved database, more sensitive and essentially free of stray radiation. Channel maps show numerous clumps and filaments that are detached from the disk, and \\citet{KalberlaWestphalen1998} argued for an extra-planar gas layer which can be characterized by a distribution with a velocity dispersion of $\\sigma = 60$ \\kms, considerably larger than the dispersion suggested by \\citet{Lockman1991}. The first high resolution data of the extra-planar gas layer at a beam-width of 9\\arcmin~ have been taken by \\citet{Lockman2002} with the Robert C. Byrd Green Bank Telescope (GBT). These observations demonstrated convincingly the nature of the extra-planar gas layer as a population of cold clumps with a typical mass of 50 \\msun. Many of these clumps appear to be surrounded by warmer envelopes. A larger sample of clouds in the lower halo was studied by \\citet{Ford2008ApJ} with the Parkes Telescope. These clumps are somewhat larger and more massive than the sample detected by \\citet{Lockman2002}. This cloud population, located close to $R \\sim 3.8 $ kpc, is interpreted as originating from a Galactic fountain. Such a model would also explain the high kinetic energy which is needed for individual clouds to reach large $z$-distances. Support for such an interpretation comes from \\citet{Stil2006}. They found fast moving clumps in the Galactic plane with velocity vectors located {\\it within} the Galactic plane, analogous to fast velocities {\\it perpendicular} to the plane as suggested as an explanation for the extra-planar gas layer. So far we discussed predominantly Galactocentric distances $R \\la 8.5$ kpc, where 8.5 kpc is the I.A.U Sun--Galactic center distance, since most of the observations are in this range. For a more general description of this phenomenon, in particular for the question whether the extra-planar gas layer is caused by a fountain flow, objects at larger distances are needed. \\citet{Kalberla2008AA} argue that extra-planar gas is present even at $R \\ga 35$ kpc. Gas at such distances can hardly originate from fountain events. Direct evidence for a population of extra-planar \\hi~ clouds outside the Solar circle was first given by \\citet{Snezana2006}. Arecibo data in the direction towards the anti-center suggest that these clouds are not restricted to the inner part of the Milky Way disk, which is similar to preliminary results with the Effelsberg telescope reported by \\citet{Kalberla2005ASPC}. In the following we intend to explore the extra-planar gas layer of the outer part of the Milky Way in some more detail. Our results are based on single-dish observations with the Effelsberg 100-m radio telescope and on interferometer observations with the VLA and the WSRT array. This paper is organized as follows. In Sect. \\ref{sec_2} we explain our selection criteria for targets that have been mapped with the 100-m telescope. Our observations are described in Sect. \\ref{sec_3}. We discuss the properties of the \\hi~ cloud sample detected by us with the 100-m telescope, and also the results from two targeted interferometer observations in Sect. \\ref{sec_4}. We find evidence for a multi-phase structure and compare in Sect. \\ref{sec_6} the derived physical parameters with theoretical models. Sect. \\ref{sec_7} gives our summary and conclusion. ", "conclusions": "\\label{sec_7} We discussed a population of \\hi~ clouds residing in the lower halo of the Milky Way, co-rotating with the Galactic disk. The sample was observed with the 100-m Effelsberg telescope. Search criteria were angular sizes $s$, the brightness temperatures \\TB and line width \\FWHM~ which are considered to be typical for halo clumps. The sample includes \\hi~ clumps with the following properties: \\begin{itemize} \\item they reside in the outer galaxy with Galactocentric radii $R$ $ 10 < R < 15 $ kpc. \\item they belong to the lower halo ($ 0.9 < z < 5.4$ kpc). \\item the gas is cold, with a median \\TKIN $ \\sim 600$ K and a line width \\FWHM=5.3\\kms. \\item the sample shows a prominent two-component structure. Cold \\hi~ cores are surrounded by an extended component with broad line emission. \\end{itemize} Two of the most prominent \\hi~ clouds were observed using synthesis arrays, the WSRT and the VLA. These high-resolution observations resolve the clouds into a conglomeration of arc-minute sized \\hi~ cores. These cores are embedded in a more diffuse medium which is detectable only with single dish telescopes. The cores contain a significant fraction of the \\hi~ mass and tend to be in pressure equilibrium with the surrounding envelopes. Taking into account the influence of turbulence onto the line widths \\FWHM, the median line width values of 3.3 \\kms and 4.3 \\kms observed at the cores implies that the \\hi~ gas is very cold. Estimating densities and pressures for clumps and surrounding envelopes, we find some scattering but also a reasonable agreement with models which predict pressure equilibrium and a multi-phase structure caused by thermal instabilities \\citep{Wolfire2003}. The clumps tend to populate unstable regions in the phase diagrams, in agreement with recent predictions of turbulence driven instabilities \\citep{Audit2005,Gazol2005}. Comparing samples observed with big single dish telescopes e.g. GBT \\citep{Lockman2002}, Effelsberg (this work), and Parkes \\citep{Ford2008ApJ}, we find similar column densities, peak temperatures, line widths and masses. Our interferometer observations imply that some of the derived parameters may be heavily biased if the small scale structure observed by us may be considered as typical for \\hi~ halo clumps." }, "0911/0911.3905_arXiv.txt": { "abstract": "{} {We aim to demonstrate the usability of mm-wavelength imaging data obtained from the APEX-SZ bolometer array to derive the radial temperature profile of the hot intra-cluster gas out to radius $r_{500}$ and beyond. The goal is to study the physical properties of the intra-cluster gas by using a non-parametric de-projection method that is, aside from the assumption of spherical symmetry, free from modeling bias.} {We use publicly available X-ray spectroscopic-imaging data in the 0.7--2 keV energy band from the \\textit{XMM-Newton} observatory and our Sunyaev-Zel'dovich Effect (SZE) imaging data from the APEX-SZ experiment at 150 GHz to de-project the density and temperature profiles for a well-studied relaxed cluster, Abell 2204. We derive the gas density, temperature and entropy profiles assuming spherical symmetry, and obtain the total mass profile under the assumption of hydrostatic equilibrium. For comparison with X-ray spectroscopic temperature models, a re-analysis of recent \\textit{Chandra} observation is done with the latest calibration updates. We compare the results with that from an unrelaxed cluster, Abell 2163, to illustrate some differences between relaxed and merging systems.} {Using the non-parametric modeling, we demonstrate a decrease of gas temperature in the cluster outskirts, and also measure gas entropy profiles, both of which are done for the first time independently of X-ray spectroscopy using the SZE and X-ray imaging data. The gas entropy measurement in the central 100 kpc shows the usability of APEX-SZ data for inferring cluster dynamical states with this method. The contribution of the SZE systematic uncertainties in measuring $T_e$ at large radii is shown to be small compared to \\textit{XMM-Newton} and \\textit{Chandra} systematic spectroscopic errors. The total mass profile obtained using the hydrostatic equilibrium assumption is in agreement with the published X-ray and weak lensing results; the upper limit on $M_{200}$ derived from the non-parametric method is consistent with the NFW model prediction from weak lensing analysis.} {} ", "introduction": "Current cosmological models are built upon two complementary approaches of astronomical observation: the statistical study of the ensemble properties in a large sample of objects (i.e. from surveys) and the detailed analysis of the individual objects for gaining better understanding of the physical processes affecting those ensemble properties. This is particularly important in the study of galaxy clusters, where extraction of cosmological parameters from large survey samples (X-ray, optical, or in the radio/mm wavebands) relies critically on our understanding of different mass observables, which depends on the detailed physical processes affecting constituent gas and galaxies. Accurately determining the thermodynamic state of the intra-cluster medium (ICM) out to a large radius is critical for understanding the link between cluster mass and observables. For over a decade, observations of the thermal Sunyaev-Zel'dovich Effect (tSZE, hereafter simply SZE; Sunyaev \\& Zel'dovich 1970, Birkinshaw 1999) have been considered as a promising complement to X-ray observations for modeling the ICM in galaxy clusters, yet only recently has it been possible to make meaningful de-projections of gas temperature and density profiles using SZE imaging data from multi-pixel bolometer arrays, in combination with X-ray data. The APEX-SZ experiment (Dobbs et al. 2006, Halverson et al. 2009) employs one of the first such powerful multi-pixel Transition-Edge Sensor (TES) bolometer cameras, and a joint analysis of the ICM properties using SZE and X-ray data has been presented by Nord et al. (2009, Hereafter NBP09) for the massive cluster Abell 2163. In this paper we use the de-projection method used in NBP09 on the prototypical relaxed cluster Abell 2204. Our non-parametric analysis does not rely on any prior physical models in the construction of temperature and density profiles (apart from the assumption of spherical symmetry), hence the results are not based on parametric model fits. We measure the ICM entropy profile, as well as demonstrate the decrease of the ICM temperature in the cluster outskirts, first time from an SZE imaging data and independently from the X-ray spectroscopy. The derived ICM and cluster properties are compared with available X-ray and lensing results to highlight the level of accuracy of this independent method. Joint SZE/X-ray de-projection analysis is expected to become a standard tool in the near future for understanding the ICM physical state, as large numbers of resolved SZE maps will be available from the new generation SZE experiments. Our analysis assumes the gas to be in thermal equilibrium to model its physical properties, but presence of multi-phase ICM due to gas clumping will drive the electron temperature lower than the ion temperature in the electron-ion plasma (Evrard et al. 1996, Nagai et al. 2000). Recent hydro-simulations by Rudd \\& Nagai (2009) have shown, with a limited sample of halo models, that this deviation is small (about 5\\%) near $r_{200}$ for a relaxed cluster. Joint SZE/X-ray analysis using interferometric measurement of the SZE with OVRO/BIMA (Reese et al. 2002) has already shown that clumping effects are not large in the cluster interior (within $r_{500}$). Jia et al. (2008) have demonstrated the effect of the gas clumping on SZE and X-ray derived gas temperatures, and also found that these two quantities are in very good agreement within $r_{500}$ for the massive relaxed cluster RXC J2228.6$+$2036. But at large radii the gas should get clumpier, due to the onset of filamentary structures. One vital goal for sensitive imaging of the SZE signal using wide-field, multi-pixel bolometer cameras, and its combination with the X-ray and weak-lensing measurements, will be to provide an ultimate tool for measuring the gas clumping and thermodynamic state near the cluster virial radius, to give a dynamic view on the growth of clusters through accretion. \\subsection{Previous SZE/X-ray joint modeling} Due to the unavailability of resolved SZE images most of previous SZE/X-ray joint analysis studies have been limited to analytical or numerically simulated cluster models with idealized noise properties. Zaroubi et al. (2001) considered a method for reconstructing the triaxial structure of clusters based on Fourier slice theorem and applied it to a set of cluster simulations. Lee \\& Suto (2004) also considered de-projection method combining SZE and X-ray data and applied to analytical cluster models. Puchwein \\& Bartelman (2006) have employed the Richardson-Lucy de-projection technique to reconstruct the ICM and probe the dynamical state of clusters from simulations, and Ameglio et al. (2007) used a joint SZE/X-ray likelihood function maximization using a Monte Carlo Markov Chain (MCMC) for a similar objective. Modeling ICM properties from real SZE observations has been limited mainly to isothermal $\\beta$-models (Cavaliere \\& Fusco-Femiano 1978). Holzapfel et al. (1997), Hughes \\& Birkinshaw (1998) used isothermal models to constrain the Hubble parameter from observations of the clusters Abell 2163 and CL 0016$+$16, respectively, and later Reese et al. (2002) extended this analysis to a sample of 18 clusters detected by OVRO/BIMA. De Filippis et al. (2005) used published SZE decrement values and X-ray imaging data to constrain the triaxial structure of clusters using isothermal $\\beta$-models. Zhang \\& Wu (2000) similarly used the $\\beta$-model to combine SZE and X-ray data to derive central gas temperature in clusters. A more detailed parametric modeling has been done by Mahdavi et al. (2007) for the cluster Abell 478, using simultaneous fits to the X-ray, lensing and SZE data assuming parametric models for dark matter, gas and stellar mass distribution, and hydrostatic equilibrium. Yoshikawa \\& Suto (1999) first used Abel's integral inversion technique, originally proposed by Silk \\& White (1978), for a non-parametric reconstruction of radial density and temperature profiles using analytical and simulated cluster models. More recently Yuan et al. (2008) has extended this method for the most X-ray luminous cluster RXC J1347.5-1145 using published $\\beta$-model fit values from SZE and X-ray measurements. Extrapolation of the density and temperature profiles to the cluster outskirts based on such parametric modeling can be problematic, in particular for clusters with a very peaked central emission such as RXC J1347.5-1145. Additionally, this cluster is considered to be a merging system (Cohen \\& Kneib 2002) where the assumptions of spherical symmetry and hydrostatic equilibrium may not be valid. The nearest approach to non-parametric modeling was made by Kitayama et al. (2004) for the same cluster, RXC J1347.5-1145, using a beta-model density profile to fit the X-ray surface brightness and obtaining fitted temperature values separately in each radial bin from their SZE imaging data. The small extent of their SZE map (less than 2 arcmin) limited the temperature modeling again to the cluster core region. \\subsection{Scope of the present work} In this paper we apply the non-parametric ICM modeling based on Abel's integral inversion technique, as presented in NBP09, to the well studied and dynamically relaxed galaxy cluster Abell 2204 ($z=0.1523$, $L_{\\mathrm{X}}=26.9\\times 10^{44}$ $h_{50}^{-2}$ ~erg s$^{-1}$ in the $0.1-2.4$ keV band, $T_{\\mathrm{X}} = 7.21 \\pm 0.25$ keV; Reiprich \\& B\\\"ohringer 2002). The only assumptions in this analysis are spherical symmetry for reconstructing temperature and density profiles, and hydrostatic equilibrium (HSE) for reconstructing the total mass profile. The primary aim is to confirm the validity of this method for modeling the ICM distribution and cluster mass -- and compare the results with those obtained from deep X-ray spectroscopic and weak lensing data -- in a cluster where the assumptions of spherical symmetry and HSE are generally accepted to be valid. We compute the \\textit{Chandra} spectral temperature profile with the latest calibration updates and compare it with the SZE-derived temperature profile. In contrast to the X-ray spectroscopic measurements from \\textit{Chandra}, the SZE-derived ICM temperature measurements near the cluster virial radius are constrained primarily by the statistical uncertainties in the SZE data. This fact demonstrates the potential for stacking the SZE signal of several relaxed clusters to put tighter constraints on the slope of the gas temperature profile in the cluster outskirts (Basu et al., in preparation). For a single cluster (Abell 2204), the depth in the APEX-SZ map allows us to model the temperature profile with meaningful errors up to $\\sim80\\%$ of the cluster virial radius (which we take to be $r_{200}$, the radius within which the mean total density is 200 times the critical density). From density and temperature profiles we derive other physical properties like total gravitational mass, gas mass fraction and the gas entropy index. The total mass modeling provides a quantitative comparison with the published X-ray and lensing results. The modeling of the gas entropy profile from SZE/X-ray imaging data is a first, and we compare the central entropy values of two clusters with different morphologies, A2204 and A2163 (APEX-SZ analysis of the latter was presented in NBP09). This comparison shows how the gas entropy in the cluster core derived from SZE/X-ray joint modeling can be used to infer the dynamical state of clusters without the need for X-ray spectroscopy. A further comparison of the baryonic fraction of the ICM between A2204 and two other dynamically complex clusters detected by APEX-SZ (Bullet and A2163) illustrates a statistically significant difference of $f_{\\mathrm{gas}}$ inside $r_{2500}$. All the scientific results in this paper are computed from the radial profiles of two observables: the SZE temperature decrement at 150 GHz, and the X-ray surface brightness in the $0.7-2$ keV band of XMM-Newton. In \\S2 and \\S3, we describe the map making and radial profile extraction steps from the X-ray and SZE data, and discuss the different systematic uncertainties associated with each profile. \\S4 describes Abel's integral inversion method and presents our primary results in the form of the radial density and temperature profiles. In \\S5 we present the other derived quantities like gas entropy and the total cluster mass profiles, and list the conclusions in \\S6. We use the currently favored $\\Lambda$CDM cosmology with the following parameters: $\\Omega_m=0.27$, $\\Omega_{\\Lambda}=1-\\Omega_m=0.73$, and the Hubble constant $H_0=70$ km s$^{-1}$ Mpc$^{-1}$. At redshift of $z=0.1523$, the angular diameter distance of Abell 2204 is 541.6 Mpc. To put the radial profiles in perspective using the characteristic cluster radii, we adopted the maximum likelihood NFW fit parameters from Corless et al. (2009), $M_{200} = 7.1\\times 10^{14} M_{\\odot}$ and $c=4.5$, which gives $r_{200}=1.76$ Mpc ($11.2'$), $r_{500}=1.16$ Mpc ($7.3'$) and $r_{2500}=0.51$ Mpc ($3.2'$). ", "conclusions": "\\begin{enumerate} \\item We describe the detailed application of a direct, nonparametric de-projection method of cluster density and temperature profiles, using APEX-SZ and XMM-Newton data. The method was presented in NBP09, the current paper builds upon the previous work by applying this technique to the well-studied relaxed cluster Abell 2204. \\item Analysis of both SZE and X-ray data have been done from their raw data sets, to create images and radial profile. In particular, we describe the creation of a set of SZE radial profiles, all consistent with the APEX-SZ measurement, to characterize the statistical uncertainties on the bin values and minimize the numerical errors in Abel's de-projection method. Our final results are dominated by the statistical uncertainties in the SZE data, the signal at $r_{200}$ is essentially an upper limit for A2204. We describe the different sources of systematic uncertainties and include them in the analysis. \\item The decreasing gas temperature in the cluster outskirts is demonstrated for the first time from SZE measurements, using a broad re-binning of the APEX-SZ (and X-ray) data. The temperature drop can be confirmed to $98\\%$ confidence level. We also compare the direct de-projected pressure profile with some parametric models, and show that the Nagai profile is adequate for modeling the gas pressure, within the current statistical uncertainties in APEX-SZ imaging of a single cluster. \\item We re-perform the X-ray spectral analysis for the ICM temperature profile from publicly available \\textit{Chandra} data, primarily to find the changes from the recent calibration updates (CALDB 4), but also to show the effect of systematic uncertainties due to the background modeling in the X-ray spectral analysis. A comparison with the projected temperature profile obtained from SZE data confirms that our SZE derived temperature values are much less affected by systematic uncertainties at large radii, in comparison with \\textit{Chandra} and \\textit{XMM-Newton}. Precise comparison between the SZE and X-ray spectroscopic measurements of the gas temperature in the cluster outskirts will be a promising method to constrain gas clumping and non-LTE effects. \\item The integrated total mass profile is computed assuming hydrostatic equilibrium for the cluster gas. The mass profile is in excellent agreement with the recent X-ray and weak lensing analyses. Our model prediction for $M_{500}$ is $(2.6\\pm 2.2) \\times 10^{14}~h^{-1} M_{\\mathrm{\\odot}}$. This is somewhat lower than the X-ray and lensing results but consistent within $1\\sigma$ errors. The upper limit on $M_{200}$ from our analysis is in good agreement with the published NFW model fit from weak lensing analysis of A2204. \\item The ICM mass fraction as function of radius is computed using the non-parametric modeling, and found to be below 0.1 within $r_{500}$. The low $f_{\\mathrm{gas}}$ value in A2204 in the cluster center is in contrast with the previous APEX-SZ measurement of this ratio for Abell 2163 and the Bullet cluster. \\item We compute the ICM entropy profile from SZE/X-ray joint analysis and confirm the general agreement with the self-similar cluster model predictions within the present statistical uncertainties. The significance of the APEX-SZ measurement of A2204 is not sufficiently high at $r\\gtrsim r_{500}$ to constrain the slope of the entropy profile in the cluster outskirts. \\item We compare the entropy profiles of Abell 2204 (relaxed) and Abell 2163 (merging system), using the same non-parametric SZE/X-ray de-projection and radial binning, and find a clear entropy difference in their central 200 kpc. This corresponds to the different dynamical states of these two clusters and seen for the first time from SZE derived $T_{\\mathrm{gas}}$ measurement. \\end{enumerate}" }, "0911/0911.1307_arXiv.txt": { "abstract": "The Phase-Induced Amplitude Apodization (PIAA) coronagraph is a high performance coronagraph concept able to work at small angular separation with little loss in throughput. We present results obtained with a laboratory PIAA system including active wavefront control. The system has a 94.3\\% throughput (excluding coating losses) and operates in air with monochromatic light. Our testbed achieved a 2.27 $10^{-7}$ raw contrast between 1.65 $\\lambda/D$ (inner working angle of the coronagraph configuration tested) and 4.4 $\\lambda/D$ (outer working angle). Through careful calibration, we were able to separate this residual light into a dynamic coherent component (turbulence, vibrations) at 4.5 $10^{-8}$ contrast and a static incoherent component (ghosts and/or polarization missmatch) at 1.6 $10^{-7}$ contrast. Pointing errors are controlled at the $10^{-3}$ $\\lambda/D$ level using a dedicated low order wavefront sensor. While not sufficient for direct imaging of Earth-like planets from space, the 2.27 $10^{-7}$ raw contrast achieved already exceeds requirements for a ground-based Extreme Adaptive Optics system aimed at direct detection of more massive exoplanets. We show that over a 4hr long period, averaged wavefront errors have been controlled to the 3.5 $10^{-9}$ contrast level. This result is particularly encouraging for ground based Extreme-AO systems relying on long term stability and absence of static wavefront errors to recover planets much fainter than the fast boiling speckle halo. ", "introduction": "\\label{sec:intro} An imaging system aimed at dection or characterization (spectroscopy) of exoplanets must overcome the large contrast beween the planet and its star. This is particularly challenging for Earth-like planets, where the contrast is $\\approx 10^{-10}$ in the visible and the angular separation is 0.1\\arcsec for a system at 10pc. Many coronagraph concepts have recently been proposed to overcome this challenge (see review by \\cite{guyo06}). Among the approaches suggested, Phase-Induced Amplitude Apodization (PIAA) coronagraphy is particularly attractive. In a PIAA coronagraph, aspheric optics (mirrors or lenses) apodize the telescope beam with no loss in throughput. A PIAA coronagraph combines high throughput, small inner working angle (2 $\\lambda/D$ for $10^{-10}$ contrast), low chromaticity (when mirrors are used), full 360 degree discovery space, and full $1 \\lambda/D$ angular resolution. Angular resolution (size of the planet's PSF in the image) is a critical performance parameter as exoplanet imaging sensitivity, even if speckles have been perfectly removed, is usually background-limited due to sky or thermal emission (near-IR and mid-IR imaging from the ground) or zodiacal and exozodiacal light (direct imaging of Earth-like planets in the visible from space). The PIAA concept, orginally formulated by \\cite{guyo03}, has since been studied in depth in several subsequent publications \\citep{trau03,guyo05,vand05,gali05,mart06,vand06,pluz06,guyo06,beli06,guyo09,lozi09}, which the reader can refer to for detailed technical information. In the first laboratory demonstration of the PIAA concept \\citep{gali05}, lossless beam apodization was demonstrated, and the field aberrations introduced by the PIAA optics were confirmed experimentally. In this first prototype, the PIAA acrilic optics lacked surface accuracy required for high contrast imaging, and since this experiment did not include active wavefront control, the high contrast imaging potential of the technique could not be demonstrated. In the present paper, we report on results obtained with a new system which includes reflective PIAA optics and wavefront control. Our prototype combines the main elements/subsystems envisionned for a successful PIAA imaging coronagraph instrument, with the exception of corrective optics required to remove the strong off-axis aberrations introduced by the PIAA optics. This last subsystem has been designed and built for another testbed, and its laboratory performance is reported in a separate paper \\citep{lozi09}. The overall system architecture adopted for our test is presented and justified in \\S\\ref{sec:archi}. The design of the main components of the coronagraphs (PIAA mirrors, masks) is also described in this section. Wavefront control and calibration are discussed in \\S\\ref{sec:wfc}. Laboratory results are presented in \\S\\ref{sec:labresults}. \\begin{figure*}[htb] \\includegraphics[scale=0.6]{f1.eps} \\caption{\\label{fig:optlayout} Optical layout of the laboratory PIAA coronagraph system. The grey shaded area shows the rigid PIAA bench on which the two PIAA mirrors are mounted. The light source is at the upper corner of the figure. The focal plane mask (near the bottom, center) separates light into the imaging channel and the coronagraphic low order wavefront sensor (CLOWFS) channel.} \\end{figure*} ", "conclusions": "The results obtained in this experiment are especially encouraging for ground-based coronagraphy. The 2 $10^{-7}$ raw contrast we have achieved in the 1.65 to 4.4 $\\lambda/D$ angular separation range already exceeds by two orders of magnitudes the raw contrast that can be hoped for in even a theoretically ideal Extreme-AO system \\citep{guyo05}. We note that with a more careful optical design and anti-reflection coated optics, our experiment could probably have reached 5 $10^{-8}$ raw contrast (level of coherent light leak in the current experiment). More importantly, we have demonstrated that with the coronagraph + wavefront control architecture adopted in our experiment, static coherent speckles can be pushed down very low (3.5 $10^{-9}$) in long exposures. Our system successfully removed long term correlations in the coherent speckles, and their averaged level in long exposure was reduced with a $1/\\sqrt{T}$ law. The combination of a high performance PIAA coronagraph and a focal-plane based wavefront control therefore appears extremely attractive for ground-based Extreme-AO. In that regard, our experiment has been a successful validation of the key technologies and control algorithms of the Subaru Coronagraphic Extreme-AO (SCExAO) system currently in assembly. The major differences between the SCExAO PIAA coronagraph and our laboratory prototype are (1) the need to design and operate a PIAA coronagraph on a centrally obscured pupil with thick spider vanes and (2) the need for corrective optics to recover a wide field of view. These two requirements have been validated in a separate laboratory experiment using the final SCExAO coronagraph optics \\citep{lozi09}. Our experiment was limited at the 2 $10^{-7}$ contrast by an optical ghost and at the 5 $10^{-8}$ contrast by turbulence or vibrations. The PIAA coronagraph could therefore not be tested to the contrast level required for direct imaging of Earth-like planets from space (approximately $10^{-10}$), although several key concepts were demonstrated, including simultaneous operation of a low-order wavefront sensor using starlight in the PSF core and high-order wavefront sensor using scattered light in the science focal plane. New calibration schemes which will be very useful for high contrast coronagraphy were also developed and validated, such as the system response matrix optimization loop, which can executed as a background task to fine-tune the system. PIAA coronagraph technologies for high contrast space applications are now being actively developed and tested at NASA Ames Research Center and NASA Jet Propulsion Laboratory. A new set of PIAA mirrors was recently manufactured to higher surface accuracy than the ones used for this demonstration, and is being integrated within the High Contrast Imaging Testbed (HCIT) vacuum chamber at NASA Jet Propulsion Laboratory. We note that the HCIT chamber has already demonstrated stability to the $10^{-10}$ contrast with a Lyot-type coronagraph \\citep{trau07}. The experiment described in this paper served as a precursor to this new step, which is aimed at reaching higher contrast (minimum goal of $10^{-9}$) in broadband light using a two-DM wavefront correction. In parallel to this effort, a highly flexible high stability air testbed at NASA Ames Research Center is coming online to further explore technology and architecture trades for PIAA systems. In addition to pushing the contrast further, future laboratory demonstration of the PIAA coronagraph will explore chromaticity and wavefront control issues unique to a PIAA coronagraph architecture. The monochromatic experiment described in this paper did not address these important points, and should therefore be considered as only a first step toward validation of the PIAA coronagraph technique for high contrast space-based exoplanet imaging mission." }, "0911/0911.4714_arXiv.txt": { "abstract": "s{ We report on the observations of cosmic rays with energies $ \\geq $ $10^{18}$ eV from Jan 2004 to April 2009 by the Pierre Auger Observatory. During this period the Observatory has grown from about 300 surface detectors to about 1600 upon its completion in November 2008. The 1600 surface detectors are overlooked by 24 fluorescence telescopes. We report on measurements of the cosmic ray spectrum, the arrival directions and the elongation rate. We also report limits for the photon and neutrino components of this cosmic radiation.} \\begin{figure} \\begin{minipage}[b]{0.47\\linewidth} \\includegraphics [width=3.0in] {blois1.jpg} \\label{fig:blois1.jpg} \\end{minipage} \\hfill \\begin{minipage}[b]{0.47\\linewidth} \\includegraphics [width=3.0in] {blois2.jpg} \\label{fig:blois2.jpg} \\end{minipage}\\\\ \\begin{minipage}[t]{0.47\\linewidth} \\caption{\\it Cosmic ray spectrum (data compiled by Simon Swordy, University of Chicago).} \\end{minipage} \\hfill \\begin{minipage}[t]{0.47\\linewidth} \\caption{\\it GZK horizons for protons and nuclei (figure courtesy Denis Allard).} \\end{minipage} \\end{figure} ", "introduction": "The cosmic radiation discovered by Hess \\cite{Hess} extends from very low energies $\\leq$ 10$^{6}$ eV to $\\geq$ 10$^{20}$ eV. The latter energy is equal to 16 joules - a macroscopic energy in a microscopic particle as the cosmic rays are principally atomic nuclei ranging from protons to iron. Figure 1 shows the full cosmic ray spectrum. It is roughly a power law falling by 30 orders of magnitude in flux over 10 orders of magnitude increase in energy. The upper end of the spectrum represents a mystery as there is no clear understanding of how Nature can accelerate atomic nuclei to such high energies. The study of this category of cosmic rays is a scientific imperative and Nature provides two important analytical tools for the investigation. First, the very highest energy cosmic rays must come from nearby. Consequently one can expect that there are a small number of sources that can contribute to the flux of the highest energy cosmic rays. Protons interact with the cosmic microwave background (CMB) losing energy while producing pions. This is the famous GZK effect. Complex nuclei are photo disintegrated by the CMB. The result of these interactions is that half of the cosmic rays with energy $\\geq$ 6x10$^{19}$ eV must come from distances less than 70 Mpc \\cite{Allard}. In Figure 2 we show the expected distribution of distances for several nuclear species on the basis of a uniform source distribution. It is noteworthy that for distances $\\geq$ 50 Mpc only protons and iron nuclei survive. In composition analysis at these high energies the assumption of two components is more than just an ansatz - it is a reasonable assumption. Second, the higher energies and shorter distances will reduce the effects of the random magnetic fields which at lower energies decouple the observed arrival direction from the true direction of the source. Thus it is quite possible that the arrival directions for energies $\\geq$ 6x10$^{19}$ will correlate with the distribution of extragalactic objects located within 100 Mpc. \\begin{figure} \\begin{minipage}[b]{0.47\\linewidth} \\includegraphics [width=2.0in] {blois3.jpg} \\label{fig:blois3.jpg} \\end{minipage} \\hfill \\begin{minipage}[b]{0.47\\linewidth} \\includegraphics [width=3.0in] {blois4.jpg} \\label{fig:blois4.jpg} \\end{minipage}\\\\ \\begin{minipage}[t]{0.47\\linewidth} \\caption{\\it Cartoon showing the two techniques for detection of air showers (figure courtesy Enrique Zas .} \\end{minipage} \\hfill \\begin{minipage}[t]{0.47\\linewidth} \\caption{\\it Cosmic ray spectra from HiRes and AGASA circa 2004 .} \\end{minipage} \\end{figure} ", "conclusions": "" }, "0911/0911.2768_arXiv.txt": { "abstract": "The preliminary results of a three-site CCD photometric campaign are reported. The $\\delta$ Scuti variable V650 Tauri belonging to the Pleiades cluster was observed photometrically for 14 days on three continents during 2008 November. An overall run of 164 hr of data was collected. At least five significant frequencies for V650 Tauri have been detected. ", "introduction": "$\\delta$ Scuti variables 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. These variables are thought to be excellent laboratories for probing the internal structure of intermediate mass stars. Intents of modelling $\\delta$ Scuti stars belonging to open clusters have been performed recently (e.g. [1], [2], [3]). Although the constraints imposed by the cluster parameters have proved to be very useful when modelling an ensemble of $\\delta$ Scuti stars, more detected frequencies in individual stars would improve current seismic studies. \\bigskip The target star V650 Tau (HD 23643, $V=7^{\\rm m}.79$, A7) was identified as a short-period pulsating variable by Breger (1972). Intensive observations performed by the STEPHI network in November 1990, revealed four frequency peaks in V650 Tau [4]. One-site CCD photometric observations carried out by [5] in November-December 1993, confirmed the results obtained by the STEPHI campaign. Since then, no new observations of V650 Tauri have been performed. \\bigskip The present paper provides preliminary observational results of a three-site campaign on V650 Tauri in 2008. ", "conclusions": "" }, "0911/0911.2597_arXiv.txt": { "abstract": "It is commonly stated that we have entered the era of precision cosmology in which a number of important observations have reached a degree of precision, and a level of agreement with theory, that is comparable with many Earth-based physics experiments. One of the consequences is the need to examine at what point our usual, well-worn assumption of homogeneity associated to the use of perturbation theory begins to compromise the accuracy of our models. It is now a widely accepted fact that the effect of the inhomogeneities observed in the Universe cannot be ignored when one wants to construct an accurate cosmological model. Well-established physics can explain several of the observed phenomena without introducing highly speculative elements, like dark matter, dark energy, exponential expansion at densities never attained in any experiment (i.e. inflation), and the like. Two main classes of methods are currently used to deal with these issues. Averaging, sometimes linked to fitting procedures a la Stoegger and Ellis, provide us with one promising way of solving the problem. Another approach is the use of exact inhomogeneous solutions of General Relativity. This will be developed here. ", "introduction": "\\label{int} The observation of its structures show that our Universe is not homogeneous. We see voids, groups of galaxies, clusters, superclusters, walls, filaments, etc. However, it is usually argued in the literature that the Universe should be nearly homogeneous at large scales which is supposed to validate the use of Friedmannian models. But how large these scales are and what nearly does imply is never precisely stated. It has however become, during the last few years, as a widely accepted fact, that the effect of the inhomogeneities cannot be ignored when one wants to construct an accurate cosmological model up to the regions where structures start forming and their evolution becomes non-linear. Three different methods have been proposed to deal with this issue: \\begin{enumerate} \\item Linear perturbation theory. However, this method is only valid when {\\it both} the curvature and density contrasts remain small, which is not the case in the non-linear regime of structure formation and where the SNe Ia are observed. \\item Averaging methods `{\\`a} la Buchert', promising, but needing to be improved (see \\cite{TB08} and references therein). \\item Exact inhomogeneous solutions, valid at all scales and exact perturbations of the Friedmann background which they can reproduce as a limit with any precision. \\end{enumerate} The use of such exact solutions shows that well-established physics can explain several of the phenomena observed in astrophysics and cosmology without introducing highly speculative elements, like dark matter, dark energy, exponential expansion at densities never attained in any experiment (inflation), and the like. Here, we foccuss on the application of a couple of exact solutions of general relativity to structure formation and evolution and to the reproduction of cosmological data. ", "conclusions": "The increasing precision of observational data implies that FLRW models must now be considered just a zeroth order approximation, and linear perturbation theory a first order approximation whose domain of validity is an early, nearly homogeneous Universe. In the nonlinear regime, which was entered since structures formed, there is no escape from the use of exact methods (or of averaging schemes aiming at investigating this issue from the standpoint of backreaction). In the era of `precision cosmology', the effect of the inhomogeneities on the determination of the cosmological models cannot be ignored. Inhomogeneous models constitute an exact perturbation of the Friedmann background and can reproduce it as a limit with any precision. This is the reason why they are fully adapted for the purpose of studying astrophysical and cosmological effects and for constructing precise models of universe. While using L--T models with a central observer to represent our `local' Universe averaged over angles around us, a giant void is not mandatory to explain away dark energy. A giant overdensity can also do the job. However while neither the void nor the overdensity are directly observable, the giant void alone can be tested with observations of the density function on our past light cone. The giant hump will need more and more precise data to be constrained. Exact inhomogeneous solutions can be employed not only for studying the geometry and dynamics of the Universe, but also to investigate the formation and the evolution of structures. They give enhanced formation efficiency and might therefore help solving the problem of structure formation pertaining to the standard model. While the L--T models have been mostly used up to now for modeling the inhomogeneities of the Universe, the need of getting rid of spherical symmetry for this purpose will lead the cosmological community to consider other solutions, and among them QSS models, for the future developments of inhomogeneous cosmology." }, "0911/0911.4843_arXiv.txt": { "abstract": "Previous velocity images which reveal flows of ionized gas along the most prominent cometary tail (from Knot 38) in the Helix planetary nebula are compared with that taken at optical wavelengths with the Hubble Space Telescope and with an image in the emission from molecular hydrogen. The flows from the second most prominent tail from Knot 14 are also considered. The kinematics of the tail from the more complex Knot 32, shown here for the first time, also reveals an acceleration away from the central star. All of the tails are explained as accelerating ionized flows of ablated material driven by the previous, mildly supersonic, AGB wind from the central star. The longest tail of ionized gas, even though formed by this mechanism in a very clumpy medium, as revealed by the emission from molecular hydrogen, appears to be a coherent outflowing feature. ", "introduction": "Dense, neutral (\\citealt{mea92}; \\citealt{hug92}) knots with ionized cometary tails are found in the central regions of planetary nebulae (PNe); the most famous being those in the Helix nebula for, at a distance of only 213 $+$30/$-$16 pc \\citep{har07}, they were easily detected and resolved by early ground--based observations (Baade and reported by \\citealt{vor68}). The incredibly clumpy nature of the neutral material in the disk of the Helix nebula (when modelled as a bi--polar PNe viewed along its axis, \\citealt{mea98}; \\citealt{mea05}) has now been revealed in spectacular fashion in the imagery in the H$_{2}$ emission line by \\citet{mat09}. Previously, \\citet{mei05} had estimated that there were 23,000 cometary knots and that inevitably tail interactions must be occurring. \\citet{mat09} show conclusively that it is an inner region in the Helix disk, towards the central star, where tails are observed from neutral globules surrounded by an outer clumpy region free of such tails. \\citet{dys03} reviewed the two broad models for the creation of the cometary tails; either they are shadowed from the ionizing radiation of the central star, and their surfaces photo-ionized by scattered Lyman photons in the nebula, or they are dynamically produced as the particle winds from the central star swept past the slowly expanding system of dense globules. A critical distinction between these models arises if the cometary tails can be seen to be flowing along their lengths away from their parent globules. Only the `dynamic' model would have this effect. In fact, a numerical simulation of the flow of a moderately supersonic particle wind past the ionized head of a globule \\citep{dys06} predicted not only the creation of a cometary tail but a moderate acceleration of ionized material parallel to the tail surface and away from the globule. In this initial model the neutral material away from the globule had a smooth density distribution. The most comprehensive optical study of the kinematics of the Helix system of globules remains that described in \\citet{mea98}. Here, spatially resolved \\nii\\ profiles, from 300 separate long-slit (163\\arcsec\\ long) positions, were obtained with the Manchester echelle spectrometer (MES -- \\citealt{mea84} but now with a CCD as the detector) on the Anglo--Australian telescope in exceptional `seeing' conditions, over three separate `blocks' of globules and their tails in the nebular core. Furthermore, a kinematical `case study' of the globule which is apparently the 2nd closest to the nebular core (Knot 14), and a length of its tail, was carried out in a range of emission lines. The principal outcome was to show that the system of central globules is concentrated in a disk expanding itself at 14\\kms. Flows parallel to the surfaces of the cometary tails of two of the most prominent globules (Knots 38 and 14) were also detected and even an acceleration of this flow along the length of the longest tail (from Knot 38) suggested. \\citet{ode07} challenged these latter assertions without making any further kinematical observations but simply because they claim that more recent HST images reveal confusing minor globules in the longest tail (Knot 38). With the dismissal of these kinematical effects they proceeded to support the shadowing theories of the creation of the tails. The aim of the present paper is to reassess the strength of the deductions from the \\citet{mea98} kinematical data set in the light of the subsequent HST and very latest H$_{2}$ imagery by \\citet{mat09}. In particular, to consider if the evidence in \\citet{mea98} for flows along the tail surfaces of Knots 38 and 14, away from the central star, is now invalidated by this HST imagery alone as suggested by \\citet{ode07}. Furthermore, the flow behind Knot 32, which was not considered hitherto, is now presented to strengthen the original suggestion that accelerating flows in the cometary tails could be ubiquitous though in a clumpy medium. \\section[]{Knots 38 and 14} Knot 38 has the longest (62\\arcsec) cometary tail and is the apparently closest knot to the central star of NGC 7293, while Knot 14 is the next closest. The heads of both have arcs of \\oiii\\ emission facing this star which indicates that they protrude into the hard radiation field of the central volume of the bi--polar, ellipsoidally--shaped, nebula that produces the characteristic helical appearance in emission lines of lower ionization species (\\citealt{mea98}; \\citealt{mea05}). The HST archival images (PI NAME: Meixner, PID:9700) of Knots 38 (HST ACS/WFC J8KR14040) and 14 (HST ACS/WFC J8KR0840) are presented respectively in Figs. 1a and 2. These should be compared with those taken with the New Technology Telescope (NNT - Chile) and presented in Meaburn et al (1998). All are in the light of the \\ha\\ plus \\NII\\ nebular emission lines. The uniquely long ionized tail of Knot 38 in Fig. 1 is prominent in both the HST and former \\citep{mea98} NTT images and appears to be a coherent structure in both i.e. it is not simply a consequence of chance superposition of a large number of fore-- or background tails along the same sightlines. Minor ionized knots appear along the tail in the image in Fig. 1a but it is the H$_{2}$ 2.12$\\mu$m image \\citep{mat09} shown in Fig. 1b that emphasises that the tail of ionized gas, being so long, is formed around an internal core composed of a large number of minor clumps of neutral material. With this as a starting point the kinematics along this tail should be re--considered by examining the \\nii\\ line profiles presented in \\citet{mea98}. The centroids of these profiles are shown in fig. 13 of that paper to change along the 62\\arcsec\\ length of this tail by a radial velocity difference (from central knot to tail end) by 10\\kms. The velocity images in figs. 4 and 12 of the same paper confirm this systematic radial velocity change in a different way. Unfortunately, the four images in different heliocentric radial velocity (\\vhel) ranges became jumbled in the production of fig. 12 in the \\citet{mea98} paper. These should be \\vhel\\ $= -$31 to $-$27 (top right), $-$25 to $-$21 (top left), $-$20 to $-$15 (bottom right) and $-$14 to $-$10 (bottom left) and all \\kms. When these four images are considered with this correction it is clear that the head of Knot 38 appears alone in the top right frame then progressively the tail appears in the subsequent three frames towards more positive velocities as the image of the head declines. The acceleration along the tail length, if regarded as a coherent feature starting at Knot 38, is very clear and not dominated by the minor confusing knots apparent in Fig. 1a. Those areas free of these along the tail length of Knot 38 in Fig. 1 clearly show this radial velocity change. The appearance of the neutral material in the tail of Knot 38 and seen in Fig. 1b suggests that the ionized outflow is in a sheath around a clumpy neutral core. Similarly, the \\nii\\ line profiles up to 5\\arcsec\\ from the head of Knot 14 (fig. 9 of \\citealt{mea98}) show a systematic change of radial velocity of their centroids of -6 \\kms. This was not covered by the H$_{2}$ imagery of \\citep{mat09}. The tail from Knot 14 was modelled kinematically as a flow parallel to the globule and tail surfaces. The HST images in Fig. 2 show that there are no significant minor ionized knots along this small length of the tail. Again an accelerating flow away from the central star is indicated. If tilted at 25\\degree\\ to the plane of the sky this change of outflow velocity amounts to 14 \\kms\\ compared to the apex of the head of the cometary knot. ", "conclusions": "" }, "0911/0911.4591_arXiv.txt": { "abstract": "Apart from the 11-year solar cycle, another periodicity around 155-160 days was discovered during solar cycle 21 in high energy solar flares, and its presence in sunspot areas and strong magnetic flux has been also reported. This periodicity has an elusive and enigmatic character, since it usually appears only near the maxima of solar cycles, and seems to be related with a periodic emergence of strong magnetic flux at the solar surface. Therefore, it is probably connected with the tachocline, a thin layer located near the base of the solar convection zone, where strong dynamo magnetic field is stored. We study the dynamics of Rossby waves in the tachocline in the presence of a toroidal magnetic field and latitudinal differential rotation. Our analysis shows that the magnetic Rossby waves are generally unstable and that the growth rates are sensitive to the magnetic field strength and to the latitudinal differential rotation parameters. Variation of the differential rotation and the magnetic field strength throughout the solar cycle enhance the growth rate of a particular harmonic in the upper part of the tachocline around the maximum of the solar cycle. This harmonic is symmetric with respect to the equator and has a period of 155-160 days. A rapid increase of the wave amplitude could give place to a magnetic flux emergence leading to observed periodicities in solar activity indicators related with magnetic flux. ", "introduction": "During solar cycle 21, a short periodicity between 152--158 days was discovered in $\\gamma$ ray flares \\citep {Rieger84}, X ray flares \\citep{Rieger84, Dennis85, Bai87, Kile91, Dimitropoulou08}, flares producing energetic interplanetary electrons \\citep{Droge90}, type II and IV radio bursts \\citep{Verma91}, and microwave flares \\citep{Bai85, Kile91}. However, this periodicity was absent during solar cycle 22 \\citep{Kile91, Bai92a, ozguc94}. The periodicity has also been detected in indicators of solar activity (sunspot blocking function, sunspot areas, ``active'' sunspot groups, group sunspot numbers) which suggest that it is associated preferentially with photospheric regions of compact magnetic field structures \\citep{Lean89, Lean90, pap90, carball90, Bou92, carball92, Verma92, oliver98, ballester99, Krivo02}. Probably, the most important, and enigmatic, feature of the periodicity is that it appears during epochs of maximum activity and that it occurs in episodes of 1 to 3 years. \\citet{rabin91} performed a study of the magnetic flux variations during solar cycle 21 which reveals the existence of quasi-periodic pulses or episodes of enhanced magnetic activity. The duration of the pulses is $\\approx$ 5 rotations during the years around maximum activity, the epoch in which the flare periodicity appears, and the comparison with magnetic field maps indicates that those pulses of activity correspond to the occurrence of complex active regions containing large sunspots \\citep{Bai87a}. \\citet{ballester02, ballester04} analyzed several data sets of, or strongly related to, photospheric magnetic flux to point out that the appearance of the near 160-day periodicity in different manifestations of solar activity during solar cycle 21 has its underlying cause in the appearance of the periodicity in the magnetic flux linked to regions of strong magnetic field. They also showed that during solar cycle 22 the periodicity does not appear in the photospheric magnetic flux records and, as a consequence, the periodicity did not appear in other solar activity indicators, while during solar cycle 23 it appeared in the photospheric magnetic flux but not in other solar activity indicators. Several mechanisms have been put forward in order to explain the existence of this periodicity. \\citet{wolff83} linked it to the interaction of rotating features (active longitude bands) resulting from g-modes with $l=2$ and $l=3$. \\citet{Bai87b} suggested that the cause of this periodicity must be a mechanism that causes active regions to be more flare productive. Later, \\citet{Bai87} concluded that it cannot be due to the interaction of \\lq \\lq hot spots\", i.e. regions where flare activity is higher than elsewhere \\citep{Bai87a, Bai88}, rotating at different rates and that the cause must be a mechanism involving the whole Sun. \\citet{Ichi85} suggested that it is related to the timescale for storage and/or escape of magnetic fields in the solar convection zone. \\citet{B90}, taking into account the possible intermittency of the periodicity, suggested that this behavior could be simulated with a damped, periodically forced non-linear oscillator, which shows periodic behavior for some values of the parameters and chaotic behavior for other values. \\citet{wolff92} argued that such periodicity can be understood in terms of the normal modes of oscillation of a nearly spherical, slowly rotating star, when two r-modes (inertial modes) couple with an interior g-mode beat. This suggestion seems to agree qualitatively with the fact that the periodicity is stronger around the activity maximum. \\citet{Bai91} and \\citet{stu92} proposed that the Sun contains a \\lq \\lq clock\", modeled by an oblique rotator or oscillator, with a period of 25.8 days and suggested that the periodicity of 154 days is just a subharmonic of that fundamental period. Later, \\citet{Bai93} modified the earlier period to the value 25.50 days, but that model seems to be very constrained by helioseismological data about the rotation of the Sun's interior. \\citet{lou00} suggested that such periodicities can be related to large-scale equatorially trapped Rossby-type waves showing that, for typical solar parameters, the periods of these waves (with n = 1 and m even) are in good agreement with the observed ones. Moreover, \\citet{lou00} has also pointed out that such waves can give rise to detectable features, such as surface elevations in the photosphere. Coincidently, \\citet{kuhn00} have reported observations made with MDI onboard SOHO and claim to have detected a regular structure of 100-m \\lq \\lq hills'', uniformly spaced over the surface of the Sun with a characteristic separation of 90,000 km. They suggest that this structure is the surface manifestation of Rossby waves, or r-modes oscillations. Finally, \\citet{Dimitropoulou08} have linked the found periodicities in different classes (B, C, M, X) of X-ray flares with the theoretical periods derived by \\citet{lou00}, pointing out that odd m periodicities are also frequent and significant. On the other hand, most of the proposed mechanisms to explain solar flares, specially the most energetic ones, accept as a prerequisite the emergence of magnetic flux \\citep{priest90, forbes91} which, by reconnection with the ambient field, triggers the destabilization of active regions. Based on this mechanism, Carbonell \\& Ballester (1990, 1992) suggested that the periodic increase in the occurrence rate of energetic flares is related to a periodic emergence of magnetic flux through the photosphere. Later, Oliver et al. (1998) showed that during solar cycle 21 there was a perfect time correlation between the intervals of occurrence of the periodicity in sunspot areas and energetic flares, and \\citet{ballester02} clearly pointed out that in cycle 21, and during the time interval in which the periodicity appeared, there was a perfect time and frequency coincidence between the impulses of high-energy flares and those corresponding to strong photospheric magnetic flux. The efficiency of the reconnection mechanism depends on the geometry of the two flux systems \\citep{Galsgaard07} and recent high resolution observations performed by \\citet{zuca08} have confirmed the suitability of the mentioned mechanism for flare production. Emerged magnetic flux is probably connected to deeper regions, namely to the tachocline, which is a thin, transition layer between differentially rotating convection zone and rigidly rotating radiative envelope. The tachocline may prevent the spreading of the solar angular momentum from the convection zone to the interior \\citep{spi92,gough98,gough07,gar07} and probably it is the place, where the large-scale magnetic field which governs the solar activity is generated/amplified. The observed periodicity of 155--160 days in the emerging flux is in the range of Rossby wave spectrum. Therefore, we suggest that the periodicity is connected to the Rossby wave activity in the tachocline. Rossby waves are well studied in the geophysical context \\citep{gill82,ped87}, however, the presence of magnetic fields significantly modifies their dynamics \\citep{zaqa07,zaqa09}. On the other hand, the differential rotation, which is inevitably present in the tachocline, may lead to the instability of particular harmonics of magnetic Rossby waves. It has been shown that the joint action of toroidal magnetic field and the differential rotation generally leads to tachocline instabilities \\citep{gilman97,cal03,dik05,gil07,gil007}. However, the stability analysis usually has been performed in an inertial frame, which complicates to extract the information about unstable Rossby modes. Therefore, it is of paramount importance to perform the stability analysis in a rotating frame. Another important point is that the consideration of a rotating frame may tighten the stability criteria as it has been suggested by \\citet{hug01}. The difference between the present analysis and that by \\citet{hug01} is the inclusion of rotation which allows us to obtain Rossby wave solutions. In this paper, we use a rotating spherical coordinate system to study the linear stability of magnetic Rossby waves in the solar tachocline taking into account the latitudinal differential rotation and the toroidal magnetic field. We perform a two dimensional analysis, which can be followed in the future by more sophisticated shallow water considerations \\citep{gil00}. We first derive the analytical conditions of instability similar to \\citet{dahlburg98} and \\citet{hug01}. Then, we perform a detailed stability analysis using Legendre polynomial expansions \\citep{longuet68} to obtain the spectrum of unstable harmonics of magnetic Rossby waves. ", "conclusions": "In summary, we have shown that the destabilization of magnetic Rossby waves in the solar tachocline is produced by the joint effect of the latitudinal differential rotation and the toroidal magnetic field. The frequencies and growth rates of unstable harmonics depend on the combination of the differential rotation parameters and the magnetic field strength. The possible increase of latitudinal differential rotation at the solar maximum may trigger the instability of symmetric harmonic with period of 155-160 days in the upper part of the tachocline. This instability has a direct correlation with magnetic flux emergence, therefore the periodicity also appears in solar activity indicators related with magnetic flux. Later on, and probably via reconnection, this periodic magnetic flux emergence triggers the observed periodicity in solar flares. The magnetic Rossby wave theory opens a new research area about the activity on the Sun and other stars, and magnetic Rossby waves can be of paramount importance for observed intermediate periodicities in solar and stellar activity \\citep{massi98, massi05}. {\\bf Acknowledgements} The authors acknowledge the financial support provided by MICINN and FEDER funds under grant AYA2006-07637. Also, the Conselleria d'Economia, Hisenda i Innovaci\\'o of the Government of the Balearic Islands is gratefully acknowledged for the funding provided under grant PCTIB2005GC3-03. T. V. Z. acknowledges financial support from the Austrian Fond zur F\\\"orderung der wissenschaftlichen Forschung (under project P21197-N16), the Georgian National Science Foundation (under grant GNSF/ST06/4-098) and the Universitat de les Illes Balears. Wavelet software was provided by C. Torrence and G. Compo \\footnote{The software is available at http://paos.colorado.edu/research/wavelets}. \\appendix" }, "0911/0911.4072_arXiv.txt": { "abstract": "{} {Our main aim is to study the influence of the initial conditions of a cloud in the intermediate/high-mass star formation process.} {We observed with the VLA, PdBI, and SMA the centimeter and millimeter continuum, \\nth\\,(1--0), and CO\\,(2--1) emission associated with a dusty cloud harboring a nascent cluster with intermediate-mass protostars.} {At centimeter wavelengths we found a strong source, tracing a \\uchii\\ region, at the eastern edge of the dusty cloud, with a shell-like structure, and with the near-infrared counterpart falling in the center of the shell. This is presumably the most massive source of the forming cluster. About $15''$ to the west of the \\uchii\\ region and well embedded in the dusty cloud, we detected a strong millimeter source, MM1, associated with centimeter and near-infrared emission. MM1 seems to be driving a prominent high-velocity CO bipolar outflow elongated in the northeast-southwest direction, and is embedded in a ridge of dense gas traced by \\nth, elongated roughly in the same direction as the outflow. We estimated that MM1 is an intermediate-mass source in the Class 0/I phase. About $15''$ to the south of MM1, and still more deeply embedded in the dusty cloud, we detected a compact millimeter source, MM2, with neither centimeter nor near-infrared emission, but with water maser emission. MM2 is associated with a clump of \\nth, whose kinematics reveal a clear velocity gradient and additionally we found signposts of infall motions. MM2, being deeply embedded within the dusty cloud, with an associated water maser but no hints of CO outflow emission, is an intriguing object, presumably of intermediate mass.} {The \\uchii\\ region is found at the border of a dusty cloud which is currently undergoing active star formation. Two intermediate-mass protostars in the dusty cloud seem to have formed after the \\uchii\\ region and have different properties related to the outflow phenomenon. Thus, a single cloud with similar dust emission and similar dense gas column densities seems to be forming objects with different properties, suggesting that the initial conditions in the cloud are not determining all the star formation process.} ", "introduction": "} It is well established that intermediate (2--8~\\mo) and high-mass ($\\gtrsim8$~\\mo) stars form in cluster environments (\\eg\\ Kurtz \\et\\ \\cite{kurtz2000}; Evans \\et\\ \\cite{evans2009}), and that at the very first stages of their formation, they are embedded within dense and cold gas and dust from their original natal cloud. However, it is not clear to what extent the initial conditions of the natal cloud are determining the star formation story of the cloud and the properties of the nascent stars of low, intermediate, and high mass, forming within it. Regarding the star formation story, there is increasing evidence that a single cloud can undergo different episodes of star formation, as suggested by studies of deeply embedded clusters observed at spatial scales similar to the cluster member separation ($\\sim5000$~AU). In these studies, the intermediate/high-mass young stellar objects (YSOs) in the forming cluster seem to be in different evolutionary stages (judging from the peak of their spectral energy distributions: Beuther \\et\\ \\cite{beuther2007}; Palau \\et\\ \\cite{palau2007a},b; Leurini \\et\\ \\cite{leurini2007}; Williams \\et\\ \\cite{williams2009}). Concerning the properties of the intermediate/high-mass YSOs forming within the cloud, a few studies toward massive star-forming regions have revealed that two objects formed in the same environment may have very different properties in the ejection of matter, with one YSO driving a highly collimated outflow nearby another YSO driving an almost spherical mass ejection (\\eg\\ Torrelles \\et\\ \\cite{torrelles2001}, \\cite{torrelles2003}; Zapata \\et\\ \\cite{zapata2008}), indicating that the initial conditions in the cloud may not be the only agent determining the formation and evolution of the intermediate/high-mass members in a forming cluster. While most of these cases have been found at very high spatial scales ($\\sim1$~AU), a broad observational base in other regions and at other spatial scales is required to properly understand the role of initial conditions of the cloud in the cluster formation process and ultimately in the formation of intermediate/high-mass stars. In this paper we show a high angular resolution study of a deeply embedded cluster whose intermediate-mass protostars show differences not only in their spectral energy distributions, but also in the ejection phenomena associated with the YSOs. The region, IRAS\\,00117+6412, was selected from the list of Molinari \\et\\ (\\cite{molinari1996}) in a search for deeply embedded clusters which are luminous ($>1000$~\\lo) and nearby (distance $<3$~kpc). The IRAS source has a bolometric luminosity of 1400~\\lo\\ and is located at a distance of 1.8~kpc (Molinari \\et\\ \\cite{molinari1996}). The millimeter single-dish image reported by S\\'anchez-Monge \\et\\ (\\cite{sanchezmonge2008}) shows strong emission with some substructure, tracing a dusty cloud, and the centimeter emission reveals an ultra-compact \\hii\\ (\\uchii) region, associated with the brightest 2MASS source of the field at the eastern border of the dusty cloud. The dusty cloud is associated with an embedded cluster reported by Kumar \\et\\ (\\cite{kumar2006}), two H$_2$O masers spots (Cesaroni \\et\\ \\cite{cesaroni1988}; Wouterloot \\et\\ \\cite{wouterloot1993}), and CO\\,(2--1) bipolar outflow emission (Zhang \\et\\ \\cite{zhang2005}; Kim \\& Kurtz \\cite{kimkurtz2006}). All this is suggestive of the dusty cloud harboring a \\uchii\\ region and a deeply embedded stellar cluster forming in its surroundings. We conducted high-sensitivity radio interferometric observations in order to study the different millimeter sources embedded in the dusty cloud. To properly characterize the protostars, we also studied the distribution of dense gas, outflow and ionized gas emission. In the present paper we show the first results obtained from \\nth\\ dense gas and CO outflow emission, focusing mainly on the intermediate/high-mass content of the protocluster. ", "conclusions": "In this paper we study with high angular resolution the centimeter, and millimeter continuum, and \\nth\\,(1--0), and CO\\,(2--1) emission of the intermediate-mass YSOs forming within a dusty cloud, with the goal of assessing the role of the initial conditions in the star formation process in clusters. Our conclusions can be summarized as follows: \\begin{enumerate} \\item A \\uchii\\ region is found at the eastern border of the dusty cloud, with a shell-like structure and a flat spectral index, $-0.03\\pm0.08$. The estimated age and mass of the underlying star is $\\sim1$~Myr and $\\sim6$~\\mo. \\item Deeply embedded within the dusty cloud, we have discovered a millimeter source, MM1, associated with a 2MASS infrared source, which is driving a CO\\,(2--1) powerful and collimated high-velocity outflow, oriented in the southwest-northeast direction. The mass derived from the millimeter continuum emission for MM1 is $\\sim3$~\\mo. MM1 is embedded within a ridge of dense gas as traced by \\nth, which seems to be rotating roughly along the outflow axis. MM1 is associated with centimeter emission, whose spectral index is compatible with an ionized wind, and at 1.2~mm splits up into different subcomponents when observed with an angular resolution of $\\lesssim1''$. From the derived outflow momentum rate, we estimated a luminosity for MM1 of 400--600~\\lo. Thus, MM1 seems to be a Class 0/I intermediate-mass YSO. \\item About $\\sim15''$ to the south of MM1, our observations revealed a dust compact condensation, MM2, lying in a dark infrared region, associated with water maser emission and a dense core traced by \\nth\\ emission. The mass from the dust emission is $\\sim1.7$~\\mo, and the \\nth\\ excitation temperature and column density are similar to the ones derived for MM1. The dense core in MM2 is rotating in the same sense as the ridge associated with MM1 and seems to be undergoing infall. We modeled the MM2 dense core as a disk-like structure with an inner radius of $\\sim1''$ and an outer radius of $\\sim5''$, with a rotation velocity in the outer radius of $\\sim0.6\\pm0.2$~\\kms, and an infall velocity at the same radius of $\\sim0.17\\pm0.02$~\\kms. The non-detection of CO at any velocity toward MM2 makes this object intriguing. \\item Although MM1 and MM2 formed within the same cloud and have similar dust and dense gas emission, their properties, specially concerning the ejection phenomenon, seem to be different and could be indicating that the initial conditions in a cloud forming a cluster are not the only agent determining the properties of the members of the cluster. \\end{enumerate}" }, "0911/0911.3652_arXiv.txt": { "abstract": "We investigate the nature and spatial variations of turbulence in the Small Magellanic Cloud (SMC) by applying several statistical methods on the neutral hydrogen (HI) column density image of the SMC and a database of isothermal numerical simulations. By using the 3rd and 4th statistical moments we derive the spatial distribution of the sonic Mach number (${\\cal M}_s$) across the SMC. We find that about 90\\% of the HI in the SMC is subsonic or transonic. However, edges of the SMC `bar' have ${\\cal M}_s\\sim4$ and may be tracing shearing or turbulent flows. Using numerical simulations we also investigate how the slope of the spatial power spectrum depends on both sonic and Alfv\\'en Mach numbers. This allows us to gauge the Alfv\\'en Mach number of the SMC and conclude that its gas pressure dominates over the magnetic pressure. The super-Alfv\\'enic nature of the HI gas in the SMC is also highlighted by the bispectrum, a three-point correlation function which characterizes the level of non-Gaussianity in wave modes. We find that the bispectrum of the SMC HI column density displays similar large-scale correlations as numerical simulations, however it has localized enhancements of correlations. In addition, we find a break in correlations at a scale of $\\sim160$ pc. This may be caused by numerous expanding shells of a similar size. ", "introduction": "\\label{intro} In the recent decade, many advances in both observations and computational models have provided new insights into the workings and evolution of the interstellar medium (ISM). The emerging picture is that interstellar turbulence plays the key role in ISM structure formation and evolution (McKee \\& Ostriker 2007). In the Galaxy and the Magellanic Clouds, the ISM is turbulent on scales ranging from less than a parsec to a few kiloparsecs (Crovisier \\& Dickey 1983; Stanimirovic et al. 1999; Deshpande et al. 2000; Dickey et al. 2001; Elmegreen et al. 2001; Elmegreen \\& Scalo 2004). Although the observational evidence for the importance of turbulence in the ISM is overwhelming, many questions remain open. For example, what are the dominant energy sources and physical processes that convert kinetic energy into turbulence (Burkert 2006)? At what scales and through which modes is turbulent energy dissipated (Heyer \\& Zweibel 2004)? How do the level and type of turbulence depend on properties of the interstellar gas (e.g. presence/absence of star formation, presence/absence of tidal effects, or the strength of magnetic field)? Since no complete theory of astrophysical turbulence exists, studying its effects on the multiphase ISM can be challenging and calls for a combination of numerical and observational efforts. Statistical studies have proved to be important in characterizing the magnetized turbulent ISM \\cite[]{Lazarian09}, however the interpretation of results is not always straight forward. Several statistical methods have been extensively used for both observational and synthetic data. These statistics include probability density functions (PDFs), wavelets, the principal component analysis, higher order moments, Velocity Coordinate Spectrum (VCS), and Velocity Channel Analysis (VCA), to name just a few (Gill \\& Henriksen 1990; Brunt \\& Heyer 2002; Kowal, Lazarian \\& Pogosyan 2000, 2004, 2006; Lazarian \\& Beresnyak 2007). Most of these statistical methods require large datasets with a large spatial or velocity dynamic range, and produce a single, mostly one-dimensional, measure. This results in the lack of spatial information about turbulent properties across a given interstellar cloud, or a galaxy, making a connection with underlying physical properties highly difficult. In this paper we explore a new method for obtaining spatial information about the level and nature of ISM turbulence on the neutral hydrogen (HI) observations of the Small Magellanic Cloud (SMC). The SMC, a dwarf irregular galaxy in the Local Group, has a highly gas rich ISM environment (see, Stanimirovic et al. 1999, henceforth known as SX99), and is an excellent candidate for ISM studies. Being nearby (60 kpc, \\cite{West91}), the SMC is distant enough for all its objects to be treated as having roughly the same distance, unlike the Milky Way where distance determination is relatively uncertain. The HI observations of the SMC, obtained using the Australia Telescope Compact Array (ATCA) and the Parkes telescope (SX99), have been used for several investigations, including the HI spatial power spectrum and the kinematic study of HI, which revealed the existence of many expanding shells of gas and three supergiant shells. The power law index of HI density and velocity distributions was derived in SX99 and \\cite{Stan01}, while the Genus statistic in \\cite{Chep08} revealed spatial variations of HI morphology. Because the SX99 SMC data set is well studied, it is a perfect candidate to investigate new statistical methods. We can acquire new information, but also test and confirm past results, as well as validate the promise of these statistical tools for further use in other observational studies. In this study, we investigate turbulent properties of the HI in the SMC by applying the higher order moments on the HI column density image. We then use a database of MHD simulations to bootstrap the spatial distribution of the sonic Mach number across the SMC. The crucial aspect of our approach is the confluence of observations and numerical simulations: only by combining the two we can retrieve the spatial variations of turbulent properties. This is the reason why we oscillate between observational and synthetic data in this paper. We also investigate whether and how interstellar shocks leave footprints on the HI gas by employing the bispectrum, a three point statistical measure, on the SMC HI column density image. Again, to interpret our results we apply the same statistics on the database of MHD simulated column density images. In particular, the paper is organized as follows. We start with \\S~\\ref{sec:background} by providing a brief summary of previous work regarding the statistical methods used in our study. In \\S~\\ref{sec:data} we describe the SMC HI column density map and the database of numerical simulations of compressible MHD turbulence used for the comparison with observations. In \\S~4 we introduce higher order moments and their dependence on the sonic and Alfvenic Mach numbers. We then apply higher order moments on the SMC HI observations to derive an image of the sonic Mach number across the SMC, in \\S~5. We compare our results with an observational estimate of the sonic Mach number of the cold neutral medium (CNM) in the SMC, based on a comparison of the spin and kinetic temperature of HI absorption profiles, in \\S~\\ref{ratio}. In \\S~\\ref{ps} we show how the power-law slope of the spatial power spectrum depends on the sonic Mach number and use this to gauge the Alfvenic Mach number of the HI gas seen in emission. In \\S~\\ref{sec:bispectrum} we present an analysis of the bispectrum of the SMC, as well as a brief discussion of the noise and windowing effects. In \\S~\\ref{sec:Discussion} we provide a discussion of our results, followed by our conclusions. ", "conclusions": "\\label{sec:Discussion} \\subsection{Turbulence properties of the HI gas in the SMC} Using 3rd and 4th statistical moments of the HI column density image and boothstreping turbulent information from a database of isothermal MHD simulations, we have mapped spatial variations of the sonic Mach number across the SMC. While most of the HI seen in emission in the SMC appears to be subsonic or transonic, several supersonic regions have emerged from our study. It is interesting that these regions do not correlate well with the most recent sites of star formation and seem to point out to large scale shearing or tidal flows. Commonly, it is believed that supernovae and superbubbles are the main drivers of galactic turbulence (McCray \\& Snow 1979), with a typical size of $\\sim100$ pc. While we do not have high enough resolution to see changes on such small scales ($\\sim10'$) in our derived map of ${\\cal M}_s$, most of the star-forming bar of the SMC appears to have subsonic or transonic properties when viewed at resolution of 30$'$. The most turbulent regions in the SMC may be tracing some kind of shearing flows between the SMC bar and the surrounding HI. This suggests that SMC's chaotic history with the LMC and our own Milky Way has probably left strong turbulent imprints on the HI gas. The lack of a turnover in the HI spatial power spectrum on the largest observed scales is also indicative of the fact that turbulent energy injection happens largely on scales larger than the size of the SMC \\cite[]{Stan01}. Similarly, Goldman (2000) suggested that the HI turbulence in the SMC was induced by large-scale flows from tidal interactions with the Milky Way and the LMC about $2\\times10^8$ yrs ago. Such large-scale bulk flows could have generated turbulence through shear instabilities, and this turbulence has not have had enough time to decay. Most of the HI in the SMC has a sonic Mach number of 1-2. This is on average at least two times smaller than what we inferred from HI absorption observations for the CNM in the SMC, ${\\cal M}_s\\sim$3.5-4. Similarly, for the CNM in the Milky Way Heiles \\& Troland (2003) found ${\\cal M}_s \\sim3$ with a large dispersion. A sonic Mach number of about 4-5 is commonly assumed for cold gas in molecular clouds (Federrath et al. 2009). For example, Heyer et al. (2006) measured from CO observations ${\\cal M}_s=4.2\\pm0.17$ for the Rosette molecular cloud, and ${\\cal M}_s=4.7\\pm0.12$ for G216-2.5. On the other hand, Hill et al. (2008) found that the distribution of the warm ionized medium (WIM) in the Milky Way can be best fit by models for mildly supersonic turbulence with ${\\cal M}_s\\sim1.4-2.4$. Turbulent properties of the HI in emission in the SMC are therefore closer to properties of the WIM in the Milky Way than properties of the CNM. This may suggest a large fraction of warm relative to cold HI being traced in HI emission. In the Milky Way, the HI is known to consist of at least two components with different temperature: the WNM with $T_{warm}=6000$~K and the CNM with $T_{cold}=70$~K. In addition, there could be a substantial amount of gas at intermediate temperatures \\cite[]{Heiles03}. Due to its lower metallicity, the HI in the SMC has different properties. Dickey et al. (2000) found $T_{cold}=40$~K, in agreement with theoretical expectations by Wolfire et al. (1995) whereby the existence of the two-phase medium is possible only at higher pressures compared with the range that applies for solar neighborhood conditions. They also estimated the fraction of cold HI in the SMC to be $\\la15$\\%. This is lower than $\\sim25$\\% found for the Milky Way. As our simulations are isothermal it is obvious to wonder how does the multi-phase HI affect our statistics and conclusions. We investigate this in Figure~\\ref{fig:ex} by producing a simulated data cube from a weighted combination of two cubes, one subsonic (${\\cal M}_{s}$=0.7) and one supersonic. The subsonic cube represents contribution from warmer gas, while the supersonic cube represents the cold gas. We combined the two cubes with different emphasis on warm vs cold gas, obtained the column density image of the resultant cube, and calculated its moments. Figure~\\ref{fig:ex} shows that for the case when the supersonic cube has ${\\cal M}_{s}$=2.0 skewness of the final cube with up to 50\\% of subsonic gas will be biased towards supersonic gas and will appear dominated by cold gas. If we increase the sonic Mach number of the supersonic cube to 4.0, the dominance of the higher turbulence is even more pronounced. A cube with up to 60-70\\% of subsonic gas and 25\\% of supersonic gas, will still have high skewness biased by the supersonic contribution. Considering that the HI column density image results in the mean ${\\cal M}_{s}=1-2$, significantly lower than what is expected for the cold HI, the fraction of the CNM along any LOS is most likely $\\la25$\\%. This supports the Dickey et al. (2000) estimate of the fraction of cold HI in the SMC being about 15\\%. \\begin{figure*}[tbh]\\centering \\includegraphics[scale=.7]{f16a.eps} \\caption{Skewness vs. percent of gas that is subsonic for column density of a $512^3$cube weighted with supersonic and subsonic gas. Supersonic gas generally dominates the skewness of the column density PDF.} \\label{fig:ex} \\end{figure*} \\subsection{Is the HI in the SMC sub-Alfv\\'enic or Super-Alfv\\'enic?} Two different statistical approaches in our study suggest that the HI gas in the SMC seen in emission is super-Alfv\\'enic. As we have shown in Figure ~\\ref{fig:power}, in addition to the sonic Mach number the spectral slope of the spatial power spectrum is sensitive to the Alf\\'venic Mach number for ${\\cal M}_{s}<2$. The sub-Alfv\\'enic models generally show steeper slopes due to large scale influence of magnetic fields. Thus, if one independently knows the sonic Mach number, it is possible to estimate the Alf\\'venic one using just the column density data. While the dependence of the spectral slope on the Alf\\'venic Mach number has not received much attention in the past, it is somewhat expected. Essentially, magnetization decreases compression in the shocks. Strong magnetic forces mix up density clumps preventing formation of isolated peaks, which results in a steeper spectrum. In addition, in the sub-Alfv\\'enic case we expect oblique shocks to be disrupted by Alfv\\'en shearing, which in turn, produces more small scale shocks \\cite[]{beresnyak05}. Another indication that the HI in the SMC is super-Alfv\\'enic comes from the bispectrum. The very sharp decrease in the bispectral amplitudes from large to small scales observed for the SMC is the closest to the trend found for simulated data for the case of ${\\cal M}_{s}\\sim2$ and ${\\cal M}_{A}\\sim2$. Detailed comparison between simulated and observed bispectra awaits future work, however this qualitative comparison is certainly encouraging. Assuming on average ${\\cal M}_{s}\\sim1-2$, the power spectrum slope suggests a super-Alfv\\'enic HI in the SMC with ${\\cal M}_{A}\\sim1-3$. This is generally in agreement with the observationally inferred strength of the magnetic field by Mao et al. (2008). Using their estimate for $B_{\\rm ext}=2$ $\\mu$G, a radius of the SMC of 2 kpc, the total hydrogen mass of $4.2\\times10^{9}$ M$_{\\odot}$, and a typical velocity dispersion of 20 \\kms (Stanimirovic et al. 2004), we estimate ${\\cal M}_{A}\\sim3$. As the Alf\\'venic Mach number shows the nature of the interplay between gas pressure and magnetic fields, it appears that the gas pressure in the SMC dominates over the magnetic pressure. \\subsection{Intermittency in mode correlations? } Our bispectrum analysis of the SMC HI data was the first attempt to apply bispectrum on observed astrophysical data. While more detailed comparison between observations and simulations awaits future work, we clearly see trends in the bispectral amplitudes similar to what was found for simulations of supersonic MHD turbulence. The most interesting finding is, however, the effect of small-scale variations in the $k_1=k_2$ correlations and a strong break in correlations at a scale of 160 pc. Such small-scale variations, or jumps, have not been seen in the bispectrum of simulated data. We can speculate about several possible scenarios that could explain their existence. The jumps could be caused by the energy injection due to processes other than turbulence affecting specific spatial scales. Alternatively, the jumps may be marking the presence of colder or multi-phase gas. Similarly, the observed break in the bispectrum at about 160 pc is intriguing. As we already pointed out, it is interesting that most expanding shells in the SMC (more than 500 were cataloged so far) have a diameter of $\\sim120$ pc. The break could be due to the lack of correlations on scales similar to the distance between two shell centers. Obviously this will require further studies. \\subsection{Limitations of the present study} A natural question to ask is how results presented in this paper depend on the resolution of numerical simulations. For example, Kritsuk et al. (2007) investigated how resolution of numerical simulations affects the power spectrum of density. These authors found that the spectral index estimates based on low resolution simulations bear large uncertainties due to the bottle neck contamination, and that the power spectra of $512^3$ simulations are substantially shallower then models with resolution of $1024^3$. However, while Kritsuk et al. (2007) only examined hydrodynamic turbulence, Beresnyak et al. (2008) showed that the slopes were very different between the MHD and pure hydrodynamic cases. For instance, the slopes for hydrodynamic simulations showed a pronounced and well defined bottleneck effect, while the MHD slopes were much less affected. This is indicative of MHD turbulence being less local than the Kolmogorov turbulence, and suggests that our simulations will be less affected by resolution. In addition, Kritsuk et al. (2007) found a difference in the slope between hydrodynamic $512^3$ and $1024^3$ simulations to be 0.17. This would result in a change of $d{\\cal M}_{s}\\sim0.5$ only and will not change our interpretation. We also add that in the case of higher statistical moments and the bispectrum BFKL confirmed trends noticed by KLB at lower resolution of 128$^3$. Another issue that should be further addressed and that could affect our results is the type of numerical forcing of turbulence. Federrath et al. (2009) recently investigated the effects of solenoidal vs. compressive (divergence-free vs. curl-free) forcing on a variety of statistics including PDFs and higher order moments. They found that both types of driving mechanisms are compatible with observations of molecular clouds however, depending on the data studied, one type could be superior then the other in terms of the statistics and reproduced observables. This implies that different regions in the SMC may exhibit statistical signatures of either compressive or solenoidally driven turbulence. \\subsection{Summary} \\label{sec:conclusions} We have investigated a new method for constraining turbulent properties of the ISM, specifically the sonic Mach number, by using the HI column density image and a database of numerical simulations with a range of sonic and Alf\\'venic Mach numbers. By applying the 3rd and 4th statistical moments on both observed and simulated data we have derived the spatial distribution of the sonic Mach number across the SMC with angular resolution of 30$'$. To provide an estimate of the Alf\\'venic Mach number we used two approaches: the spatial power spectrum and the bispectrum. Using the database of numerical simulations we have shown that the spatial power spectrum varies with both the sonic and Alf\\'venic Mach numbers. If the sonic number is known the Alf\\'venic number can be constrained from this dependence. The bispectrum shows the level of correlation between turbulent eddies of different size and depends greatly on the sonic Mach number, and somewhat on the Alf\\'venic Mach number. By comparing the bispectra of observations and simulations we have gauged the importance of magnetic fields relative to the gas pressure in the SMC. The following results were discussed in the paper. \\begin{itemize} \\item Skewness and kurtosis of the HI column density generally correlate well and are within the range expected from MHD simulations. This suggests that departures from Gaussianity could be interpreted as being governed by MHD turbulence. \\item Most of the HI in the SMC bar and the Eastern Wing is subsonic or transonic with ${\\cal M}_{s}\\sim0-2$. Sites of most recent star formation have ${\\cal M}_{s}\\sim1$. Regions with the highest skewness and kurtosis, which could be interpreted as having ${\\cal M}_{s}\\sim4$, correspond to the edges of the bar. The most turbulent regions are most likely tracing tidal or shearing flows. The fraction of the SMC with different turbulent properties is: 10\\% with ${\\cal M}_{s}>2$, 80\\% with $0<{\\cal M}_{s}<2$, and about 10\\% with very low values of ${\\cal M}_{s}$. \\item Using HI absorption profiles from Dickey et al. (2000) we have estimated that the CNM in the SMC is highly supersonic with ${\\cal M}_{s}=3.5-4$. This is at least a factor of two higher than what we measured from the higher statistical moments for the HI gas seen in emission. One possible reason for this discrepancy could be that HI emission is dominated by warm gas and the fraction of the CNM in the SMC is $\\la20$\\%. \\item The slope of the spatial power spectrum and the bispectrum suggest that the HI in the SMC is super-Alf\\'venic with ${\\cal M}_{A}\\sim1-3$. This is implies that the gas pressure dominates over the magnetic pressure. \\item The bispectrum of the HI column density shows large scale wave correlations suggesting a large scale energy injection mechanism. Contrary to simulations which show a smooth decrease of wave-wave correlations from large to small scales, the SMC bispectrum shows localized enhancements of correlations and at least one prominent break at $\\sim160$ pc. We speculate that the multi-phase medium, and/or energy injection by processes other than turbulence, could be responsible for the correlation jumps. The break on the other hand appears at a scale similar to the diameter of the majority of expanding shells in the SMC. \\end{itemize}" }, "0911/0911.4134_arXiv.txt": { "abstract": "In this paper we examine the contribution of galaxies with different infrared (IR) spectral energy distributions (SEDs) to the comoving infrared luminosity density, a proxy for the comoving star formation rate (SFR) density. We characterise galaxies as having either a {\\em cold} or {\\em hot} IR SED depending upon whether the rest-frame wavelength of their peak IR energy output is above or below $90\\,\\um$. Our work is based on a far-IR selected sample both in the local Universe and at high redshift, the former consisting of {\\it IRAS} $60\\,\\um$-selected galaxies at $z< 0.07$ and the latter of {\\it Spitzer} $70\\,\\um$ selected galaxies across $0.1< z\\le 1$. We find that the total IR luminosity densities for each redshift/luminosity bin agree well with results derived from other deep mid/far-IR surveys. At $z<0.07$ we observe the previously known results: that moderate luminosity galaxies ($L_{IR}<10^{11}\\,L_\\odot$) dominate the total luminosity density and that the fraction of cold galaxies decreases with increasing luminosity, becoming negligible at the highest luminosities. Conversely, above $z=0.1$ we find that luminous IR galaxies ($L_{IR}>10^{11}\\,L_\\odot$), the majority of which are cold, dominate the IR luminosity density. We therefore infer that cold galaxies dominate the IR luminosity density across the whole $0< z< 1$ range, hence appear to be the main driver behind the increase in SFR density up to $z\\sim1$ whereas local luminous galaxies are not, on the whole, representative of the high redshift population. ", "introduction": "The rise in the comoving star formation rate (SFR) density up to $z\\sim1$ and its subsequent flattening \\citep{Lilly:96, Madau:96} has now been well studied at several wavelengths \\citep[e.g.][and references therein]{Bunker:04, Hopkins:06}. While the global picture to $z=1$ has been well constrained by observation, the details of how or why this change occurs remain largely unknown. There now is evidence that star formation depends on both galaxy mass \\citep[][]{Feulner:05, Juneau:05} and environment \\citep[][]{Lewis:02,Elbaz:07} presenting a more complicated picture of galaxy evolution than simple evolution of the luminosity function. The infrared (IR), and particularly the far-IR, is one of the most powerful tracers of star formation as IR luminosity directly scales with SFR \\citep[][and references therein]{Kennicutt:98b} and far-IR emission originates from regions of cold dust and gas that constitute the fuel for an on-going burst of star formation. Another advantage of selecting sources at long wavelengths is the low contribution by active galactic nuclei (AGN), if present, to the total IR luminosity \\citep{Alexander:03, Clements:08}. Studies of the distant Universe at IR wavelengths remain limited and the most sensitive probe has been surveys with the {\\it Spitzer} $24\\,\\um$ band. However, this wavelength is relatively far from the peak of all but the hottest IR galaxies and progressively shifts to shorter wavelengths at higher redshifts where strong spectral features in the observed frame can also complicate matters. The few studies done with deep {\\it Spitzer} $70\\,\\um$ imaging have found rapid evolution in the total IR luminosity function \\citep{Huynh:07b, Magnelli:09}. When the IR luminosity function is integrated at different redshifts it is found that the IR luminosity density increases rapidly up to $z=1$ in a similar fashion to the SFR density. Nevertheless, the relative contribution by luminosity to the IR luminosity density changes with redshift with starbursts ($10\\le\\log(L_{IR}/L_\\odot)< 11$) dominating locally, but with luminous IR galaxies (LIRGs: $11\\le\\log(L_{IR}/L_\\odot)< 12$) and ultra-luminous IR galaxies (ULIRGs: $12\\le\\log(L_{IR}/L_\\odot)< 13$) becoming increasing contributors at higher redshifts \\citep{LeFloch:05,Magnelli:09}. In \\citet[][hereafter S09]{Symeonidis:09} we studied the IR spectral energy distributions (SEDs) of a sample of $70\\,\\um$ selected galaxies at $z\\ge0.1$, using a sample of local, $z<0.1$, {\\it IRAS} $60\\,\\um$ selected galaxies for comparison. We fitted the mid to far-IR photometry of both samples with models from \\citet{Siebenmorgen:07}. We found that the majority of the $70\\,\\um$ sources had IR SEDs which peaked, in $\\nu\\times F_\\nu$, at longer wavelengths than galaxies from the local sample with similar luminosities. In the local sample we observed a shift of the IR SED peak to shorter wavelengths with increasing luminosity \\citep[as has previously been noted:][]{Sanders:96, Chapman:03a, Rieke:09}, whereas the $70\\,\\um$ sample had a wide range of peak wavelengths the distribution of which varied little with luminosity. The observation that the IR SEDs of luminous, distant galaxies were on average different to their local analogues was in contrast to other recent results, \\citet[e.g.][]{Magnelli:09} who concluded there was no significant change in IR SED of luminous galaxies with redshift. Models of galaxy evolution in the IR have tended to assume that the SEDs of high redshift luminous sources follow the luminosity trend seen in local sources \\citep[][]{Lagache:05,Pearson:05,LeBorgne:09,RowanRobinson:09}. However, as shown in S09, the range of IR SEDs for luminous galaxies is much wider at high redshifts than seen locally. In this paper we examine the contribution of galaxies with different IR SEDs to the comoving IR luminosity density (IRLD). We will compare our results to earlier work \\citep{LeFloch:05} by examining the contribution to the IRLD by luminosity, but this time with a sample selected at $70\\,\\um$ rather than $24\\,\\um$. This selection enables us to use a more robust selection of IR luminous galaxies as $70\\,\\um$ lies closer to the peak of typical IR SEDs. Hence, the total IR luminosity can also be estimated more accurately, especially with constraints (mainly detections) from $160\\,\\um$ data. We examine the IRLD in more detail by focusing on the contribution by IR SED type within each bin. By obtaining an estimate of the peak of the IR SED in rest-frame $\\nu\\times F_\\nu$, we can characterise these IR bright sources as being either {\\em cold} or {\\em hot} depending on whether the SED peaks above or below $90\\,\\um$ ($32.2\\,$K for a blackbody). We choose this wavelength for two reasons. Firstly, local IR galaxies appear to have a warm IR component ($\\bar{\\lambda}\\sim 60\\,\\um$) associated with dust around young star forming regions and a cooler ``cirrus'' component ($\\lambda\\ge\\,100\\,\\um$) associated with more extended dust heated by the interstellar radiation field \\citep{Lonsdale:87}. Secondly, this $90\\,\\um$ cut also marks a divide between the SEDs of most local ULIRGs and those of cold SMGs seen at high redshift, most of which would be classified as cold by our definition \\citep{Chapman:05}. We note that there is a linear relationship between the {\\it IRAS} colours often used to define the IR galaxies as `cold' or `hot' \\citep[e.g.][]{Chapman:03a} and the peak wavelength used in this work. We present our far-IR galaxy sample in \\S2, describe our determination of the comoving IRLD in \\S3. We present our results in \\S4 and discuss them in \\S5. Throughout we use the current `concordance' cosmology: $\\Omega_M = 1 - \\Omega_{\\Lambda} = 0.3$, $\\Omega_0 = 1$, and $H_0 = 70\\, \\kmpspMpc$ \\citep{Spergel:03}. \\begin{figure} \\centerline{ \\psfig{file=fig1.ps,width=8.5cm,angle=270} } \\caption{The redshift/total ($8-1000\\,\\um$) IR luminosity distribution of the local ($z<0.07$) BGS sample and our {\\it Spitzer} $70\\,\\um$ sample ($z>0.1$) with sources characterised as cold or hot depending on whether their IR SED peaks above or below $90\\,\\um$ (rest-frame in $\\nu\\times F_\\nu$). The dotted lines indicate the binning chosen to investigate the comoving IRLD by IR luminosity and SED.} \\label{fig:lz} \\end{figure} ", "conclusions": "In S09 we demonstrated that the IR SEDs of individual LIRGs and ULIRGs at $0.1\\le z\\le 1.0$ peak at longer wavelengths, and span a wider range, than local galaxies at similar luminosities. Other changes in the properties of star forming galaxies have occured since $z\\sim1$. Star formation shifts from massive galaxies at high redshifts to progressively less massive galaxies at lower redshifts \\citep[e.g.][]{Cowie:96,Juneau:05}. A shift is also seen in environment of the most strongly star forming galaxies with star formation tending to occur in over-dense environments at $z\\sim 1$ \\citep{Elbaz:07}, a reverse of the trend seen locally. While the dependence of star formation on stellar mass and environment inform us about the star formation history and potential or current galaxy mergers, information about the IR SED is a direct measure of the physical conditions in galaxies underlying major episodes of star formation. In this paper we have expanded the results of S09 by deriving IR luminosity densities at different luminosities and epochs. We observe that the total IR luminosity densities derived from a $70\\,\\um$ sample are broadly consistent with results derived from shorter wavelengths (Fig~\\ref{fig:lumden}). The low IR luminosity density of the $z\\sim0.2$ starburst bin compared to other results is likely due to the effect of sample variance over the small cosmic volume probed. We also separate our sample into galaxies that can be characterised as having cold or hot SEDs and study their contribution to each redshift/luminosity bin. We find two striking new results. Firstly, we observed a rapid change in the make up of LIRG and ULIRG bins: locally they are dominated by hot galaxies, but at $z>0.2$ they are dominated by cold galaxies. Secondly, cold galaxies appear to be the dominant contributor to the total IR luminosity density over $0\\le z\\le 1.0$; when starburst galaxies dominate at $z<0.2$ we find that they are mainly cold and by the time LIRGs and ULIRGs begin to dominate at higher redshifts they are also mainly cold. Hence, a cold mode of star formation has dominated galaxy evolution since $z\\sim1$ and is likely responsible for the increase in the star formation rate density up to $z=1$. While Fig.~\\ref{fig:lumden} does dramatically show the rise of cold galaxies in the LIRG and ULIRG bins at higher redshifts we note that hot galaxies are still found at those redshifts. In fact the results presented here do not rule out a rise with redshift in the number (and luminosity density) of hot galaxies. Therefore, the distribution of types of SEDs remains quite wide at high redshift. Fig.~\\ref{fig:frac} most strikingly demonstrates the evolution of the relative contribution of cold galaxies to each redshift/luminosity bin. The rise in the contribution of cold galaxies to the LIRG and ULIRG bins is most rapid between $z\\sim0$ and $z\\sim0.3$. We now consider if AGN could be responsible for the change in the fraction of cold galaxies. As a powerful AGN would make the IR SED hotter the presence of an AGN can not explain the prevalence of cold galaxies at high redshift. AGN could, however, be the cause of the high fraction of hot galaxies locally. Studies have shown that the incidence of AGN are low in local LIRGs, $10-20\\%$, (Petric et al., 2009, in press), but while it is higher in ULRIGs, $\\sim40\\%$ \\citep{Farrah:07} the $8-1000\\,\\mu$m IR luminosity remains dominated by star formation. We note that at higher redshift the incidence of AGN, and their contribution to total IR luminosity, also remains low \\citep{Alexander:03, Clements:08}. Hence, AGN are not a significant contributor to the IR luminosity of the sources studied here and therefore are not responsible for any of the trends with redshift. The prominence of cold galaxies discovered here might suggest that there could be a substantial contribution to the IR emission of luminous galaxies at high redshift from a cold cirrus component akin to that observed in local IR galaxies. Such an interpretation is in agreement with the observation that many of the distant, cold, SMGs are extended on scales of $\\sim10\\,$kpc \\citep{Chapman:04} in contrast with the compact ULIRGs seen locally, and is also consistent with the change in the mean IR SED of luminous galaxies as inferred from the observed $70\\,\\um$ to radio flux density ratio of high redshift star forming galaxies \\citep{Seymour:09}. In addition, changes in dust properties (such as opacity, grain distribution etc.) with redshift are also a possible cause of the larger spread in SED types at high redshift. Most galaxy evolution models \\citep[e.g.][]{Lagache:05,Pearson:05,LeBorgne:09,RowanRobinson:09} assume that the IR SED is only dependent on IR luminosity, i.e. they typically use certain templates for star forming galaxies of a given luminosity, often using local galaxies like Arp 220 (which can be characterised as having a hot IR SED by our definition) for the most luminous population. This choice was often necessary as there were few constraints on any change with redshift until now. Models based on cooler SEDs could , for example, reconcile the flat or decreasing star formation rate density above $z=1$ and the far-IR and sub-millimetre source counts. In conclusion, we have found that the star formation history of the Universe is dominated by cold galaxies across $0$ 0.15 mag) at the level of the horizontal branch (HB) of NGC 147 produced significant errors in the period determinations - especially for the shorter period RR Lyraes (P $<$ 0.4 d). However, the effect of the photometric errors on the period determination was not addressed because the L-K method ignores the photometric errors in the process of identifying the optimal period. Saha et al's (1990) paper appears to be the only previous study of the RR Lyrae population in NGC 147. We are motivated to revisit the properties of these stars for two reasons. First, we have had good success in using a light curve template-fitting algorithm (Layden \\& Sarajedini 2000) to determine the properties of RR Lyraes such as periods, amplitudes, and mean magnitudes. This method includes the photometric errors in the analysis and has been streamlined and redesigned to be more user-friendly by Mancone \\& Sarajedini (2008). This will allow us to refine the determination of the RRL periods and place better constraints on the total number of such variables in NGC 147. Second, there are archival imaging data from the Hubble Space Telescope (HST) Wide Field Planetary Camera 2 (WFPC2) for NGC 147 that may allow us to update the list of RR Lyraes published by S90. Although the WFPC2 data provide accurate photometry at the level of the HB in NGC 147, they exhibit poor phase coverage with a time baseline of $\\sim$0.4 day. However, it is still useful because it could help to identify the shorter period, lower amplitude RR Lyrae stars that may not have been detected by S90. The next section describes the observational datasets that we will analyze. Sections 2 and 3 make it clear that neither the S90 data nor the WFPC2 data are ideal for the purpose of studying the RR Lyraes in NGC 147, but they do complement each other nicely. As discussed in Section 4, it is important to carry out simulations to fully understand the biases and caveats inherent in the results obtained from each of the datasets. The conclusions are presented in Section 5 as well as the case for future work to better characeterize the variable stars in NGC 147 and its M31 dwarf satellite cousin NGC 185. ", "conclusions": " 1. The $g$-band photometry from the work of Saha et al. (1990) likely possesses an adequate observational baseline and available epochs. However, our simulations showed that the photometric errors at the level of the horizontal branch significantly hinder the accurate determination of the pulsation periods of the RR Lyrae candidates in NGC 147. 2. Our template light curve fitting technique (FITLC) detected 36 probable RR Lyrae candidates from HST/WFPC2 archival data. However, our simulations reveal that the short observational baseline and small number of observations severely affect the accurate characterization of RR Lyrae periods longer than $\\sim$0.4 days, which are essentially the ab-type RR Lyraes. 3. The $g$-band photometry and the WFPC2 archival data analyzed herein present two extreme cases often found in period finding studies - good phase coverage but with large photometric errors, and high quality photometry with poor phase coverage. Our investigation of these two extreme cases not only provides a good reference for interpreting the pulsation properties of RR Lyrae variables in other similar situations, but also calls attention to a strong need for new high quality time-series observations of NGC 147. Thus, while we can confidently assert that NGC 147 contains RR Lyrae variables, and therefore a population older than $\\sim$10 Gyr, it is not possible at this time to use the pulsation properties of these RR Lyraes to study other aspects of this old population." }, "0911/0911.2712_arXiv.txt": { "abstract": "We present the first comprehensive study of short-timescale chromospheric H$\\alpha$ variability in M dwarfs using the individual 15 min spectroscopic exposures for $52,392$ objects from the Sloan Digital Sky Survey. Our sample contains about $10^3-10^4$ objects per spectral type bin in the range M0--M9, with a total of about $206,000$ spectra and a typical number of 3 exposures per object (ranging up to a maximum of $30$ exposures). Using this extensive data set we find that about $16\\%$ of the sources exhibit H$\\alpha$ emission in at least one exposure, and of those about $45\\%$ exhibit H$\\alpha$ emission in all of the available exposures. As in previous studies of H$\\alpha$ {\\it activity} ($L_{\\rm H\\alpha}/L_{\\rm bol}$) we find a rapid increase in the fraction of active objects from M0--M6. However, we find a subsequent decline in later spectral types that we attribute to our use of a spectral type dependent equivalent width threshold. Similarly, we find saturated activity at a level of $L_{\\rm H\\alpha}/L_{\\rm bol}\\approx 10^{-3.6}$ for spectral types M0--M5, followed by a decline to about $10^{-4.3}$ in the range M7--M9. Within the sample of objects with H$\\alpha$ emission, only 26\\% are consistent with non-variable emission, independent of spectral type. The H$\\alpha$ {\\it variability}, quantified in terms of the ratio of maximum to minimum H$\\alpha$ equivalent width ($R_{\\rm EW}$), and the ratio of the standard deviation to the mean ($\\sigma_{\\rm EW}/\\langle {\\rm EW}\\rangle$), exhibits a rapid rise from M0 to M5, followed by a plateau and a possible decline in M9 objects. In particular, $R_{\\rm EW}$ increases from a median value of about 1.8 for M0--M3 to about 2.5 for M7--M9, and variability with $R_{\\rm EW}\\gtrsim 10$ is only observed in objects later than M5. For the combined sample we find that the $R_{\\rm EW}$ values follow an exponential distribution with $N(R_{\\rm EW})\\propto {\\rm exp}[-(R_{\\rm EW}-1)/2]$; for M5--M9 objects the characteristic scale is $R_{\\rm EW}-1\\approx 2.7$, indicative of stronger variability. In addition, we find that objects with persistent H$\\alpha$ emission exhibit smaller values of $R_{\\rm EW}$ than those with intermittent H$\\alpha$ emission. Based on these results we conclude that H$\\alpha$ variability in M dwarfs on timescales of 15 min to 1 hr increases with later spectral type, and that the variability is larger for intermittent sources. Future studies using this large sample will address the variability timescales, the variability of other chromospheric emission lines (e.g., H$\\beta$, \\ion{Ca}{2} H\\&K), and the origin of the highest amplitude events. ", "introduction": "The study of magnetic activity in fully convective low mass stars and brown dwarfs (spectral types late-M, L, and T) has progressed in recent years from the question of whether these objects produce stable fields to a quantitative investigation of how the fields are produced and dissipated. Unlike the solar-type $\\alpha\\Omega$ dynamo, which operates in the transition region between the radiative and convective zones (the tachocline; \\citealt{par55}), objects later than spectral type M3 can only support a convective dynamo. Numerical simulations of such a mechanism are still at an early stage, but they suggest that large-scale axisymmetric fields can indeed be generated in fully convective objects, at least for conditions that roughly correspond to mid-M dwarfs ($M\\sim 0.3$ M$_\\odot$; \\citealt{bro08}). Thus, observational constraints on the scale, geometry, and dissipation of the fields are essential. Several observational techniques are now being used to address these questions, including Zeeman measurements in Stokes $I$ and $V$, which probe the strength of the integrated surface fields and their large-scale topology, respectively \\citep{rb07,dmp+08,mdp+08}, and activity indicators such as radio, X-ray, and H$\\alpha$ emission, which trace the dissipation of the field and hence its strength and geometry (e.g., \\citealt{whw+04,ber06,bbf+09}). The activity indicators also provide insight into magnetic heating, and their temporal variability can potentially trace the field properties on small scales that are inaccessible to the Zeeman measurements. The current results from Zeeman measurements point to a transition from mainly toroidal and non-axisymmetric fields in M0--M3 dwarfs to predominantly poloidal axisymmetric fields in mid-M dwarfs \\citep{dmp+08,mdp+08}, with field strength of $\\sim 0.1-3$ kG \\citep{rb07,dmp+08,mdp+08}. Studies of activity indicators point to a rapid decline in X-ray and H$\\alpha$ activity (i.e., $L_{X,{\\rm H\\alpha}}/L_{\\rm bol}$) at about spectral type M7, and uniform radio luminosity (i.e., increasing $L_{\\rm rad}/L_{\\rm bol}$) at least to spectral type L3 (\\citealt{bbf+09} and references therein). The disparate trends may be due to a decoupling of the magnetic fields from the increasingly neutral atmospheres, or to a shift in the magnetic field configuration. While X-ray and radio observations provide powerful insight into the nature of the magnetic fields, H$\\alpha$ chromospheric emission, which traces gas at temperatures of $\\sim 10^4$ K, is more easily accessible for large samples as a by-product of standard optical spectroscopic observations. In recent years, large spectroscopic samples of M dwarfs have become available through dedicated studies and large-scale surveys such as the Sloan Digital Sky Survey (SDSS). These extensive samples have led to several important results concerning chromospheric activity. First, the fraction of objects that exhibit H$\\alpha$ emission increases rapidly from $\\sim 5\\%$ in the K5--M3 dwarfs to a peak of $\\sim 80-100\\%$ around spectral type M7, followed by a subsequent decline to a few percent in the L dwarfs \\citep{gmr+00,whw+04,whb+08}. Second, while the level of activity increases with both rotation and youth in F--K stars, it reaches a saturated value of $L_{\\rm H\\alpha}/L_{\\rm bol}\\approx 10^{-3.6}$ in M0--M6 dwarfs, followed by a rapid decline to $L_{\\rm H\\alpha}/L_{\\rm bol}\\approx 10^{-5}$ by spectral type L0 \\citep{hgr96,gmr+00,whw+04}, and a breakdown of the rotation-activity relation \\citep{mb02}. Finally, a small fraction ($\\lesssim 5\\%$) of late-M and L dwarfs have been serendipitously observed to exhibit H$\\alpha$ flares that reach the saturated emission levels found in the early-M dwarfs \\citep{lkc+03}. To uniformly address the latter point --- H$\\alpha$ variability --- we recently carried out spectroscopic monitoring observations of about 40 M4--M8 dwarfs with known H$\\alpha$ emission \\citep{lbk09}. With observations of about 1 hr per source and a time resolution of $5-10$ min, we found that about $80\\%$ of these objects exhibit H$\\alpha$ variability on a wide range of timescales, ranging in amplitude from tens of percent up to a factor of about 5. Indeed, the timescale distribution for variability ``events'' is nearly flat from 10 min to 1 hr, with fewer events on timescales below 10 min. The variability amplitudes follow an exponential distribution with a characteristic scale of ${\\rm Max(EW)/Min(EW)}-1\\approx 0.7$. Finally, we found tentative evidence for increased variability with later spectral type. Here, we extend our study of H$\\alpha$ variability by three orders of magnitude using SDSS time-resolved spectroscopic data for a sample of 52,392 M0--M9 dwarfs (Knapp et al. 2009 in preparation; hereafter Paper I). For the first time we take advantage of the individual 15 min exposures, with a typical number of 3 exposures per source. Using these data we can thus probe H$\\alpha$ variability on timescale of $15$ min to $\\sim 1$ hr. In this paper we focus on the H$\\alpha$ variability amplitudes and their relation to the H$\\alpha$ activity level ($L_{\\rm H\\alpha}/L_{\\rm bol}$) and spectral type. We also re-investigate the relation between H$\\alpha$ activity and spectral type using the individual 15 min spectra, rather than the SDSS pipeline-combined spectra, which were used in previous work \\citep{whw+04,whb+08}. In \\S\\ref{sec:obs} we define our sample selection and describe the analysis method for measuring H$\\alpha$ equivalent widths. The results for H$\\alpha$ variability and activity are described in \\S\\ref{sec:results}. Future papers will focus on the variability timescales, the highest amplitude events, the variability of other chromospheric emission lines (e.g., higher-order Balmer lines, \\ion{Ca}{2} H\\&K, \\ion{He}{1} lines), and the relation between chromospheric variability and various source properties, such as age. ", "conclusions": "\\label{sec:conc} We study the variability of the H$\\alpha$ chromospheric emission line in M dwarfs using the individual 15 min spectroscopic exposures from SDSS for an unprecedentedly large sample of 52,392 stars. The typical timescale probed by these data is 15 min to 1 hr. About $16\\%$ of the sources exhibit H$\\alpha$ emission in at least one exposure, and of those about $45\\%$ have H$\\alpha$ emission in all of the available exposures. Only $26\\%$ of all the objects with H$\\alpha$ emission are consistent with steady emission, spread relatively uniformly from M0 to M9. The detection fraction as a function of spectral type exhibits the known trend of a sharp increase from a few percent in M0--M3 to tens of percent in later objects. However, unlike the results of previous studies based on SDSS pipeline-combined spectra and a fixed equivalent width threshold of 1 \\AA\\ \\citep{whw+04,whb+08}, we find a peak detection fraction of $\\sim 40\\%$, with a steady decline beyond M6. We attribute this trend to our use of a spectral type dependent detection threshold of $\\sim 3-6$ \\AA\\ ($3\\sigma$) in M5--M9 objects. The H$\\alpha$ variability, quantified in terms of $R_{\\rm EW}$, exhibits a substantial increase with later spectral type. For M dwarfs as a whole, the distribution of $R_{\\rm EW}$ values appears to follow an exponential with $N(R_{\\rm EW})\\propto{\\rm exp}[-(R_{\\rm EW}-1)/2]$; the characteristic scale is about 2.7 for M5--M9. We also find that objects with partial detections exhibit a wider distribution of $R_{\\rm EW}$ values than those with persistent H$\\alpha$ emission. These results indicate that H$\\alpha$ variability increases with later spectral type, even as H$\\alpha$ activity declines, and that M dwarfs with intermittent H$\\alpha$ activity are more variable than those with persistent H$\\alpha$ emission. In the context of chromospheric heating, these trends suggest that the magnetic energy input typically varies by a factor of about 2 on timescales of $\\sim 15$ min to $\\sim 1$ hr. Moreover, the increased variability as a function of later spectral type, with a particularly large increase beyond spectral types M3--M4, hints at an overall shift in the magnetic field, leading to less uniform field dissipation. Taken at face value, this conclusion suggests that the relative contribution from small scales, which are likely to dissipate on short timescales, increases for later M dwarfs. This possible trend is in conflict with preliminary results from Zeeman measurements of a few objects in the range M0--M5, which point to an increase in the relative contribution from large-scale fields \\citep{dmp+08,mdp+08,rb09}. Alternatively, the increase in H$\\alpha$ variability, particularly for sources with intermittent emission, may be due to an increase in the stochastic dissipation of the field in the presence of increasingly neutral atmospheres. In this scenario, heating of the bulk chromospheric plasma is suppressed by the decoupling of the field from the atmosphere (leading to a decline in $L_{\\rm H\\alpha}/L_{\\rm bol}$), but small-scale stochastic heating still takes place (leading to an increase in the H$\\alpha$ variability). The role of these two scenarios may become clearer as samples of late-M dwarfs with Zeeman measurements become available." }, "0911/0911.0521_arXiv.txt": { "abstract": "We study, within an effective field theory framework, $O(E^{2}/\\Mpl^{2})$ Planck-scale suppressed Lorentz invariance violation (LV) effects in the neutrino sector, whose size we parameterize by a dimensionless parameter $\\eta_{\\nu}$. We find deviations from predictions of Lorentz invariant physics in the cosmogenic neutrino spectrum. For positive $O(1)$ coefficients no neutrino will survive above $10^{19}~\\eV$. The existence of this cutoff generates a bump in the neutrino spectrum at energies of $10^{17}~\\eV$. Although at present no constraint can be cast, as current experiments do not have enough sensitivity to detect ultra-high-energy neutrinos, we show that experiments in construction or being planned have the potential to cast limits as strong as $\\eta_{\\nu} \\lesssim 10^{-4}$ on the neutrino LV parameter, depending on how LV is distributed among neutrino mass states. Constraints on $\\eta_{\\nu} < 0$ can in principle be obtained with this strategy, but they require a more detailed modeling of how LV affects the neutrino sector. ", "introduction": "Over the last fifteen years there has been consistent theoretical interest in possible small deviations from the exact local Lorentz Invariance (LI) of general relativity as well as a flourishing of observational tests. The theoretical interest is driven primarily by ideas in the Quantum Gravity (QG) community that Lorentz invariance may not be an exact local symmetry of the vacuum. The possibility of outright Lorentz symmetry violation (LV) or a different realization of the symmetry than in special relativity has arisen in string theory~\\cite{KS89, Ellis:2008gg}, Loop QG~\\cite{LoopQG,Rovelli:2002vp,Alfaro:2002ya}, non-commutative geometry~\\cite{Carroll:2001ws, Lukierski:1993wx, AmelinoCamelia:1999pm, Chaichian:2004za}, space-time foam~\\cite{AmelinoCamelia:1997gz}, some brane-world backgrounds~\\cite{Burgess:2002tb}, and condensed matter analogues of ``emergent gravity''~\\cite{Analogues}. As well, allowing the fundamental theory of gravity to be non-relativistic in the ultraviolet can make gravity renormalizable while avoiding some of the other pathologies that plague renormalizable gravitational actions with higher derivative terms~\\cite{Horava:2009uw}. Constructing useful tests for the various active models and ideas is therefore vital. Since there are so many theoretical models around, a good approach is to work within a calculable framework where all possible LV terms are parameterized. Each theory then picks out a certain combination of terms which can be constrained. In this vein, a standard method is to simply analyze a Lagrangian containing the standard model fields and all LV operators of interest that can be constructed by coupling the standard model fields to new LV tensor fields that have non-zero vacuum expectation values\\footnote{There are other approaches to either violate or modify Lorentz invariance, that do not necessarily yield a low energy EFT (see \\cite{AmelinoCamelia:2008qg} and refs therein). However, these models do not easily lend themselves to particle physics constraints as the dynamics of particles is less well understood and hence we do not consider them here. In particular, we remark here that ideas of deformation, rather than breaking, of the Lorentz symmetry (see, e.g., \\cite{AmelinoCamelia:2000mn}) do not have an ordinary-EFT formulation, hence they cannot be tested with the arguments presented in this work.}. All renormalizable LV operators that can be added to the standard model in this way are known as the Standard Model Extension (SME)~\\cite{Colladay:1998fq}. These operators all have mass dimension three or four and can be further classified by their behavior under CPT. Higher mass dimension operators can be systematically explored as well, which is useful in case the naive EFT hierarchy breaks down due to other new physics (e.g.~SUSY) or quantum gravity introducing a custodial mechanism for the renormalizable operators in the infrared. The CPT odd dimension five kinetic terms for QED coupled to a non-zero vector field were written down in~\\cite{Myers:2003fd} while the full set of dimension five operators with a vector were analyzed in~\\cite{Bolokhov:2007yc}. The dimension five and six CPT even kinetic terms for QED for particles coupled to a non-zero background vector, which we are primarily interested in here, were partially analyzed in~\\cite{Mattingly:2008pw}. The full set of dimension five and six operators for QED has recently been introduced in~\\cite{Kostelecky:2009zp}. It is notable that SUSY forbids renormalizable operators for matter coupled to non-zero vectors~\\cite{GrootNibbelink:2004za} but permits certain nonrenormalizable operators at mass dimension five and six. Many of the parameterized LV operators have been very tightly constrained via direct observations (see~\\cite{Mattingly:2005re,Jacobson:2005bg, AmelinoCamelia:2008qg, Liberati:2009pf} for reviews). In particular, the dimension five and six CPT even operators for LV with a vector field have been recently directly\\footnote{Notice that all these operators can be indirectly constrained by EFT arguments~\\cite{Collins:2004bp}, as higher dimension LV operators induce large renormalizable ones if we assume no other relevant physics enters between the TeV and $\\Mpl$ energies. SUSY, however, is an example of new relevant physics that can change this.} constrained in the hadronic sector by exploiting ultra-high energy cosmic-ray observations \\cite{Maccione:2009ju} performed by the Pierre Auger Observatory (PAO). Indeed, the construction and successful operation of this instrument has brought UHECRs to the interest of a wide community of scientists and it is expected to allow, in the near future, the assessment of several problems of UHECR physics and also to test fundamental physics (in particular Lorentz invariance in the QED sector) with unprecedented precision \\cite{Galaverni:2007tq,Maccione:2008iw,Galaverni:2008yj}. The UHECR constraints \\cite{Kifune:1999ex,Aloisio:2000cm,AmelinoCamelia:2000zs,Stecker:2004xm,GonzalezMestres:2009di,Scully:2008jp,Maccione:2009ju,Stecker:2009hj} rely on the behavior of particle reaction thresholds with LV, which are one of the best methods in the EFT approach to constrain nonrenormalizable LV operators. Many LV operators give modified dispersion relations for free particles, where the energy as a function of momentum deviates slightly from the special relativistic form. For threshold reactions what matters is not the size of the LV correction to the energy compared to the absolute energy of the particle, but instead the size of the LV correction to the mass of the particles in the reaction. Hence the LV terms usually become important when their size becomes comparable to the mass of the heaviest particle. If the LV term scales with energy as $E^n$, then this critical energy is $E_{cr} \\sim \\left(m^{2}\\Mpl^{n-2}\\right)^{1/n}$ \\cite{Jacobson:2002hd}. According to this reasoning, the larger the particle mass the higher is the energy at which threshold LV effects come into play. This is why $\\gtrsim \\TeV$ electrons and positrons, but not protons, can be used to constrain $n=3$ LV \\cite{Maccione:2007yc}, and why UHE protons are needed to obtain constraints on hadronic LV with $n=4$ scaling (which corresponds to CPT even mass dimension five and six operators). From this point of view, neutrinos, with their tiny mass of order $m_{\\nu} \\simeq 0.01~\\eV$ \\cite{pdg}, are in principle the most suited particles to provide strong constraints on LV, at least for reactions involving {\\em only} neutrinos. One such reaction is that of neutrino oscillation. Indeed, for a decade neutrino oscillations have proven to be excellent tests of the SME and other LV models~\\cite{Coleman:1998ti,Kostelecky:2009zp,GonzalezGarcia:2005xw,Diaz:2009qk,Yang:2009ge} as when the LV corrections are near the neutrino mass, the oscillation pattern can change as a function of energy, direction and mass. ICECUBE may even be able to probe dimension six operators with time of flight techniques with TeV neutrinos from distant Gamma Ray Bursts~\\cite{GonzalezGarcia:2006na}. UHECR experiments have also the capability of placing constraints on SME parameters by exploiting neutrino oscillations \\cite{Bhattacharya:2009tx}\\footnote{We notice that in the same work \\cite{Bhattacharya:2009tx} the process of neutrino decay is discussed, but in a different context than what we consider here.}. One can construct complementary and in some measures even more sensitive neutrino tests of higher dimension operators by leveraging observables which deviate more strongly from their special relativistic values as the neutrino travel distance increases. One such observable is found to be related to UHE neutrino spectrum observations. Despite the threshold being low for LV effects to kick in, neutrinos with ultra-high energy are necessary to achieve a signal, as they interact so weakly that the phase space for a LV reaction must be huge to generate an appreciable rate. This requirement implies that, since the LV terms and hence the phase space grow with energy, very large energies are needed. Indeed, for renormalizable operators, where the phase space does not grow quickly enough, reactions of the type we consider here never achieve the necessary rate given current bounds on their coefficients, even for UHE neutrinos over cosmological distances~\\cite{Coleman:1997xq,Coleman:1998ti}. However, for the CPT even non-renormalizable operators the situation is different, and we find that for energies of order $10^{17}~\\eV$ there can be a significant modification of the spectrum. Next generation neutrino detectors such as ANITA \\cite{Gorham:2008yk} and SuperEUSO \\cite{Petrolini:2009cg,Santangelo:2009et} are sensitive to neutrinos of energies $>10^{19}~\\eV$. Further experiments, like the planned ARIANNA \\cite{Barwick:2006tg,Barwick:2009zz} and IceRay \\cite{Allison:2009rz}, will cover the range $10^{17}\\div 10^{20}~\\eV$. In the present work we study how limits on the absolute scale of non-renormalizable CPT even LV neutrino parameters can be obtained from UHE neutrino observations. In particular we will determine the energy scale where the neutrino spectrum might begin to deviate as a function of the size of the LV coefficients. With three neutrino species as well as assorted light leptons possibly involved, the spectrum as a function of various LV parameters can only be computed by a detailed parameterized numerical search. At this stage, where LV in this sector is only speculative, we feel that such a search is unwarranted. However, we will show the ``best case'' scenario for a LV neutrino signal with dimension six operators as well as a scenario where a signal in the UHE spectrum is much more subtle even though LV is still relatively strong. This paper is structured as follows. In section~\\ref{sec:theo} we describe the theoretical LV framework in which we will derive the LV effects on the UHE neutrino spectrum. In sections~\\ref{sec:nuLV} and \\ref{sec:nusplitting} we will give general information on the standard understanding of UHE neutrino generation by UHECRs, and we will present and detail the main LV reactions possibly affecting their spectrum. Section~\\ref{sec:bestcase} is devoted to present results in our test cases. In Section \\ref{sec:otherprocesses} we discuss the possible role of other processes than neutrino splitting. In section~\\ref{sec:conclusions} we will report our final remarks and conclusions. ", "conclusions": "\\label{sec:conclusions} In this work we have investigated possible signals of higher dimension LV CPT even operators in the UHE neutrino spectrum, in particular the effect of the neutrino splitting on the UHE neutrino spectrum. This process provides a clean test as it does not involve other LV operators apart from the neutrinos' ones. In addition, since the dominant neutrino field is left-handed, there are no complications due to different LV for different chirality of fermion, as there are in the UHECR case, and one ends up with one LV coefficient for each neutrino mass eigenstate. In the flavor blind scenario, where every mass eigenstate undergoes splitting approximately at the same energy, there is both a precocious fall off of the neutrino flux at UHE as well as a significant excess in the UHE neutrino flux at energies as low as $10^{17.5}$ eV. Noticeably, this kind of energies are well within reach of current and future UHECR experiments and about order of magnitude below those so far used for LV tests with hadronic and electromagnetic UHECRs. According to our study, existing or planned UHE neutrino experiments have the potential to probe LV in the neutrino sector for coefficients $\\eta \\gtrsim 10^{-4}$. However, we have discovered a serious difficulty with deriving constraints in the absence of a positive LV signal event. The distribution in mass eigenstates is roughly equal for UHE neutrinos in realistic mixing scenarios. Although it might seem somewhat unnatural that LV has different effects on different mass states of the same particle field, if this is the case then it is possible that LV can exist/be strong for one mass eigenstate yet be almost invisible for UHE neutrino detectors. In particular, the observation of a neutrino flux up to some maximal energy does not imply a firm conclusion on LV, at least with present accuracy. However, let us note that if a neutrino is ever detected at energy $E_{\\nu}$, we can rule out a flavor blind $\\eta_{\\nu} \\gtrsim (E_{\\nu}/10^{18.8}~\\eV)^{-13/4}$ according to Eq.~(\\ref{eq:constraint_naive}). On the other hand, the bump feature can lead to constraints on $\\eta_{\\nu} > 0$, as this bump should be observable in any such scenario of LV in the neutrino sector. In this case, to obtain a $O(1)$ constraint on LV requires at least a 50\\% accuracy in the determination of the neutrino spectrum in the energy range $10^{17}\\div10^{19}~\\eV$, which can be achieved by future experiments. \\ack We thank J.~Kelley and B.~McElrath for useful discussions. This work was supported by the Deutsche Forschungsgemeinschaft through the collaborative research centre SFB 676 ``Particles, Strings and the Early Universe: The Structure of Matter and Space-Time''. LM acknowledges support from the State of Hamburg, through the Collaborative Research program ``Connecting Particles with the Cosmos'' within the framework of the LandesExzellenzInitiative (LEXI)." }, "0911/0911.0698_arXiv.txt": { "abstract": "Following an initial explosion that might be launched either by magnetic interactions or neutrinos, a rotating magnetar radiating according to the classic dipole formula could power a very luminous supernova. While some $^{56}$Ni might be produced in the initial explosion, the peak of the light curve in a Type I supernova would not be directly related to its mass. In fact, the peak luminosity would be most sensitive to the dipole field strength of the magnetar. The tail of the light curve could resemble radioactive decay for some time but, assuming complete trapping of the pulsar emission, would eventually be brighter. Depending on the initial explosion energy, both high and moderate velocities could accompany a very luminous light curve. ", "introduction": "\\lSect{intro} The role of rotation in powering the explosion of supernovae has long been debated \\citep[e.g.,][]{Hoy46,Leb70,Ost71,Aki03}. Most recent work has focused on the possibility that a rotating neutron star could, by way of a magnetic interaction, be the energy source for {\\sl exploding} a massive star. While that issue is far from resolved, and recent headway has been made in exploding these same stars using neutrino transport, it is certain that a large fraction of supernova explosions produce rotating neutron stars and that those neutron stars frequently have large magnetic fields. Magnetars are a class of neutron stars with field strengths 10$^{14}$ to 10$^{15}$ Gauss and more. They may constitute 10\\% of the neutron star birthrate \\citep{Kou98}. It is likely that fast rotation in the collapsing iron core is responsible for creating the large magnetic field \\citep{Dun92}, so it is reasonable to expect that the birth of rapidly rotating, highly magnetic neutron stars is commonplace. It is also generally assumed that these rapidly rotating neutron stars are magnetically braked by dipole emission early on, accounting for the slow periods observed in anomalous x-ray pulsars and soft gamma-ray repeaters \\citep{Dun92,Kou98}. Since the initial rotational energy of such stars at birth must have been large and 1000 years later is small, where did the difference go? Might it have been emitted in some observable form? As a rough approximation, assume that the neutron star radiates its rotational energy away at a rate given by the traditional dipole formula for pulsars. For a typical moment of inertia of 10$^{45}$ g cm$^2$, the rotational energy of a neutron star with period, P$_{\\rm ms}$, in milliseconds is \\begin{equation} \\begin{split} E &= \\frac{1}{2} I \\omega^2 \\\\ &\\approx 2 \\times 10^{52} {\\rm P}_{\\rm ms}^{-2} \\ {\\rm erg}. \\end{split} \\lEq{pulsener} \\end{equation} The approximate energy loss for dipole radiation is given by the Larmor formula \\citep[e.g.,][]{Lan80}, \\begin{equation} \\begin{split} \\frac{d E}{d t} &= \\frac{2}{3 c^3} \\left(B R^3 \\ {\\rm Sin} \\, \\alpha \\right)^2 \\left(\\frac{2 \\pi}{{\\rm P}}\\right)^4 \\\\ &\\approx 10^{49} B_{15}^2 {\\rm P}_{\\rm ms}^{-4} \\ {\\rm erg \\ s^{-1}}. \\end{split} \\lEq{dipole} \\end{equation} Here $B_{15}$ is the surface dipole field in 10$^{15}$ Gauss, $R \\approx 10^6$ cm is the neutron star radius, and $\\alpha$ is the inclination angle between the magnetic and rotational axes, taken arbitrarily to be 30 degrees. At face value, these simple formulae suggest large luminosities for magnetars during the first few weeks of their lives. For example, the magnetar in SGR 1627-41 is estimated to have a current spin-down luminosity of $\\sim 4 \\times 10^{34}$ erg s$^{-1}$ at an age of 2.2 ky implying a magnetic field $B \\, {\\rm Sin} \\, \\alpha \\ \\sim 2 \\times 10^{14}$ Gauss \\citep{Esp09}. Extrapolated back to when the magnetar was 10 days old, \\eq{dipole} implies a luminosity of $2 \\times 10^{44}$ erg s$^{-1}$. Provided the initial rotation rate was rapid compared with its value at 10 days, this result is independent of the initial rotation rate. Here we explore the observational characteristics of Type Ib/c supernovae in which pulsars with magnetar-like characteristics have been embedded. The brightest supernovae actually come from neutron stars with fields that, for a magnetar, are relatively modest, $\\sim 10^{14}$ Gauss. Larger fields imply that most of the magnetar's rotational energy is dissipated early on, adding to the kinetic energy and mixing, but not appreciably affecting the luminosity. These luminous magnetar-powered supernovae might easily be confused with other forms of ``hypernovae'' where the luminosity has a radioactive origin. Alternatively, the lack of such emission constrains the existence of any rapidly rotating magnetar. This could be an important constraint in the context of gamma-ray bursts where the the possibility of a magnetar power source is currently debated. ", "conclusions": "Pulsars with magnetar-like magnetic fields can contribute appreciably to the light curves of Type Ib and Ic supernovae. In fact, it would be surprising if they never did. For moderate field strengths, the light curve is very luminous and long lasting and might be confused with those of pair-instability supernovae or circumstellar interaction. Indeed, \\citet{Mae07} have suggested a pulsar as the possible source powering the second (principal) maximum of the light curve of SN 2005bf. If magnetars are the central engine that powers the long-soft class of gamma-ray bursts \\citep[e.g.,][]{Woo06}, then it is reasonable to expect that they may contribute to the light curves of the supernovae that accompany them. On the one hand, this might facilitate the magnetar paradigm because the production of the necessary large amount of $^{56}$Ni has proven problematic \\citep{Buc09}. Having an alternate explanation that involves magnetar energy input would solve this problem. On the other hand, if a magnetar contribution to the light curve can be ruled out based on the spectrum and late time light curve, one must wonder how the magnetar is so effectively concealed. Is the rotational energy extracted with such high efficiency in the first few seconds that the magnetar forever afterwards rotates slowly, or do rotating magnetars not emit according to the popular dipole formula during the first year? Perhaps they do not. The absence of a pulsar contribution to the luminosity of typical supernovae is easy to understand. A pulsar born with a 14 ms period (10$^{50}$ erg), would only have a pulsar luminosity of $6 \\times 10^{39}$ erg s during the time when it is bright. This is small compared with the 10$^{42}$ - 10$^{43}$ erg s$^{-1}$ resulting from shock energy released by recombination (Type II supernova) or radioactivity (Type I supernova). But the extremely faint optical luminosity of SN 1987A, $< 8 \\times 10^{33}$ erg s$^{-1}$ seventeen years after its birth \\citep{Gra05} is very difficult to reconcile with any model with a young active pulsar. While dust extinction could be considerable, the lack of a point source in SN 1987A remains a mystery." }, "0911/0911.2181_arXiv.txt": { "abstract": "{ Neutron-induced reaction rates, including fission and neutron capture, are calculated in the temperature range $10^{8} \\leq T({\\rm K})\\leq 10^{10}$ within the framework of the statistical model for targets with the atomic number $84\\leq Z\\leq 118$ (from Po to Uuo) from the neutron to the proton drip-line. Four sets of rates have been calculated, utilizing - where possible - consistent nuclear data for neutron separation energies and fission barriers from Thomas-Fermi (TF), Extended Thomas-Fermi plus Strutinsky Integral (ETFSI), Finite-Range Droplet Model (FRDM) and Hartree-Fock-Bogolyubov (HFB) predictions. Tables of calculated values as well as analytic seven parameter fits in the standard REACLIB format are supplied$^0$. We also discuss the sensitivity of the rates to the input, aiming at a better understanding of the variations introduced by the nuclear input. } ", "introduction": "\\label{sec:intro} Investigations of nucleosynthesis processes make use of reaction networks including thousands of nuclei and tens of thousands of reactions. Most of these reactions occur far from stability and thus cannot yet be directly studied in the laboratory. In addition most of the nuclear properties including reaction rates, which are also required for the calculation of cross sections and astrophysical reaction rates, are not experimentally known either. Therefore, predictions based on theoretical models are necessary. While close to stability partial experimental information is available, relying fully on theoretical information leads to relatively large variations in computed cross sections far from stability. This is especially true for the region of fissionable nuclei, which is the focus of the present investigation. In the past, a series of efforts were applied to calculate neutron-capture rates for r-process nucleosynthesis and other astrophysical applications \\citep[e.g.,][and references therein]{A72,HWF76,WFH78,sar82,thi87,cowan91,rafkt00,most05,gori08}. Fission has often been neglected in astrophysical calculations. In early applications to astrophysical nucleosynthesis, usually only one mode was considered, beta-delayed fission \\citep{TMK83} or a phenomenological model of spontaneous fission \\citep{gori99,frei99,cowan99}. However, it was shown recently that neutron-induced fission is more important than beta-delayed fission in r-process nucleosynthesis \\citep{pan2002,panfkt04,fiss07}. Thus, the need to provide a compilation of neutron-induced fission rates is obvious. Initial investigations have been undertaken by \\cite{pan05} and \\cite{talys09}. Here we present extended calculations of neutron-induced fission rates for different predictions of masses and fission barriers. The present work also completes existing nuclear neutron-capture rate sets by extending the works of \\cite{rafkt00} and \\cite{pan05} to the region $84\\leq Z\\leq 118$ in order to provide the necessary input for nucleosynthesis studies under high neutron densities. As in \\citet{pan05}, the statistical model approach of Wolfenstein-Hauser-Feshbach \\citep{Wol51,HF52} for compound nuclear reactions was used, but employing more recent and complete data and predictions for masses, spins, and fission barriers. \\footnotetext{Tables 3-18 with these data are only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/} Nuclear mass and fission barrier predictions have a strong model dependence, and none of the existing models can reproduce all experimentally known data. Moreover, the fission process itself is complicated, and extended calculations for neutron-induced fission across the nuclear chart have to be done carefully. Here, we aim to provide rates for studying the endpoint of the r-process and the possible production of super-heavy elements. By comparing rates obtained with different choices of mass and fission barrier predictions we attempt to give a measure of the involved variations. Astrophysical models, providing the nucleosynthesis conditions, bear large variations in themselves. This is especially true for the r-process, for which the astrophysical site is still unknown despite decades of study. For a realistic and exhaustive exploration of synthesis conditions, simulations do not only have to vary astrophysical parameters, but also have to include a variation range of involved reaction rates given by different mass and fission barrier models. Our paper is structured as follows. In Sect. \\ref{sec:statmod} we briefly describe the statistical model used in the calculations as well as the nuclear input data and give a comparison of cross sections or rates for a number of experimentally known nuclei with existing experimental information and other theoretical models. These methods are then applied to supplement the rate sets of \\citet{rafkt00} of $(n,\\gamma)$-rates for chemical elements with $Z>83$ and predict neutron-induced fission cross sections and rates (where available in comparison to experiments). Section \\ref{sec:rates} presents these results and shows the sensitivity with respect to mass models and fission barriers employed. Rate fits for utilization in astrophysical calculations are discussed in Sect. \\ref{sec:rates_2}. In Sect. \\ref{sec:yields} we give a brief discussion and some examples of the mass distribution of fission fragments, which will be provided in an extended way in a forthcoming paper. The final Sect. \\ref{sec:conclusion} contains conclusions and a summary. The explanation of the tables and their structure are given in Appendix {\\bf A}. The complete tables of reaction rates and their fits are found at CDS in electronic form. \\begin{figure*} \\begin{center} \\hspace{3.5mm}\\includegraphics*[width=.32\\textwidth]{11967fg01a.eps} \\hspace{0mm}\\includegraphics*[width=.32\\textwidth]{11967fg01b.eps} \\hspace{0mm} \\includegraphics*[width=.325\\textwidth]{11967fg01c.eps} \\hspace*{2mm}\\includegraphics*[width=.335\\textwidth]{11967fg01d.eps} \\hspace{2mm}\\includegraphics*[width=.305\\textwidth]{11967fg01e.eps} \\hspace{3mm} \\includegraphics*[width=.305\\textwidth]{11967fg01f.eps} \\hspace*{0mm} \\includegraphics*[width=.345\\textwidth]{11967fg01g.eps} \\hspace{1mm} \\includegraphics*[width=.305\\textwidth]{11967fg01h.eps} \\includegraphics*[width=.325\\textwidth]{11967fg01i.eps} \\end{center} \\caption{ Present predictions of energy-dependent $(n,f)$ cross sections $\\sigma_{nf}(E)$ for some target nuclei of U, Np and Pu calculated in the framework of different mass and fission barrier predictions (ETFSI, TF, HFB-14) and experimental data, marked $B_{\\rm f}^{\\rm exp}$ as well. Experimentally measured cross-sections were used after JENDL-3.3 \\citep{jendl05}, averaged by the code JANIS \\citet{janis07}, displayed by a black line. All the predictions are given for a ground-state population. Our previous results \\citep{pan05} are shown as well. } \\label{ex_jeja} \\end{figure*} ", "conclusions": "\\label{sec:conclusion} We provide predictions of neutron-induced fission rates and $(n,\\gamma)$-rates for a wide range of astrophysical temperatures ($10^{8} \\leq T({\\rm K})\\leq 10^{10}$) and targets (proton- to neutron- drip-line for $84 \\leq Z \\leq 118$, i.e. from Po to Uuo ) in the framework of the Wolfenstein-Hauser-Feshbach model, making use of a variety of different mass and fission barrier predictions \\citep{mysw99,mamdo01,etfsi,frdm,homo80,talys09}. The astrophysical (stellar) reaction rates were fitted as in previous works \\citep{thi87,rafkt00} in the common REACLIB seven parameter form, and these parameters are also tabulated. This provides the basis for r-process nucleosynthesis calculations where the abundance predictions for the highest mass numbers as well as the effect of fission cycling are strongly dependent on the interplay of neutron capture and fission. In order to give an impression of the reliability of the results, we compared them with experiment and with available independent predictions before exploring the currently unreachable regions of the nuclear chart with a variety of theoretical predictions for nuclear masses and fission barriers (FRDM, ETFSI, TF, HFB). An extended comparison of neutron-induced fission rates with experiment and with available independent predictions was done. The dependence of rates on nuclear input data, most of all fission barriers, is high. Astrophysical nucleosynthesis yield predictions, especially in the transuranium region, should take into account these large differences in order to explore the variations involved. For this reason extended tables for neutron-induced fission rates as well neutron capture rates are presented for different mass and fission barrier predictions in fitted form for nucleosynthesis calculations. Their structure is given in the Appendix {\\bf A} (note that the full rate and fit tables are available at the CDS). Given that fission predictions far from stability have not been tested yet, and even close to stability none of the existing models has yet been proven to be superior (see Fig. \\ref{U_Pu_sigv_nf}). Nucleosynthesis calculations should probably continue to use a variety of these models. A further requirement for nucleosynthesis modeling in the region of fissioning nuclei is the knowledge of the mass distribution of fission products. This work is in progress (see Sect. \\ref{sec:yields})." }, "0911/0911.0968_arXiv.txt": { "abstract": "{} { Our primary goal is to search for planets around intermediate mass stars. We are also interested in studying the nature of radial velocity (RV) variations of K giant stars. } { We selected about 55 early K giant (K0 - K4) stars brighter than fifth magnitude that were observed using BOES, a high resolution spectrograph attached to the 1.8 m telescope at BOAO (Bohyunsan Optical Astronomy Observatory). BOES is equipped with $I_2$ absorption cell for high precision RV measurements. } { We detected a periodic radial velocity variations in the K0 III star \\gam1leo with a period of P = 429 days. An orbital fit of the observed RVs yields a period of P = 429 days, a semi-amplitude of K = 208 \\mps, and an eccentricity of $e = 0.14.$ To investigate the nature of the RV variations, we analyzed the photometric, Ca II $\\lambda$ 8662 equivalent width, and line-bisector variations of \\gam1leo. We conclude that the detected RV variations can be best explained by a planetary companion with an estimated mass of m $\\sin i = 8.78 M_{Jupiter}$ and a semi-major axis of $a = 1.19$ AU, assuming a stellar mass of 1.23 \\Msun. } {} ", "introduction": "Since the first exoplanet around a main sequence star was discovered in 1995, more than 300 exoplanets have been detected (http://exoplanet.eu/catalog.php). Among them, the majority of exoplanets candidates has been detected by using radial velocity methods around late F-G and K main-sequence stars. The existence of planets around intermediate mass early-type stars and their planetary parameters have not been investigated well due to their fast stellar rotational velocities and the lack of an appropriate number of sharp absorption lines in the stellar spectra. When intermediate mass stars evolve toward the red giant stage, they go through the G and K-giant phase where many sharp absorption lines appropriate for high precision RV measurements are available. Therefore, G and K giant stars are suitable targets for detecting exoplanets with the RV technique. There are several ongoing exoplanet survey projects around giant stars (Setiawan et al. 2005, Hatzes et al. 2005, Sato et al. 2007, Johnson et al. 2007, Lovis \\& Mayor 2007, Niedzielski et al. 2007, and Liu et al. 2008). Now, more than 20 exoplanets have been detected around giant stars and from this sample some statistical studies on the planetary systems around giants have been made (Pasquini et al. 2007, Hekker et al. 2008, and Sato et al. 2008). We started our precise RV survey using the 1.8m telescope at BOAO (Bohyunsan Optical Astronomy Observatory) in 2003 in order to search for exoplanets and to study the pulsations and stellar surface activity of K-giant stars. Our sample consists of 55 K0 - K4 giant stars. Most of them are brighter than fifth magnitude. Results from the first 3 years of our survey show that the majority (more than 90\\%) of our sample of K-giants show RV dariations. Among them, we found spectroscopic binaries, pulsating variables (see Kim et al. 2006), and several periodic RV variable stars. Our survey confirmed an exoplanet around $\\beta$ Gem (Han et al. 2008) discovered by Hatzes et al. (2006) and Reffert et al. (2006). In this paper we present a new exoplanet detection around the giant star \\gam1leo. ", "conclusions": "There is no doubt that a 429-d periodic variation is present in the RV measurements of {\\gam1leo}. This variation has persisted for almost 5 cycles with no change in phase or amplitude. To clarify the cause of the periodic RV variation, we investigated the possibility of any correlation between the RV variation and other stellar variations such as Hipparcos photometry, EW of Ca II $\\lambda$ 8662 line, line bisectors, and Ca II H line profile. Within the accuracy of our measurements, we found no convincing evidence of such correlation. Though we leave the real cause of the detected 429-d periodic RV variation open, we conclude that the observed periodic RV variation is best explained by an unseen companion with an estimated mass of m $\\sin i = 8.78 \\pm 0.2 M_{Jupiter}$. The non-negligible eccentricity of 0.14 found from the orbital solution also lends some support to a Keplerian orbital motion as the cause of the periodic RV variations. But we would like to emphasize that continued RV observations of \\gam1leo are essential to clarify the true nature of the RV variation. Perhaps the best way to confirm the orbital hypothesis is to detect the astrometric perturbation. For \\gam1leo the expected astrometric perturbation is several tenths of milli arcseconds, which is about the limit of current ground-based astrometric techniques. The nature of the 1340-d period found in the residual RVs remains unclear. We do not see any evidence for this period in either Hipparcos photometry, Ca II EW, or line bisectors. At face value this would argue for an additional companion, but this is not certain. Continued RV measurements are needed to confirm a possible second companion. We also found a significant power at P = 8.5 days in the residual RVs. The 8.5-d period is close to the estimated fundamental radial pulsation period. To confirm the reality of the 8.5-d period, more observations with better sampling are needed. It is known that planet harboring giants does not show metal-rich tendency found from dwarf stars. The low metalliicity of \\gam1leo strengthens this trend further. \\gam1leo adds one more exoplanet discovered around a binary star system. Since there are not still many exoplanet found around binary system, we consider our discovery as a valuable addition to the field." }, "0911/0911.0473_arXiv.txt": { "abstract": "We investigate the recovery chances of highly spinning waveforms immersed in LIGO S5-like noise by performing a matched filtering with $10^{6}$ randomly chosen spinning waveforms generated with the LAL package. While the masses of the compact binary are reasonably well recovered (slightly overestimated), the same does not hold true for the spins. We show the best fit matches both in the time-domain and the frequency-domain. These encompass some of the spinning characteristics of the signal, but far less than what would be required to identify the astrophysical parameters of the system. An improvement of the matching method is necessary, though may be difficult due to the noisy signal. ", "introduction": " ", "conclusions": "" }, "0911/0911.5076.txt": { "abstract": "% { The Bulge is the least understood major stellar population of the Milky Way. Most of what we know about the formation and evolution of the Bulge comes from bright giant stars. The underlying assumption that giants represent all the stars, and accurately trace the chemical evolution of a stellar population, is under debate. In particular, recent observations of a few microlensed dwarf stars give a very different picture of the evolution of the Bulge from that given by the giant stars. } { We aim to resolve the apparent discrepancy between Bulge metallicity distributions derived from microlensed dwarf stars and giant stars. Additionally, we aim to put observational constraints on the elemental abundance trends and chemical evolution of the Bulge. } { We perform a detailed elemental abundance analysis of dwarf stars in the Galactic bulge, based on high-resolution spectra that were obtained while the stars were optically magnified during gravitational microlensing events. The analysis method is the same as for a large sample of F and G dwarf stars in the Solar neighbourhood, enabling a fully differential comparison between the Bulge and the local stellar populations in the Galactic disc. } { We present detailed elemental abundances and stellar ages for six new dwarf stars in the Galactic bulge. Combining these with previous events, here re-analysed with the same methods, we study a homogeneous sample of 15 stars, which constitute the largest sample to date of microlensed dwarf stars in the Galactic bulge. We find that the stars span the full range of metallicities from $\\rm [Fe/H]=-0.72$ to $+0.54$, and an average metallicity of $\\rm\\langle [Fe/H]\\rangle=-0.08\\pm 0.47$, close to the average metallicity based on giant stars in the Bulge. Furthermore, the stars follow well-defined abundance trends, that for $\\rm [Fe/H]<0$ are very similar to those of the local Galactic thick disc. This suggests that the Bulge and the thick disc have had, at least partially, comparable chemical histories. At sub-solar metallicities we find the Bulge dwarf stars to have consistently old ages, while at super-solar metallicities we find a wide range of ages. Using the new age and abundance results from the microlensed dwarf stars we investigate possible formation scenarios for the Bulge. } {} ", "introduction": " ", "conclusions": "" }, "0911/0911.0190_arXiv.txt": { "abstract": "{A hypothetical time-variation of the gravitational constant $G$ would make neutron stars expand or contract, so the matter in their interiors would depart from beta equilibrium. This induces non-equilibrium weak reactions, which release energy that is invested partly in neutrino emission and partly in internal heating. Eventually, the star arrives at a stationary state in which the temperature remains nearly constant, as the forcing through the change of $G$ is balanced by the ongoing reactions. Using the surface temperature of the nearest millisecond pulsar (PSR J0437$-$4715) inferred from ultraviolet observations and results from theoretical modelling of the thermal evolution, we estimate two upper limits for this variation: (1) $|\\dot G/G| < 2 \\times 10^{-10}~\\mathrm{yr}^{-1},$ if the fast, ``direct Urca'' reactions are allowed, and (2) $|\\dot G/G|<4\\times 10^{-12}~\\mathrm{yr}^{-1},$ considering only the slower, ``modified Urca'' reactions. The latter is among the most restrictive upper limits obtained by other methods. ", "introduction": "A number of theorists have proposed that the so-called ``fundamental constants\" of Nature may vary with cosmological time, ever since Dirac suggested that the gravitational force may be weakening \\citep{dirac37}. Several experiments aimed at constraining the time variation of $G$ have been conducted, and their results can be grouped by the time scales they probe (see \\citealt{reisenegger07} for a list with references): (1) \\emph{Human lifetime} ($\\sim 10$~yr) experiments rely on real-time monitoring of distances within the Solar System, white-dwarf oscillation periods, and pulse arrival times of isolated and binary pulsars, (2) \\emph{long time} ($\\sim 10^{9-10}$~yr) measurements use stellar astrophysics and paleontology, and (3) \\emph{cosmological} ($\\gtrsim 10^{10}$~yr) experiments use the Big Bang nucleosynthesis and Cosmic Microwave Background anisotropies to compare the value of $G$ in the early Universe to the present value. Generically, they set constraints on $|\\dot{G}/G|$ down to $\\sim 10^{-12}$~yr$^{-1}$, although the comparison between different timescales depends on the assumed form of the function $G(t)$. Here, we review our previously introduced method for setting constraints on $\\dot{G}$ \\citep{jofre06}, to our knowledge the only one so far to probe timescales $\\sim 10^{7-9}~\\mathrm{yr}$, intermediate between the \\emph{human} and \\emph{long} timescales mentioned above. It relies on the change in the internal structure of a neutron star induced by a change in $G$, which, together with the slow response timescale of weak interactions, result in internal heating and an increase in the observed surface temperature. Neutron stars have mean densities exceeding nuclear saturation density, $\\rho_{\\rm nuc}\\sim 3\\times 10^{14}$~g~cm$^{-3}$. In their outer layers, they are composed of heavy atomic nuclei and free electrons, giving way to free neutrons, free protons, muons, and potentially more exotic particles as the density increases inward. A short time after their formation, their internal temperatures drop orders of magnitude below the Fermi energies of free particles ($\\sim 10-100$~MeV), and so their structure is well approximated by zero-temperature models. Nonetheless, weak interactions still play an important role, as characteristic temperatures $10^{6-8}$~K yield non-negligible neutrino emission. Given that the equation of state of matter above nuclear density still remains poorly known, researchers construct models to calculate the evolution of the thermal content of neutron stars and compare their predictions with X-ray observations, in order to set constraints on the models for dense matter (e.g., \\citealt{yakovlev04}). In these thermal evolution models, the structure of the star is calculated assuming an equation of state, and the evolution of the internal temperature is obtained by considering losses due to different neutrino emission processes from the stellar interior, as well as thermal electromagnetic radiation from the surface. We have previously explored \\emph{rotochemical heating}, the effect of the progressive loss of rotational support of millisecond pulsars has on their internal structure, which leads to weak interaction processes (beta decays) and net heating \\citep{reisenegger95,reisenegger97,fernandez05}. A time variation of the gravitational constant has an analogous effect: it compresses the star and therefore also causes heating (\\emph{gravitochemical heating}, \\citealt{jofre06}). In what follows, we discuss how this process constrains $\\dot{G}$. ", "conclusions": "\\emph{Gravitochemical heating} sets constraints on the time variation of the gravitational constant, using the fact that a non-zero change would generate internal heating in neutron stars, which for nearby cases can be detected with existing telescopes. These constraints are the only ones on timescales $10^{7-9}$~yr. If it could be assured that the observed neutron stars cannot have direct Urca reactions, these constraints would be of the same order as the best ones available on other time scales. This is not the case at the moment, but could become so as neutron star interiors become better understood from other studies. In particular, observed temperatures of other old neutron stars (such as millisecond pulsars) will provide useful information." }, "0911/0911.2689_arXiv.txt": { "abstract": "We describe a new algorithm for the ``perfect'' extraction of one-dimensional spectra from two-dimensional (2D) digital images of optical fiber spectrographs, based on accurate 2D forward modeling of the raw pixel data. The algorithm is correct for arbitrarily complicated 2D point-spread functions (PSFs), as compared to the traditional optimal extraction algorithm, which is only correct for a limited class of separable PSFs. The algorithm results in statistically independent extracted samples in the 1D spectrum, and preserves the full native resolution of the 2D spectrograph without degradation. Both the statistical errors and the 1D resolution of the extracted spectrum are accurately determined, allowing a correct $\\chi^2$ comparison of any model spectrum with the data. Using a model PSF similar to that found in the red channel of the Sloan Digital Sky Survey spectrograph, we compare the performance of our algorithm to that of cross-section based optimal extraction, and also demonstrate that our method allows coaddition and foreground estimation to be carried out as an integral part of the extraction step. This work demonstrates the feasibility of current- and next-generation multi-fiber spectrographs for faint galaxy surveys even in the presence of strong night-sky foregrounds. We describe the handling of subtleties arising from fiber-to-fiber crosstalk, discuss some of the likely challenges in deploying our method to the analysis of a full-scale survey, and note that our algorithm could be generalized into an optimal method for the rectification and combination of astronomical imaging data. ", "introduction": "Optical fibers offer a compelling advantage for astronomical survey spectroscopy. By allowing the light from targets over a wide field of view on the sky to be rearranged into a compact format and fed to any number of spectrographs in parallel, they can provide a vast multiplex advantage over imaging spectrographs. For this reason, fiber technology has been adopted for a number of major survey programs such as the Las Campanas Redshift Survey (LCRS: \\citealt{schech_lcrs}), the Two-degree Field Survey (2dF: \\citealt{col_2df}), the Sloan Digital Sky Survey (SDSS: \\citealt{york_sdss}), and the recently initiated Baryon Oscillation Spectroscopic Survey (BOSS: \\citealt{sch_boss}) of the SDSS3 project. The use of fiber spectrographs for faint-object spectroscopy has however been restrained by concerns over throughput and systematic limitations on the quality of subtraction of night-sky emission foregrounds. The spectroscopic extractions of the SDSS multi-fiber instrument have established a high standard, but significant systematic shortcomings remain. Methods to characterize and partially remove sky-subtraction residuals in SDSS fiber spectra can mitigate the problem somewhat (e.g.\\ \\citealt{bol04, wild05}), but do not substitute for a formally correct sky-subtraction model. The faintest spectroscopic galaxy surveys have tended to make use of multi-slit imaging spectrographs (e.g., \\citealt{cow96, steidel03, dav03, gdds1, lef05}), along with sky-subtraction techniques such as nod-and-shuffle \\citep{cuil94, glaz01} or B-spline-based modeling of the two-dimensional sky spectra in the slits (e.g., \\citealt{kel03}). Nod-and-shuffle techniques in particular are ill-suited to most fiber spectrographs, require at least double the detector area, and furthermore reduce the background-limited signal-to-noise by a factor of $1/\\sqrt{2}$ due to their subtraction of data from data. This paper outlines an algorithmic framework for the modeling and extraction of optical and near-infrared astronomical fiber spectroscopy to the limit of photon noise and native instrumental resolution. By comparing our method with the standard techniques of optimal extraction currently in wide use, we identify and resolve key systematic barriers to ``perfect'' extraction. As a result, multifiber spectroscopy emerges as a clear and compelling technique for current and future generations of faint-galaxy spectroscopic surveys, even well below the brightness of the night sky at all wavelengths. This algorithm will deliver significant benefits to the reanalysis of the original SDSS (hereafter SDSS1) archive and to the ongoing analysis of the BOSS survey, both for core redshift-survey goals and for projects that aim to select rare emission-line objects from within the regions of the spectrum dominated by OH line emission from the night sky (e.g., \\citealt{bol04, wil05, bol06, wil06, bol08}). In considering this subject, we will make a clear distinction between the problems of ``calibration'' and ``extraction'': calibration is the description of the way in which any set of astronomical and environmental stimuli translate into the responses of the digital detectors (which we assume here to be pixellated charge-coupled devices, or CCDs); extraction is the reconstruction of particular stimuli from particular responses. More strictly speaking, we view an ``extracted spectrum'' not as a model for the flux of the observed source itself, but rather as a properly calibrated one-dimensional compression of the instrumental response to an observation of that source. When executed and reported correctly, an extracted spectrum permits a statistically valid $\\chi^2$ test of an input model spectrum against the extracted pixels. This paper specifically illustrates a method for carrying out this sort of perfect extraction assuming that perfect calibration is available. We do not mean to trivialize the problem of calibration, and in our concluding remarks we will discuss the relationship of our extraction method to current and future spectroscopic calibration regimes. The paper is organized as follows. Section~\\ref{mod2d} frames the problem of extraction in terms of image modeling, lays out the first part of our algorithm, and compares its performance with that of standard extraction techniques in terms of the quality of their respective models to the raw 2D data. Section~\\ref{resol} addresses the issue of resolution and covariance in the extracted spectra, and presents the second part of our method, which establishes optimal properties in both these regards. Section~\\ref{test4x3} presents a more realistic demonstration of our algorithm on simulated data, illustrating several additional strengths and subtleties of the method. Finally, Section~\\ref{dc} provides a discussion and conclusions. We will observe the following conventions in this paper. Without loss of generality, we will assume that spectroscopic traces run roughly parallel to CCD columns (i.e., ``vertically''), with wavelength increasing with row number and cross-sectional position increasing with column number. We will denote vectors in lowercase bold-face type ($\\mathbf{f}$) and matrices in uppercase bold-face type ($\\mathbf{A}$). We assume all errors have a Gaussian distribution, and we assume no formal priors on fitted model parameters. ", "conclusions": "\\label{dc} The extraction algorithm that we have described does not come without a price. In the presence of fiber-to-fiber cross-talk, the standard row-wise optimal extraction will couple extracted amplitudes between fibers in a single row, leading to a banded matrix of dimension equal to the number of fibers that must be inverted; this process must then be repeated for each row in each exposure. Our 2D modeling extraction, on the other hand, couples extracted amplitudes between fibers, wavelengths, and exposures. Thus the matrix to invert for the solution set of spectra has sides of dimension equal to \\begin{equation} N_{\\mbox{spectra}} \\times N_{\\mbox{samples per spectrum}} \\times N_{\\mbox{exposures}}~. \\end{equation} For one SDSS1 pointing, this would correspond to 320 fibers (one of the two spectrographs), approximately 4000 sampling points, and three exposures: i.e., a square matrix nearly 4 million to a side. With a brute-force approach this matrix could never be stored, let alone inverted, with any hardware of the present or foreseeable future. The way forward to determining the extracted spectra will no doubt lie in exploiting the sparseness of the inverse covariance matrix to reduce storage and computation, and to apply an iterative method such as the conjugate gradient to solve for the extracted spectra. To determine the resolution with which to reconvolve the extracted spectrum, which formally requires the inversion of the full inverse covariance matrix $\\mathbf{C}^{-1}$ , the practical solution will be to invert a sufficient subspace of $\\mathbf{C}^{-1}$ surrounding each spectrum (or subsegment of a spectrum) to define an acceptably accurate approximation to the desired resolution matrix. The exact requirements will depend upon the degree of cross-talk between neighboring fibers in the spectrograph under consideration. Even with these strategies, a usable implementation of our algorithm for real multifiber survey data may require high-performance parallel computing, depending on the computational expense of the necessary matrix-element calculations. Most immediately, we plan to implement the strategy outlined in this paper to the extraction of spectra from the BOSS Survey. We also plan to conduct a reanalysis of the SDSS1 archive, to provide the definitive version of this important spectral database with the best possible extracted resolution, signal-to-noise, and foreground subtraction. These techniques also offer promise for the upcoming Apache Point Observatory Galaxy Evolution Experiment (APOGEE: \\citealt{ap_apogee}) of the SDSS3, a high-resolution, near-infrared multifiber survey of red giant stars in our Galaxy that aims to constrain their evolutionary history as traced by multiple chemical abundances. This extraction strategy will also form a key part of the technical feasibility of the BigBOSS survey \\citep{sch_bigboss}, which proposes to use a 5000-fiber spectrograph fed by a $3^{\\circ}$ field-of-view focal plane positioner system on a 4m-class telescope to measure the baryon acoustic scale and redshift-space distortions over the redshift range $0.2 < z < 3.5$. Although the implementations for these different surveys will differ in detail, we believe that the software engine for the core extraction computations can be written in a general-purpose form. The application of this technique to slit spectroscopy may be possible as well, although the preservation of object spatial information by the slit makes the problem substantially more complicated. In all cases, the full power of this extraction can only be realized with sufficiently accurate calibration. The current standard calibrations for fiber spectroscopy are exposures of uniform spatial illumination by flat-spectrum incandescent lamps (``flats'') and multi-emission line gas-discharge lamps (``arcs''). Within the framework of our extraction algorithm, the former will determine the relative sensitivity of the individual fibers and pixels, and the latter will determine the spectrograph PSF shape and position as a function of illuminating wavelength in each fiber. Assuming there are no systematic offsets between the calibrations and the science exposures, and assuming that the variation of the spectrograph PSF is smooth enough with wavelength to be well-sampled by the arc frames (which are sparsely distributed in wavelength), these calibrations should contain sufficient information to determine the calibration matrix $\\mathbf{A}$. We may proceed by extracting the arcs and flats together, with each one described by a single spectrum projected through an initial guess for $\\mathbf{A}$, and then optimize the parameters of $\\mathbf{A}$ by non-linear iteration so as to improve the quality of the 2D extraction models to convergence. A more direct and ambitious approach to the determination of the calibration matrix would be to illuminate the facility calibration screen with either a high-wattage monochrometer or a tunable laser (c.f.\\ \\citealt{st06, cra09}), and to step the illumination source through wavelength in subsequent exposures so as to map out $\\mathbf{A}$ explicitly. In practice, with the exception of spectrograph systems that are very stable thermally and mechanically, there is likely to be some shifting of the fiber positions and focus on the CCDs between the calibration and science frames. In this case, ``tweaks'' to the parameters of $\\mathbf{A}$ will be derived from the shapes and positions of the fiber traces and night-sky emission lines in the science frames. Finally, we note that the calibration may be significantly improved by incorporating all available knowledge about the optical design of the telescope and instrument, rather than treating the system as a black box to be specified entirely by empirical calibration data. Putting aside the computational challenges that arise from the consideration of continuously two-dimensional input data, the method we have described can also be generalized into an optimal recipe for the rectification and combination of multiple CCD imaging exposures: i.e., taking the $\\mathbf{f}$ of \\S\\ref{mod2d} to be a \\textit{two-dimensional} image model to be extracted from the data. As with the spectroscopic application, the resulting extractions would have optimal resolution, statistically independent extracted image samples, and a definition in terms of the optimization of a well-motivated scalar objective function describing the quality of a model fit to all of the individual exposures. The implementation would be non-trivial, but the benefits could be great. A significant challenge on the calibration side is that the details of the imaging PSF, which must be known, will depend upon the spectral energy distributions of the imaged objects, which will in general be unknown. In summary, we have demonstrated a new spectrum extraction algorithm for optical and near-infrared astronomical fiber spectroscopy. Given sufficiently accurate calibration, this method can extract spectra to the statistical noise limit in the presence of arbitrarily complicated point-spread functions and arbitrarily high and wavelength-varying foregrounds. The extracted spectra have optimal and precisely quantified resolution and signal-to-noise, along with statistically uncorrelated pixels, for any number of sub-exposures in combination together. As such, statistically accurate $\\chi^2$ tests may be made between the extracted data and theoretical models of the input object spectrum. This algorithm represents a fundamental improvement upon the current state-of-the-art methods in use for the extraction of fiber spectroscopy, and thus motivates a serious and positive reevaluation of the promise of fiber-fed spectrographs for next-generation ground-based faint-object surveys." }, "0911/0911.2530_arXiv.txt": { "abstract": "We present an unbiased deep \\OII emission survey of a cluster XMMXCS~J2215.9-1738 at $z=1.46$, the most distant cluster to date with a detection of extended X-ray emission. With wide-field optical and near-infrared cameras (Suprime-Cam and MOIRCS, respectively) on Subaru telescope, we performed deep imaging with a narrow-band filter $NB912$ ($\\lambda_c$ = 9139\\AA, $\\Delta\\lambda$=134\\AA) as well as broad-band filters ($B$, $z'$, $J$ and $K_s$). From the photometric catalogues, we have identified 44 \\OII emitters in the cluster central region of $6'\\times6'$ down to a dust-free star formation rate of 2.6 \\Msun/yr (3$\\sigma$). Interestingly, it is found that there are many \\OII emitters even in the central high density region. In fact, the fraction of \\OII emitters to the cluster members as well as their star formation rates and equivalent widths stay almost constant with decreasing cluster-centric distance up to the cluster core. Unlike clusters at lower redshifts ($z\\la1$) where star formation activity is mostly quenched in their central regions, this higher redshift 2215 cluster shows its high star formation activity even at its centre, suggesting that we are beginning to enter the formation epoch of some galaxies in the cluster core eventually. Moreover, we find a deficit of galaxies on the red sequence at magnitudes fainter than $\\sim$$M^*+0.5$ on the colour-magnitude diagram. This break magnitude is brighter than that of lower redshift clusters, and it is likely that we are seeing the formation phase of more massive red galaxies in the cluster core at $z\\sim1$. These results may indicate inside-out and down-sizing propagation of star formation activity in the course of cluster evolution. ", "introduction": "\\label{sec;intro} Galaxy formation and evolution is strongly dependent on environment and on galaxy mass (e.g., \\citealt{dre97,tan05,coo06,tas09}). Galaxies in high density regions are systematically older than those in lower density regions, and massive galaxies are on average older than lower mass galaxies in stellar ages. It seems that massive galaxies in the high density regions form first and galaxy formation activity is propagated to lower density regions and to lower-mass galaxies with time. Such dependence on environment and mass can be understood by intrinsic effects and extrinsic effects. Intrinsic effects are those determined by the initial condition of galaxy formation and extrinsic effects are those in effect during the evolution of galaxies after formation. The environmental dependence can be partly understood by the intrinsic effect called ``galaxy formation bias'' (e.g., \\citealt{cen93}) where high density regions should have started off from the largest initial density fluctuations that collapse first in the universe, and galaxy formation takes place earliest in such regions and subsequent evolution is prompted. In lower density regions, however, galaxy formation is delayed and time scales of star formation and mass assembly are longer. Likewise, the mass dependence of galaxy formation called ``down-sizing'' (e.g., \\citealt{cow96,kod04a}) may be partly understood by the scaled-down version of the bias on galactic scale, where more massive galaxies today correspond to higher initial density fluctuations on galactic scale and their formation such as star formation and assembly of building blocks take place earlier than less massive galaxies. Galaxies are also subject to external effects from their surrounding environments, such as galaxy-galaxy interactions/mergers and ram-pressure stripping in dense environments (e.g., \\citealt{aba99,qui00} ). Such interactions may enhance and/or quench the star formation activity in galaxies preferentially in high density regions, which would therefore result in conspicuous dependence of galaxy properties on environments. However, the relative importance of the intrinsic effects and the extrinsic effects is almost totally unknown yet. One of the most effective methods to verify the existence of the intrinsic effects is to go back in time and directly see the galaxies in the distant universe. By doing so, we may eventually reach the epoch when galaxies are forming rapidly in the biased cluster core, while galaxies are not yet formed or only slowly forming in lower density regions. Also, as we go back in time, host galaxies of star formation would be shifted to higher mass systems and we may eventually see the active star formation in action in massive galaxies in the cluster core. In the low redshift universe, star formation activity is a monotonically decreasing function of local density. However, we have not yet known the dependence of star formation activity on environments at high redshifts in detail. According to recent studies, it seems that star formation is in fact biased at high density region at high redshifts. \\citet{elb07} reported for the first time in the GOODS North/South surveys that at $z\\sim1$ the galaxies at denser environment tend to have higher star formation rates (SFRs), in contrast to the local Universe. \\citet{coo08} and \\citet{ide09} also showed similar trends for \\OII emitters at $z\\sim1$ in DEEP2 and COSMOS surveys, respectively. At a slightly lower redshift, $z=0.81$, based on the mid-infrared observation of the RX~J1716.4+6708 cluster, \\citet{koy08} reported that the star formation activity is probably enhanced in the medium density region, such as cluster outskirts or galaxy groups, rather than in the highest density region. \\citet{pog08} also suggested that SFRs of galaxies may have a peak at intermediate densities at $z$=0.4--0.8 based on the EDisCS survey. These observational findings may imply that the environment that hosts active star formation is shifted towards denser regions at higher redshifts. On the other hand, it is well-known as the ``Butcher-Oemler (B-O) effect'' that more distant clusters show higher fraction of blue galaxies up to $z\\sim0.5$, suggesting the enhancement of star formation activity in higher-$z$ clusters (e.g., \\citealt{bo78,bo84,mar01}). However, the fraction of blue galaxies decreases with cluster centric radius, and red galaxies still dominate in the core regions at $z\\sim0.5$ (\\citealt{kod01,ell01}). Such studies of the B-O effect in clusters have been extended up to $z\\sim1$. \\citet{pos01} found that the fraction of active galaxies in the central regions of clusters is higher at $z\\sim$0.7--0.9 compared to $z\\sim$0.2--0.5. However, \\citet{nak05} suggested that the fraction of galaxies with strong \\OII emission in the cluster cores at $z<1.0$ is not significantly dependent on redshift. It thus seems that we have not yet reached a bursting phase of massive galaxy formation in cluster cores even at $z\\sim1$, although the global activity of star formation within clusters is already enhanced. In this paper, we present an \\OII emitter survey of the XMMXCS~J2215.9-1738 cluster (hereafter 2215 cluster) with a narrow-band filter $NB912$ on Suprime-Cam (Figure \\ref{fig;transmission}), and discuss spatial distribution of star formation activity (see also \\citealt{koy09}, which presents our \\Ha emitter survey of the RX~J1716.4+6708 cluster at $z=0.81$). The 2215 cluster is the most distant cluster to date at $z$=1.46 with a detection of extended X-ray emission (\\citealt{sta06}). \\citet{hil07,hil09} have confirmed dozens of the cluster members by spectroscopy, and found that velocity dispersion of the member galaxies is $\\sigma=580\\pm140$ [km s$^{-1}$]. With the $NB912$ filter, we can survey \\OII emissions from the galaxies with the line-of-sight velocities between $-$2794 $<$ $\\Delta V_{\\rm los}$ [km s$^{-1}$] $<$ 1598 with respect to the velocity centre of the cluster. Our $NB912$ filter thus perfectly matches this cluster, and should be able to detect all the \\OII emission lines from the cluster members to a certain flux limit without introducing any bias (Figure \\ref{fig;transmission}). Therefore, our survey is unique, and the 2215 cluster is an ideal target for us to investigate the environmental dependence of star formation activity over a wide range in environment at this high redshift. The X-ray luminosity and inter-cluster medium temperature for the 2215 cluster are $L_X=4.4^{+0.8}_{-0.6}\\times10^{44}$[erg\\ s$^{-1}$], and $kT=7.4^{+2.6}_{-1.8}$[keV], respectively (\\citealt{sta06}). The luminosity is fainter than what is expected from the temperature compared to the local $L_X-T$ relation. \\citet{hil07} thus point out that this cluster may experience a merger within the last few Gyr. It is also found that colour-magnitude diagram in this cluster shows red sequence (\\citealt{sta06,hil09}). \\citet{hil09} investigated the morphologies of bright member galaxies, and found that about 60\\% of members are E or S0 galaxies. Even at $z=1.46$, cluster core is already dominated by early-type galaxies, as far as morphology is concerned. The structure of this paper is as follows. The observations and data reduction are described in \\S~\\ref{sec;obs}. In \\S~\\ref{sec;selection}, we describe how we select the \\OII emitters associated to the cluster at $z=1.46$ from the photometric catalogues. We show the spatial distribution, SFRs, and equivalent widths of the \\OII emitters, and investigate the star forming activity in the 2215 cluster in \\S \\ref{sec;results}. We also discuss the deficit of faint red galaxies and its evolution. A summary is given in \\S \\ref{sec;summary}. Throughout this paper, magnitudes are in the AB system, and we adopt cosmological parameters of $h=0.7$, $\\Omega_{m0}=0.3$ and $\\Omega_{\\Lambda 0}=0.7$. Vega magnitudes in $J$ and $K_s$, if preferred, can be obtained from our AB magnitudes using the relations: $J$(Vega)=$J$(AB)$-$0.92 and $K_s$(Vega)=$K_s$(AB)$-$1.80. \\begin{figure} \\begin{center} \\includegraphics[width=70mm]{transmission.eps} \\end{center} \\caption{A transmission curve of the $NB912$ filter ($\\lambda_c$=9139\\AA, $\\Delta\\lambda$=134\\AA). Upper x-axis shows peculiar velocity with respect to the velocity centre of the cluster at $z=1.46$. The histograms show the velocity distribution of the spectroscopically confirmed member galaxies of the 2215 cluster (\\citealt{hil09}). } \\label{fig;transmission} \\end{figure} ", "conclusions": "\\label{sec;results} \\subsection{Star formation activity in a cluster core} \\label{sec;sf_activity} \\begin{figure} \\begin{center} \\includegraphics[width=70mm]{r_fraction_linear_oii.eps} \\end{center} \\caption{Fraction of \\OII emitters to the cluster members as a function of radius from a cluster centre. The contribution of field galaxies is statistically subtracted from our colour-selected cluster member candidates. The error-bars indicate the statistical errors associated to both \\OII emitters and the cluster members. } \\label{fig;oii_fraction} \\end{figure} \\begin{figure} \\begin{center} \\includegraphics[width=84mm]{comparison_2215-1716.eps} \\end{center} \\caption{Close-up views of the central $2'\\times2'$ regions for XCS2215 cluster at $z=1.46$ (left) and RXJ1716 cluster at $z=0.81$ (right). Blue open squares show \\OII emitters for 2215 cluster and \\Ha emitters for RXJ1716 cluster, respectively. Dots show cluster member candidates selected in \\S\\ref{sec;clmember} for the 2215 cluster, and member galaxies selected by photometric redshifts of $\\Delta z=0.76-0.83$ for RXJ1716 cluster (\\citealt{koy07}), respectively. Both panels show the areas of similar physical scales (0.51Mpc/arcmin at $z=1.46$ and 0.45Mpc/arcmin at $z=0.81$, respectively). } \\label{fig;1716} \\end{figure} \\begin{figure*} \\includegraphics[width=55mm]{radius_sfr_oii_linear.eps} \\includegraphics[width=55mm]{radius_ssfr_oii_linear.eps} \\includegraphics[width=55mm]{radius_ew_oii_linear.eps} \\caption{ SFRs(left), specific SFRs (SSFRs) (middle) and observed equivalent widths (right) for \\OII emitters as a function of distance from the cluster centre. SFRs are derived and corrected for dust extinction using the relations in \\citet{mou06} (see text). Stellar masses are estimated using the equation of \\citet{dad04}. Equivalent widths are calculated with the equations (\\ref{eq;fline}) and (\\ref{eq;fcont}). Blue open squares show 44 \\OII emitters, and 4 red \\OII emitters are marked by magenta filled square. Typical errors on each value are shown in the upper or the lower right corner on each panel. The errors on SFR and equivalent width are derived from magnitude errors in $NB912$ and $z'$, and $\\sigma(\\Delta \\log({\\rm M_\\star}))=0.20$ is applied to the error in stellar mass (see text). In each diagram, the solid line connects median values in four radius bins, and the errors contain both the typical error on each value and the standard deviation in each bin. } \\label{fig;sfr_ew} \\end{figure*} Figure \\ref{fig;map_member} shows that there are many \\OII emitters in the central region of the 2215 cluster. If we assume that \\OII lines are emitted from ionized gas in/around the star forming regions, it seems that the 2215 cluster is still actively forming stars even at its core region. This is not the case in lower-$z$ clusters where star forming activity is much lower in the central region. To be more quantitative, we estimate a fraction of \\OII emitters to cluster members as a function of cluster-centric distance in Figure~\\ref{fig;oii_fraction}. Here, the amount of remaining contamination from foreground/background galaxies are estimated from the SDF sample, and it is statistically subtracted from the colour-selected sample of cluster member candidates. Figure \\ref{fig;oii_fraction} suggests that the 2215 cluster maintains a high fraction of star forming galaxies in the central region, $\\sim 30$\\%, even at the most inner part within 0.25~Mpc in physical scale (1'=0.51Mpc at $z=1.46$). This fraction is higher than the \\OII fraction in the cluster cores in \\citet{nak05}, which is $\\la 20$\\% at $z<1.0$. \\citet{lid08} recently reported a distribution of spectroscopic sample of \\OII emitters in a cluster XMMU~J2235.3-2557 at $z=1.39$, a similar redshift to that of the 2215 cluster. While there are no galaxies in the very core within 90~kpc in radius that show star forming activity, \\OII lines are detected from more than half of galaxies near the centre. This result is similar to ours on the \\OII emitters in 2215 cluster at a similar (slightly higher) redshift of 1.46. It seems likely that distant X-ray detected clusters at $z\\ga1.4$ are actively forming new stars even in the central region within a few hundreds kpc in radius. We have recently conducted a narrow-band \\Ha emitter survey for RX~J1716.4+6708 cluster at $z$=0.81 (\\citealt{koy09} in preparation). This survey reveals quite a different situation where no \\Ha emitters are observed in the core region within a radius of 0.25~Mpc (Figure \\ref{fig;1716}). A difference in spatial distribution of the emitters is very impressive. The \\Ha emitters in the 1716 cluster are selected with a combination of $NB119$ ($\\lambda_c$ = 11885\\AA, $\\Delta\\lambda$=141\\AA) and $J$ filters on Subaru/MOIRCS. The \\Ha survey reached to the depth of a limiting line flux of $\\sim$4.1$\\times$10$^{-17}$ erg s$^{-1}$ cm$^{-2}$, which corresponds to a dust-free SFR of $\\sim1.0$\\Msun yr$^{-1}$. The \\Ha survey for the 1716 cluster is, thus, more sensitive to star forming galaxies with slightly lower SFR than the \\OII survey for the 2215 cluster, presented in this paper. Although the lines are different between the two clusters, \\Ha and \\OII lines, considering the fact that \\Ha line is much less affected by dust extinction than \\OII line, the intrinsic difference in spatial distribution of the emitters would be more significant. Furthermore, \\citet{koy09} find that the fraction of star forming galaxies decreases as one goes to denser region, which is different from what we see for the 2215 cluster. Similar trends are also found in lower redshift clusters at $z$=0.4--0.8 (\\citealt{kod04b,pog08}). \\citet{koy08,koy09} have also unveiled the star forming activities hidden by dust in the 1716 cluster based on the mid-infrared imaging with AKARI. These studies conclude that star formation activity has already been quenched in the cluster core, while it comes to an peak in the medium density regions away from the cluster core. These facts may imply that we find galaxy formation bias in the highest density region at $z\\sim1.5$. Recent studies also support the biased star formation in a relatively dense environment at $z>1$ (\\citealt{elb07,coo08,ide09}). The difference of critical environments in star formation at various redshifts may suggest that galaxy formation bias plays an important role in the dependence of galaxy properties on environments. In such a comparison between different clusters at different redshifts, we must keep in mind that the size and mass of the clusters can be different. In fact, the bolometric X-ray luminosity of 1716 cluster is $\\sim3$ times larger than that of 2215 cluster (\\citealt{ett04,sta06}). Also, the sub-structures of 1716 cluster is more prominent (\\citealt{koy07}). We cannot therefore conclude whether the difference seen in spatial distribution is largely due to time evolution or due to different masses or characteristics. But it may be the case that we are witnessing the evolution in star forming activity in the core of high-$z$ clusters from $z\\sim1.5$ to $z\\sim0.8$. On the other hand, \\citet{fin05} report a diversity in distribution of \\Ha emitters in the cores of clusters at $z$=0.7--0.8. Two among their three clusters at similar redshifts show a decrease in \\Ha fraction toward the cluster centre, while the other cluster shows an opposite trend. This fact may imply that it is not only the star formation bias that causes the high fraction of \\OII emitters in the core of the 2215 cluster. It is thus crucial to observe more clusters at $z>1.0$, and to evaluate to what extent the presence of star forming galaxies in the cluster core is a general property at $z>1$. The $L_X-T$ relation of the 2215 cluster suggests that it is possible that it experienced a merger event within the last few Gyr (\\citealt{hil07}). This can be another possible reason for the high fraction of \\OII emitters, since the cluster merging event might cause an enhancement of star formation activity in galaxies in the cluster core. Also, since the distribution of \\OII emitters is projected on celestial sphere, it may be possible that some member galaxies in the outskirt of the cluster happen to be seen in the direction to the cluster centre, which may result in apparent high fraction of \\OII emitters at the centre. However, the number density of galaxies in the outskirt is likely to be much lower than that of galaxies in the core region (Figure \\ref{fig;bzk}(b)), and such projection effect would be small in the direction to the centre. It is also possible that an active galactic nucleus (AGN) enhances \\OII line flux. Recent studies suggest that a fraction of AGNs in clusters increases with redshifts. \\citet{gal09} have found an overdensity of AGNs within a radius of 0.25~Mpc in clusters at $z>0.5$, and that the density of X-ray selected AGNs in clusters at $1.01.4$ using K20 survey data. Mass to $K$-band luminosity ratio is derived from multi-wavelength data from $U$ to $K$ with \\citet{sal55} IMF (\\citealt{fon04}). The relation is calibrated with $z-K$ colour, which can reduce the dispersion of derived stellar mass to $\\sigma(\\Delta \\log({\\rm M_\\star}))=0.20$. We also derive specific SFR by dividing SFR by stellar mass. Figure \\ref{fig;sfr_ew}(b) shows the specific SFRs for \\OII emitters as a function of radius from a centre. The observed equivalent width of an \\OII emission is also derived from \\OII line flux and continuum flux density (Equations (\\ref{eq;fline}) and (\\ref{eq;fcont})). Figure \\ref{fig;sfr_ew}(c) shows the observed equivalent widths for the \\OII emitters as a function of radius from a centre. Figure \\ref{fig;sfr_ew} suggests that star formation activities of the \\OII emitters in denser regions are as active as those in the outskirt of the cluster. The equivalent widths are not correlated with the position of galaxies in the cluster. These facts also suggest active star formation in the central region of the 2215 cluster. On the other hand, the specific SFRs may show a mild correlation in the sense that \\OII emitters at the inner regions tend to have lower specific SFRs. Since the similar correlation is not seen in their SFRs, this is due to the fact that massive galaxies tend to reside near the centre of the cluster. This tendency may be due to the galaxy formation bias that massive galaxies are formed in the cluster core, or may be a result of efficient mass assembly due to merging that can be more effective in high density regions. Figure \\ref{fig;sfr_ew}(b) also shows that the red \\OII emitters tend to have lower specific SFRs. This may be because we underestimate their intrinsic SFRs due to the strong dust extinction for the red galaxies. Otherwise, this may imply that these galaxies are just quenching their star formation activities. \\subsection{Colour-magnitude diagram} \\label{sec;cmd} \\begin{figure*} \\includegraphics[width=70mm]{kzkcolor_clmember-Kdet_center.eps} \\includegraphics[width=70mm]{lf_complete_clmemred_center_Ks-det.eps} \\caption{(a) Left panel: Colour-magnitude diagram of $z'-K_s$ vs. $K_s$ for the $K_s$-detected galaxies in the central $2'\\times2'$ region. Black dots shows cluster member candidates. Red solid line show the expected location of the colour-magnitude relation of passively galaxies galaxies at $z=1.46$ formed at $z_{form}$=5 (\\citealt{kod98}). We define the red sequence galaxies as those falling between the two broken lines, $\\Delta(z'-K_s)=\\pm0.3$ around the colour-magnitude relation. Long-dashed and dotted lines are 3$\\sigma$ and 5$\\sigma$ confidence levels in colours, respectively. Blue open squares show 18 \\OII emitters in this region. Three dot-dashed lines indicate iso-stellar mass curves for $2\\times10^{11}$\\Msun, $10^{11}$\\Msun and $5\\times10^{10}$\\Msun, respectively, which are drawn based on the equation given in \\citet{dad04}. (b) Right panel: $K_s$-band luminosity function of the red sequence galaxies in the central $2'\\times2'$ region. The contribution of field galaxies, estimated from the SDF sample, is statistically subtracted. Error-bars show Poisson errors. } \\label{fig;cmd_lf} \\end{figure*} Figure \\ref{fig;cmd_lf}(a) shows a colour-magnitude diagram of $z'-K_s$ vs. $K_s$ for cluster member candidates in the central $2'\\times2'$ region. Note that the $K_s$-selected galaxy sample is used in this section. The use of the $NB912$-selected sample may cause incompleteness in the number of galaxies at faint $z'$ magnitudes. The solid line in the figure shows the expected location of the colour-magnitude relation (CMR) of passively evolving galaxies at $z=1.46$ formed at $z_{form}=5$ inferred from the \\citet{kod98} model which is calibrated to reproduce the CMR of elliptical galaxies in the Coma cluster at $z=0$; \\begin{equation} z'-K_s=-0.088K_s+4.56. \\label{eq;cmr} \\end{equation} We define the red sequence galaxies as those falling in-between $\\Delta(z'-K_s)=\\pm0.3$ from the predicted CMR as shown by the broken lines in Figure \\ref{fig;cmd_lf}(a). The number of the red sequence galaxies seems to decrease at magnitude fainter than $K_s\\sim$21.5. The long-dashed and dotted lines show the 5$\\sigma$ and 3$\\sigma$ confidence levels of $z'-K_s$ colours, respectively. Therefore the decrease in the number of the red galaxies is not driven by incompleteness. \\citet{sta06} and \\citet{hil09} also examine colour-magnitude diagrams, and find that there is a well defined CMR of red galaxies in the 2215 cluster. In \\citet{sta06}, however, shallowness of the data does not allow us to discuss the faint end of the red sequence. \\citet{hil09} use deeper data to investigate the CMR, and colour-magnitude diagram of Figure~6 in \\citet{hil09} shows a deficit of red galaxies at magnitudes fainter than $K_s\\sim21.5$. This is consistent with our result. Since all the cluster member candidates are plotted in Figure \\ref{fig;cmd_lf}(a), some galaxies actually do not belong to the cluster. In order to correctly evaluate the deficit of the red galaxies, we derive the field-corrected $K_s$-band luminosity function in the central $2'\\times2'$ region by statistically subtracting the field contamination. The detection completeness is also corrected for as follows. Artificial objects with Gaussian profile are randomly generated and embedded in the raw $K_s$ image, and detection of these objects is conducted with the same manner as in \\S\\ref{sec;selection_nbcatalogue}. The assumed distribution of brightness of the artificial objects is uniform within each magnitude bin. The completeness thus estimated in each bin is always as high as $\\ga$90\\%. The resulting luminosity function of the red sequence galaxies is shown in Figure \\ref{fig;cmd_lf}(b). It is clear that the number of red member galaxies decreases at $K_s\\la 21.5$. This magnitude corresponds to $K_s^*$+0.5 with respect to the passively evolving galaxies at $z=1.46$ \\citep{kod98}. In the 2215 cluster at $z=1.46$, the red sequence is visible only down to $\\sim$$K_s^*$+0.5 and truncated at that magnitude. Such truncation is seen, if any, at fainter magnitudes ($\\ga M^*$+0.5) in the cores of lower redshift clusters. For example, \\citet{and06} find no deficit of galaxies on red sequence down to $M^*+3.5$ in the MS1054.4-0321 cluster at $z$=0.83. \\citet{koy07} suggest that the build-up of the CMR depends on X-ray luminosity of clusters at $z\\sim0.8$, and find that the CMR is established down to $M^*+2.0$ even for a X-ray fainter cluster RXJ1716+6708 at $z=0.81$. \\citet{lid04} and \\citet{tan08} study CMR for the RDCS~J1252.9-2927 cluster at $z$=1.24. While the red sequence galaxies only appear down to $K\\sim22$ in sub-clumps in the outskirt of the cluster (\\citealt{tan08}), faint red galaxies clearly exist down to $K\\sim24$ in the main cluster (\\citealt{lid04}). \\citet{lid08} also find that there are red faint galaxies down to $K\\sim24.5$ in the core region of the XMMU~J2235.3-2557 cluster at $z$=1.39. We also note that \\citet{kod07} show that the CMR of proto-clusters at much higher redshifts ($2\\la z\\la 3$) becomes less conspicuous, and even the bright-end of the red sequence seems to disappear in proto-clusters at $z\\ga 2.5$. In Figure \\ref{fig;cmd_lf}(a), we mark the \\OII emitters with blue open squares. Most of the \\OII emitters have bluer colours as expected since they are likely to be still forming stars. The galaxies that are fainter in $K_s$ hence less massive galaxies tend to be slightly bluer. This may suggest that star formation activity in more massive galaxies tends to be truncated at earlier times. We draw three iso-stellar mass curves of $2\\times10^{11}$\\Msun, $10^{11}$\\Msun\\ and $5\\times10^{10}$\\Msun, respectively, using the equation given in \\citet{dad04}. If the \\OII emitters cease their star formation, they would move along these curves until they reach on to the red sequence. If an \\OII emitter has a SFR of 50\\Msun/yr, which is close to the median SFR in our \\OII emitter sample, and keeps this rate constant until $z=1.0$, its stellar mass would increase by $\\Delta$M$_\\star$$\\simeq7\\times10^{10}$\\Msun. If we assume that an \\OII emitter gradually becomes red while increasing its stellar mass at a constant SFR, the faint end of the red sequence would be filled up by $z\\sim1.0$. If this is the case, the \\OII emitters with $\\la5\\times10^{10}$\\Msun\\ may be good progenitors of the faint galaxies on the red sequence. However, we do not know yet the mechanisms of changing galaxy colours and quenching their star formation. As described in the last paragraph of \\S\\ref{sec;sf_activity}, contribution of AGNs to the \\OII emitters on the red sequence can be large (\\citealt{yan06}). Perhaps AGN feedback is one of the key mechanisms to reduce the star formation activities. Combining with the previous studies of CMR (see above), we come up with the following scenario of formation of red sequence in clusters at $z\\sim1.5$. The most massive galaxies brighter than $K_s^*$ (i.e., $>10^{11}$\\Msun) are formed at $z>2$, and become red by quenching the star formation in early epoch. Some galaxies on the red sequence may still be keeping residual star formation activities. On the other hand, less massive galaxies are actively growing at $z<2$, and we hardly see galaxies fainter than $K_s^*$ on the red sequence. Star forming galaxies with $\\sim5\\times10^{10}$\\Msun\\ at $z\\sim1.5$ may evolve into faint red sequence galaxies by $z\\sim1.0$. This suggest down-sizing propagation of star formation in high redshift clusters. We performed a unique, unbiased \\OII line survey of star forming galaxies in the XMMXCS~J2215.9-1738 cluster at $z=1.46$, which is currently the most distant cluster ever identified with a detection of extended X-ray emission. We have obtained wide-field optical ($B$, $z'$, $NB912$) and near-infrared ($J$ and $K_s$) data with Suprime-Cam and MOIRCS, respectively. With a combination of $NB912$ narrow-band filter ($\\lambda_c$ = 9139\\AA, FWHM=134\\AA) and the $z'$-band filter, we detect 69 $NB912$ emitters in the central $6'\\times6'$ region where near-infrared data are also available. Among them, 44 emitters are identified as \\OII emitters associated to the cluster based on the $B-z'$ and $z'-K_s$ colours, down to a dust-free star formation rate of 2.6 \\Msun yr$^{-1}$ (3$\\sigma$). We find that many \\OII emitters reside in the central high density region even within a radius of 0.25 Mpc (physical scale). We also find that the fraction of \\OII emitters to cluster members remains high up to the core region. This suggests that the 2215 cluster is still actively forming stars even at the central region, in contrast to lower redshift clusters, where old passively evolving elliptical galaxies dominate. This indicates an inside-out propagation of star formation in high redshift clusters, and we may be eventually beginning to enter the epoch of biased galaxy formation in the densest region at $z=1.46$. SFRs, specific SFRs, and equivalent widths are derived for 44 \\OII emitters. It is found that the emitters have similar SFRs and equivalent widths irrespective of the location within the cluster. It seems however that the specific SFRs tend to decrease slightly toward the cluster core, probably due to the fact that more massive \\OII emitters exist near the centre. We may be approaching to the formation phase of massive galaxies at the cluster core. Moreover, the colour-magnitude diagram in the 2215 cluster shows a deficit of red sequence galaxies fainter than $\\sim$$M^*+0.5$, while the red sequence in lower redshift clusters extends to much fainter magnitudes. While some bright \\OII emitters are located on the red sequence, all the faint \\OII emitters with $>$$M^*$ have blue colours. It is likely that those blue \\OII emitters become redder once they truncate their star formation, and they would eventually reach and fill the faint end of the red sequence at lower redshifts. This indicates a down-sizing propagation of star formation in high redshift clusters." }, "0911/0911.0703_arXiv.txt": { "abstract": "The combination of high spatial and spectral resolution in optical astronomy enables new observational approaches to many open problems in stellar and circumstellar astrophysics. However, constructing a high-resolution spectrograph for an interferometer is a costly and time-intensive undertaking. Our aim is to show that, by coupling existing high-resolution spectrographs to existing interferometers, one could observe in the domain of high spectral and spatial resolution, and avoid the construction of a new complex and expensive instrument. We investigate in this article the different challenges which arise from combining an interferometer with a high-resolution spectrograph. The requirements for the different sub-systems are determined, with special attention given to the problems of fringe tracking and dispersion. A concept study for the combination of the VLTI (Very Large Telescope Interferometer) with UVES (UV-Visual Echelle Spectrograph) is carried out, and several other specific instrument pairings are discussed. We show that the proposed combination of an interferometer with a high-resolution spectrograph is indeed feasible with current technology, for a fraction of the cost of building a whole new spectrograph. The impact on the existing instruments and their ongoing programs would be minimal. ", "introduction": "In recent years optical interferometers have proven that they can produce excellent science in the field of stellar and circumstellar astrophysics. Over the same period high-resolution spectrographs have enabled the discovery of the first extra-solar planets, and contributed substantially to great progress in the field of asteroseismology. A number of current interferometric instruments have some spectroscopic capabilities. For example the mid- and near-infrared instruments MIDI (The Mid-Infrared instrument, at the VLTI) and AMBER (Astronomical Multiple BEam Recombiner, at the VLTI) provide spectral resolutions of up to $R\\sim250$ and $R\\sim12 000$, over bandpasses of $\\sim 5$~$\\mu$m and $\\sim 50$~nm, respectively. At the CHARA array, the Vega (Visible spEctroGraph and polArimeter) project is under construction with a spectral resolution of $R\\sim30 000$ and a bandpass of $\\sim 50$~nm. For science results obtained with spectrally resolved interferometry see for example \\cite{vakili1998} and \\cite{weigelt2007}. Unfortunately the combination of very high spectral resolution over a bandpass greater than a few tens of nanometer, to enable a real analog to classical Echelle spectroscopy with interferometric spatial resolution is not yet available. In building a dedicated high-resolution Echelle spectrograph for an existing interferometer such as the VLTI, one would face several challenges. First of all, building a high-resolution spectrograph is a very costly and time-intensive undertaking. In addition it would be hard to justify building such an instrument only for use with an interferometer, as current optical interferometers are still restricted to very bright objects in comparison with single telescopes. Furthermore, high-resolution spectrographs are usually large instruments, while the space available in the beam combining laboratories of interferometers is often limited. Therefore, it is unlikely that such a dedicated instrument will be built in the near future. In this article we advocate a different approach. By using an existing spectrograph and only building an interface between it and an interferometer at the same site, the combination of high spectral and spatial resolution could be achieved on a much shorter timescale, and for a fraction of the cost of a complete new instrument. The pre-existing infrastructure would need to consist of two telescopes, delay lines for path compensation, a fringe sensing unit to acquire and stabilize the fringes, and a high-resolution spectrograph on the same site. These conditions are already fulfilled, or will be fulfilled in the very near future, at several observatories. The two most promising sites are: \\begin{itemize} \\item In the Southern hemisphere at Paranal Observatory with the VLTI in combination with the UVES spectrograph at Unit Telescope 2 (UT2), or the High-Resolution IR Echelle Spectrometer (CRIRES) spectrograph at UT1. \\item In the Northern hemisphere at Mauna Kea with the Keck Interferometer (KI) and HIgh Resolution Echelle Spectrometer (HIRES) at Keck I telescope or the NIRSPEC spectrograph at Keck II. Also at Mauna Kea, the OHANA (Optical Hawaiian Array for Nanoradian Astronomy) interferometer is currently under development, and will allow the combination of several other pairs of telescopes at the Manual Kea site. \\end{itemize} In each case, three additional hardware components would be required: \\begin{enumerate} \\item A beam combiner that accepts two input beams from the telescopes and feeds the outputs carrying the fringe signals (coded as intensity variations) into fiber feeds; \\item Fibers that connect the interferometer to the spectrograph; \\item A fiber head that feeds the light from the fibers into the spectrograph. \\end{enumerate} If separate telescopes are available for interferometry (as in the case of the Auxiliary Telescopes of the VLTI), the impact on the single-telescope-use of the spectrograph would be minimal, as it could be used in the interferometric mode during times when other instruments are scheduled for use with the main telescope. The outline of this article is as follows. In Section~\\ref{sect:science} the scientific motivation for the proposed setup will be discussed. Section~\\ref{instrument} will give an overview of the challenges in designing and building the proposed instrument, and their possible solutions. In Section~\\ref{sect:test_case:_vlti_uves} we will give more information about the proposed combination of the VLTI with the UVES spectrograph, and its expected performance. Partial and preliminary results of this work have been presented in \\cite{quirrenbach2008}. Section \\ref{sect:other_interferometer-spectrograph_pairings} highlights the main points for other possible interferometer-spectrograph pairings, and Section~\\ref{Conclusion} gives our conclusions. ", "conclusions": "\\label{Conclusion} This article shows that the combination of high-resolution spectroscopy with long-baseline interferometry gives access to hitherto unobservable properties of stellar surfaces and circumstellar matter. Furthermore the article shows how this combination can be achieved without building a major new instrument; only a beam combiner and a fiber link are needed. No instrument components are required which would be expensive or time-consuming to make. The use of external fringe tracking is essential for an instrument like this as it enables long integration times. Time-varying longitudinal dispersion could severely limit the possibilities of such an instrument. Dispersion compensation techniques are investigated in this article and a solution is presented which allows integration times up to a few minutes. The implementation of this approach is shown for the example combination UVES-VLTI. The resulting instrument would differ from other instruments or efforts taken to achieve high spatial and spectral resolution. It would offer spectral resolution nearly a factor 2 higher than for any other interferometric instrument, and over a wide spectral range of a few hundred rather than a few tens of nanometers. It offers the same possibilities to an astronomer as a high-resolution Echelle spectrograph behind a single telescope does, plus the high spatial resolution due to the interferometer, albeit only for bright targets. However it is important to realize that this concept, the combination of two existing instruments to create a new one, is not limited to a specific location or instrument, but rather is an approach which can be followed at different observatories in both hemispheres. It is worth pointing out that the measurements taken with the proposed instruments are differential in nature (e.g. change of visibility amplitude and visibility phase over spectral lines), allowing many interferometric calibration problems to be circumvented. Therefore science relying on absolute V$^{2}$ measurements would need additional data form other instruments. However as noted in Section~\\ref{sect:science}, differential measurements contain a wealth of astronomical information in the optical/IR regime." }, "0911/0911.1533_arXiv.txt": { "abstract": "We apply ionization balance and MHD calculations to investigate whether magnetic activity moderated by recombination on dust grains can account for the mass accretion rates and the mid-infrared spectra and variability of protostellar disks. The MHD calculations use the stratified shearing-box approach and include grain settling and the feedback from the changing dust abundance on the resistivity of the gas. The two-decade spread in accretion rates among Solar-mass T~Tauri stars is too large to result solely from variations in the grain size and stellar X-ray luminosity, but can plausibly be produced by varying these parameters together with the disk magnetic flux. The diverse shapes and strengths of the mid-infrared silicate bands can come from the coupling of grain settling to the distribution of the magneto-rotational turbulence, through the following three effects. First, recombination on grains 1~$\\mu$m or smaller yields a magnetically-inactive dead zone extending more than two scale heights from the midplane, while turbulent motions in the magnetically-active disk atmosphere overshoot the dead zone boundary by only about one scale height. Second, grains deep in the dead zone oscillate vertically in wave motions driven by the turbulent layer above, but on average settle at the rates found in laminar flow, so that the interior of the dead zone is a particle sink and the disk atmosphere will become dust-depleted unless resupplied from elsewhere. Third, with sufficient depletion, the dead zone is thinner and mixing dredges grains off the midplane. The last of these processes enables evolutionary signatures such as the degree of settling to sometimes decrease with age. The MHD results also show that the magnetic activity intermittently lifts clouds of small grains into the atmosphere. Consequently the photosphere height changes by up to one-third over timescales of a few orbits, while the extinction along lines of sight grazing the disk surface varies by factors of two over times down to a tenth of an orbit. We suggest that the changing shadows cast by the dust clouds on the outer disk are a cause of the daily to monthly mid-infrared variability found in some young stars. ", "introduction": "} Concentration of the primordial solid material is a fundamental requirement for planet formation. The interstellar medium contains about 1\\% solids by mass in the form of dust grains, and the Sun has a similar abundance of heavy elements, while the terrestrial planets consist almost entirely of refractory material. Even the most gas-rich Solar system planet, Jupiter, has between 1.5 and 6~times the Solar abundance of heavy elements \\citep{sg04}. A basic concentration process operating in the Solar nebula and other protostellar disks is the settling of dust particles due to the component of the stellar gravity perpendicular to the disk plane. The grains settle at different speeds depending on their ratio of mass to area, leading to mutual collisions. If the disk gas is laminar and the initially sub-micron particles readily stick together, then most of the solid material settles into a thin midplane layer within a few tens of thousands of years at 1 to 10~AU \\citep{w80,nn81,ns86}. Such rapid loss of the grains conflicts with observations of scattered starlight showing some dust remains suspended in the atmospheres of million-year-old protostellar disks \\citep{bs96,sm03,ws04}. However mid-infrared colors suggest the dust abundance is reduced near the disk photosphere, consistent with the incorporation of some material into larger bodies in the interior \\citep{fh06}. The smaller particles can remain aloft if stirred by turbulence in the gas \\citep{dd04b}. Turbulence is also commonly invoked to drive the observed flow of material onto the central star. In particular, the turbulence resulting from the magneto-rotational instability \\citep[MRI;][]{bh91,bh98} can provide the necessary outward transfer of orbital angular momentum. However, the MRI requires a sufficient level of ionization for coupling the disk gas to the magnetic fields. The weak internal heating means thermal ionization is effective only very near the star, while the interstellar cosmic rays and stellar X-rays ionize only the top and bottom surfaces of the disk, failing to penetrate the interior \\citep{g96,ig99}. Furthermore, the recombination of the free charges on grain surfaces is so efficient that a small mass fraction of sub-micron particles is sufficient to shut off MRI turbulence in much of the region where the planets formed \\citep{sm00}. As a result, protostellar disks consist of a laminar dead zone sandwiched between two turbulent active layers. The dead zone extends from 0.1 to 15~AU in the minimum-mass Solar nebula, given micron-sized grains and ionization by the median stellar X-ray luminosity \\citep{td09}. This paper deals with the consequences of the active and dead layers for dust settling and disk evolution. We investigate how the gas flows govern the distribution of the grains, and how the grains in turn affect the dead zone size, using the methods outlined in \\S\\ref{sec:methods} to make ionization balance and 3-D MHD calculations of small patches of the disk around a T~Tauri star. Since a wealth of data is now available on the abundance, size and composition of dust in the surface layers of protostellar disks overlying the regions where planets are likely to form \\citep{fh06,sw06,bh08,wl09}, we examine the observational signatures of the movements of the grains, seeking to understand the following issues. \\begin{enumerate} \\item What is the cause of the large measured range in accretion rate at a given stellar mass (\\S\\ref{sec:massflow})? Solar-mass T~Tauri stars grow at rates between about $10^{-9}$ and $10^{-7}$~M$_\\odot$ \\citep[compiled by][]{ml05}, while simple magnetic accretion models have a fixed column of accreting gas set by the penetration depth of the ionizing radiation, implying a unique mass flow rate \\citep{hd06}. \\item Why do T~Tauri stars show such diverse mid-infrared spectra (\\S\\ref{sec:movements})? The variety in the silicate band strengths and shapes indicates wide ranges in the abundances of small and crystalline particles. High crystalline mass fractions appear more often in systems having colors consistent with grains concentrated near the midplane \\citep{fh06,wl09}. \\item What are the origins of the variability over intervals of days to years in the 10-$\\mu$m silicate feature \\citep{wh00,ww04,bl09} and nearby continuum \\citep{la08,mf09} (\\S\\ref{sec:variable})? The shorter variation timescales are substantially faster than the orbital periods at the AU radii where the feature forms. \\end{enumerate} A summary, discussion and conclusions are in \\S\\ref{sec:conc}. ", "conclusions": "} We have made resistive MHD calculations of a patch of disk in orbit around a young star, treating simultaneously for the first time three processes that together regulate the distribution of magneto-rotational turbulence. First, the dust drifts through the gas under the gravity and gas drag forces. Second, the dust alters the resistivity of the gas, and third, the resistivity controls the evolution of the magnetic fields driving the gas flows. The three processes combine in quite different ways in the disk atmosphere and interior. In the atmosphere, the high flux of ionizing radiation from the star allows good coupling between gas and magnetic fields even when the dust abundance is high. Intermittency in the turbulence and in the distribution of dust result from fluctuations in the magnetic field strength. During intervals of weak fields, the magnetic forces drive turbulence that carries dust particles high in the atmosphere. The turbulent mixing of the grains is faster than either the settling or the outflows resulting from the open boundaries. When the fields are stronger, their large pressure prevents magneto-rotational turbulence and reduces the gas density, so that the particles settle quickly. At the same time in the dead zone near the disk midplane, recombination on grain surfaces reduces the ionization fraction below the level needed to sustain turbulence. Grains settle at the laminar rate, with superimposed vertical oscillations due to the propagation of waves excited in the turbulent layers. Particles falling into the laminar layer do not return to the disk atmosphere and will eventually concentrate near the midplane, in a layer whose thickness is not less than the minimum imposed by the vertical oscillations. A dead zone enriched in solids may provide a favorable environment for the growth of larger bodies. Many T~Tauri systems show a combination of 10-$\\mu$m silicate emission and shallow millimeter spectral slope that is consistent with micron-sized grains in the disk atmosphere and millimeter-sized or larger particles in the interior, though it should be noted that the two wavelengths probe different distances from the star \\citep{dc06,pp08}. The turbulent and dead zones exchange dust and gas when turbulent motions overshoot the mutual boundary. The mixing extends about one scale height into the dead zone, as shown by the penetration of the dust-enriched material in figures~\\ref{fig:xgvsz1um} to~\\ref{fig:xgvsz100um}. The motions can perhaps be described in terms of Alfv\\'en waves driven by the breakup of channel flows in the disk atmosphere \\citep{si09}. The net effect is a transfer of dust to the dead zone, because settling proceeds fastest in the disk atmosphere. During the early stages of the concentration of the solid material, turbulence thus speeds up settling. The turbulent layer will eventually become dust-free unless replenished through radial transport within the disk or rain-out from an outflow or from the circumstellar envelope. Well-mixed micron or sub-micron grains with the interstellar mass fraction produce a dead zone reaching more than two scale heights from the midplane at 5~AU in the minimum-mass solar nebula. However, grains 10~$\\mu$m or larger yield a dead zone ending below one scale height. Turbulent mixing then penetrates to the midplane, consistent with the finding by \\cite{fp06} that diffusion adequately describes the effects of the gas motions on particles in a dead zone with a small vertical extent. While we were able to run the case with 1-$\\mu$m grains only about 60~orbits, the continuing loss of dust into the dead zone without resupply will ultimately yield a depletion factor in the surface layers exceeding the value of about ten needed for mixing to reach the midplane. The rate at which mass flows through the patch of disk toward the star is the product of the turbulent layer depth with the mean accretion stress. Both these quantities increase with the grain size, the ionizing flux from the star, and the magnetic flux threading the disk. Depletion of the dust in the surface layers will have a similar effect to an increase in the grain size. The two-decade spread in accretion rate among T~Tauri stars cannot easily be produced by grain growth and the observed spread in stellar X-ray output, but can be understood if in addition the objects have a modest range of disk magnetic field strengths. The mass accretion rate resulting from the magnetic stresses will vary with radius, leading to the build-up of material in some annuli \\citep{al01,zh09}, an effect not captured in our shearing-box calculations. Apparently the stellar X-rays both increase the disk accretion rate within 5~AU (figure~\\ref{fig:activeht}) and increase the rate at which material escapes from the disk into a photoevaporating wind at 10-40~AU \\citep{de09}. Furthermore the ionization of the inner disk could contribute to a wind driven by magnetic activity \\citep{si09}. All these effects tend to reduce the surface density of gas in the planet-forming region faster around stars with high X-ray luminosities. Whether it is possible in this way to account for the tendency of non-accreting T~Tauri stars to have higher X-ray luminosities \\citep{ns95,sn01,fd03,pk05,tg07} or for an inverse correlation between X-ray luminosity and stellar accretion rate \\citep{de09} remains unclear. We suggest that the diversity in the mid-infrared silicate bands among classical T~Tauri stars is a consequence of the concentration of solid material associated with planet formation. The growth of larger bodies requires a loss of dust, leading generally to a declining recombination cross-section. The smallest grains provide most of the recombination when the particles have a size spectrum appropriate for the interstellar medium \\citep{sm00}, and will dominate still more in the disk atmosphere, where settling removes the larger grains. The disruption of aggregates through collisions (\\S\\ref{sec:stirring}) thus slows the decline. However, unless the grains making up the aggregates also fragment so that the minimum particle size becomes smaller, the loss of cross-section is likely to continue, since even 1-$\\mu$m particles show appreciable settling in the turbulent disk atmosphere. The decreasing recombination rates will allow the active layers to expand into the disk interior, and if the mass column is not too great, turbulent mixing will eventually reach the midplane, returning the second-generation debris from planet formation to the atmosphere. Questions for future investigation include when and where the mixing reaches the midplane, and whether the observed correlation between the degree of settling and the crystalline mass fraction \\citep{wl09} can be explained by the return to the atmosphere of minerals formed near the midplane at high temperature and pressure in planetesimal interiors, impacts, or shocks. Any such correlation will be blurred by the fact that the degree of dust depletion required for mixing to reach the midplane depends on the column of gas, the X-ray luminosity of the star and the magnetic flux delivered to the disk from the surroundings. The mid-infrared variability of protostellar disks has been proposed to arise from the changing illumination of dust in the disk atmosphere. Explaining timescales shorter than a week is a challenge, since the dust in the disks around young Solar-mass stars extends inward only to about 0.1~AU \\citep{mc03,ab05,a08}, where the orbital period is 12~days. The results of the MHD calculations demonstrate that magnetic activity moves dust from place to place in the disk atmosphere on several timescales. The optical depth along rays grazing the surface of the patch of disk frequently varies by a factor two within a tenth of an orbit as regions of enhanced dust density pass by. Over longer time intervals of a few orbits, the photosphere moves up and down by as much as a scale height due to changes in the balance between settling and stirring. Clearly the effects of the magnetic activity on the dust distribution should be considered when attempting to understand the infrared variability. Properly accounting for the shadows cast on the disk surface requires models of the inner edge of the dust distribution, perhaps informed by the time-steady picture that has been developed for Herbig~Ae stars \\citep{dd01,in05}. Our results suggest that magneto-rotational turbulence near the dust sublimation radius can cast shadows that vary over timescales as short as a day." }, "0911/0911.3536_arXiv.txt": { "abstract": "It is widely believed that the cosmological redshift is not a Doppler shift. However, Bunn \\& Hogg have recently pointed out that to settle properly this problem, one has to transport parallely the velocity four-vector of a distant galaxy to the observer's position. Performing such a transport along the null geodesic of photons arriving from the galaxy, they found that the cosmological redshift is purely kinematic. Here we argue that one should rather transport the velocity four-vector along the geodesic connecting the points of intersection of the world-lines of the galaxy and the observer with the hypersurface of constant {\\em cosmic time}. We find that the resulting relation between the transported velocity and the redshift of arriving photons is {\\em not\\/} given by a relativistic Doppler formula. Instead, for small redshifts it coincides with the well known non-relativistic decomposition of the redshift into a Doppler (kinematic) component and a gravitational one. We perform such a decomposition for arbitrary large redshifts and derive a formula for the kinematic component of the cosmological redshift, valid for any FLRW cosmology. In particular, in a universe with $\\Omega_\\mathrm{m} = 0.24$ and $\\Omega_\\Lambda = 0.76$, a quasar at a redshift $6$, at the time of emission of photons reaching us today had the recession velocity $v = 0.997c$. This can be contrasted with $v = 0.96c$, had the redshift been entirely kinematic. Thus, for recession velocities of such high-redshift sources, the effect of deceleration of the early Universe clearly prevails over the effect of its relatively recent acceleration. Last but not least, we show that the so-called {\\em proper\\/} recession velocities of galaxies, commonly used in cosmology, are in fact radial components of the galaxies' four-velocity vectors. As such, they can indeed attain superluminal values, but should not be regarded as real velocities. ", "introduction": "\\label{sec:intro} A standard interpretation of the cosmological redshift in the framework of the Friedman-Lema{\\^\\i}tre-Robertson-Walker (FLRW) models is that it is an effect of the expansion of the Universe. This interpretation is obviously correct since $1 + z = a(\\tO)/a(\\tE)$, where $z$ is the value of the redshift, $a(t)$ is the scale factor of the Universe and $\\tE$ and $\\tO$ are respectively the times of emission and observation of a sent photon. In semi-popular literature (e.g. Kaufmann \\& Freedman 1999; Franknoi, Morrison \\& Wolff 2004; Seeds 2007), but also in professional (e.g. Harrison 2000; Abramowicz \\etal\\ 2007), one can often find statements that distant galaxies are `really' at rest and the observed redshift is caused by the `expansion of space'. According to other authors (e.g. Peacock 1999; Whiting 2004; Chodorowski 2007a,b; Bunn \\& Hogg 2009, hereafter BH9), such statements are misleading and cause misunderstandings about the cosmological expansion. A presentation of the ongoing debate in the literature on this issue is beyond the scope of the present work; a (possibly non-exhaustive) list of papers includes Davis \\& Lineweaver (2001), Davis \\& Lineweaver (2004), Barnes \\etal\\ (2006), Francis \\etal\\ (2007), Lewis \\etal\\ (2007), Lewis \\etal\\ (2008), Gr{\\o}n \\& Elgar{\\o}y (2007), Peacock (2008), Abramowicz \\etal\\ (2009), Chodorowski (2008), Cook \\& Burns (2009). On the other hand, there is broad agreement that the cosmological redshift is not a pure Doppler shift; the gravitational field must also generate a gravitational shift. Gravity can be neglected only locally, in the local inertial frame (LIF) of an observer. It turns out that for small redshifts, the cosmological redshift can be decomposed into a Doppler shift and a Newtonian gravitational one (Bondi 1947). The latter is a shift induced by the Newtonian gravitational potential. Can the cosmological redshift be decomposed into a Doppler shift and a gravitational shift (not necessarily Newtonian) for an {\\em arbitrary\\/} value of the redshift? This is the question which we want to deal with in this Paper. Formally, the answer is no. There is no invariant definition of the recession velocity of a distant galaxy in general relativity (GR). This velocity is a relative velocity of the galaxy and the observer, and in curved spacetime there is no unique way to compare vectors at widely separated points. A natural way to define the recession velocity is to {\\em transport parallely\\/} the velocity four-vector of the distant galaxy to the observer, but the result will depend on the chosen path. (This is just the definition of curvature.) In practice, however, as a `preferred' path one can choose a {\\em geodesic\\/} connecting the galaxy and the observer. Moreover, in FLRW models there is a natural foliation of spacetime, into space-like hypersurfaces of constant {\\em cosmic time}. In our Paper, as the geodesic we will adopt the geodesic lying on such a hypersurface, i.e.\\ connecting the points of intersection of the world-lines of the galaxy and the observer with the hypersurface of constant cosmic time. In a seminal study of the cosmological redshift, BH9, following Synge~(1960) and Narlikar~(1994), adopted another geodesic for the parallel transport: the null geodesic along which the photon is travelling from the source to the observer. This approach results in one `effective' velocity, while we think it is important to make a distinction between the velocity {\\em at the time of emission\\/} and the velocity {\\em at the time of observation}. These two velocities are obtained by transporting parallely the velocity four-vector of the source respectively on the hypersurface of constant $\\tE$ and $\\tO$. Not surprisingly, our result differs from that obtained by BH9. However, unlike theirs, ours correctly reproduces the small-redshift decomposition of the cosmological redshift into a Doppler component and a gravitational component, mentioned above. Specifically, our decomposition coincides with that of Bondi (1947) for isotropic {\\em and\\/} homogeneous matter distribution. This Paper is organized as follows. In Section~\\ref{sec:transport} we transport parallely the velocity four-vector of a distant galaxy to the observer. From the transported vector we calculate the recession velocity of the galaxy, which turns out to depend on the galaxy's comoving distance and the assumed background cosmological model. In Section~\\ref{sec:specific} we find specific relations between the cosmological redshift and its Dopplerian component for two particularly simple cosmological models: the empty model and the Einstein-de Sitter model. In Section~\\ref{sec:PoE}, using only the Principle of Equivalence, we derive the small-redshift decomposition of the cosmological redshift and find it to be identical with that obtained in Section~\\ref{sec:specific} using generally relativistic approach. In Section~\\ref{sec:rec} we calculate the recession velocities in the two models mentioned above, as well as for the currently favoured, flat non-zero $\\Lambda$ model with $\\Omega_{\\mathrm{m}} = 0.24$. We find also that the velocities are {\\em subluminal\\/} in {\\em all\\/} FLRW cosmological models and for {\\em all\\/} values of the redshift. A comparison of the results obtained in this Paper with some earlier works on the subject is given in Section~\\ref{sec:disc}. Summary is presented in Section~\\ref{sec:summ}. ", "conclusions": "recession velocities of all galaxies inside our particle horizon are subluminal. This fact is perhaps more fundamental than the specific value of recession velocity for a given redshift." }, "0911/0911.4120_arXiv.txt": { "abstract": "{Earlier work has suggested that large-scale dynamos can reach and maintain equipartition field strengths on a dynamical time scale only if magnetic helicity of the fluctuating field can be shed from the domain through open boundaries.}% {Our aim is to test this scenario in convection-driven dynamos by comparing results for open and closed boundary conditions.} {Three-dimensional numerical simulations of turbulent compressible convection with shear and rotation are used to study the effects of boundary conditions on the excitation and saturation of large-scale dynamos. Open (vertical-field) and closed (perfect-conductor) boundary conditions are used for the magnetic field. The shear flow is such that the contours of shear are vertical, crossing the outer surface, and are thus ideally suited for driving a shear-induced magnetic helicity flux.}% {We find that for given shear and rotation rate, the growth rate of the magnetic field is larger if open boundary conditions are used. The growth rate first increases for small magnetic Reynolds number, $\\Rm$, but then levels off at an approximately constant value for intermediate values of $\\Rm$. For large enough $\\Rm$, a small-scale dynamo is excited and the growth rate of the field in this regime increases as $\\Rm^{1/2}$. Regarding the nonlinear regime, the saturation level of the energy of the total magnetic field is independent of $\\Rm$ when open boundaries are used. In the case of perfect conductor boundaries, the saturation level first increases as a function of $\\Rm$, but then decreases proportional to $\\Rm^{-1}$ for $\\Rm\\ga 30$, indicative of catastrophic quenching. These results suggest that the shear-induced magnetic helicity flux is efficient in alleviating catastrophic quenching when open boundaries are used. The horizontally averaged mean field is still weakly decreasing as a function of $\\Rm$ even for open boundaries.}% {}% ", "introduction": "The classical view of the effect of shear in large-scale dynamos is that it promotes the generation of magnetic fields, and that it is instrumental in exciting oscillatory solutions (e.g.\\ Parker \\cite{P55}; Steenbeck \\& Krause \\cite{SK69}). While these aspects remain unchanged, a much more subtle effect of shear has been discovered in studies of the saturation mechanism of the large-scale dynamo. If the saturation of the dynamo occurs due to magnetic helicity conservation, a fully periodic or closed system experiences catastrophic quenching with the energy of the large-scale magnetic field decreasing in inverse proportion to the magnetic Reynolds number (Vainshtein \\& Cattaneo \\cite{VC92}; Cattaneo \\& Hughes \\cite{CH96}; Brandenburg \\cite{B01}). A way to avoid catastrophic quenching is to allow magnetic helicity fluxes to escape from the system through open boundaries, thus allowing the large-scale fields to saturate at equipartition strength in a dynamical time scale (Blackman \\& Field \\cite{BF00}; Kleeorin et al.\\ \\cite{KMRS00}). One of the most promising mechanisms, introduced by Vishniac \\& Cho (\\cite{VC01}), operates by driving a flux of magnetic helicity along the isocontours of shear. Forced turbulence simulations with shear have confirmed that the quenching of the $\\alpha$-effect is up to 30 times weaker when closed boundaries are replaced by open ones (Brandenburg \\& Sandin \\cite{BS04}). Recent numerical experiments with forced turbulence (Brandenburg \\cite{B05}; Yousef et al.\\ \\cite{YHSea08a,YHSea08b}; Brandenburg et al.\\ \\cite{BRRK08}) and convection (K\\\"apyl\\\"a et al.\\ \\cite{KKB08}, hereafter Paper~I; Hughes \\& Proctor \\cite{HP09}; see also K\\\"apyl\\\"a et al.\\ \\cite{KKB10}) with imposed large-scale shear flows have established the existence of a robust large-scale dynamo. The origin of this dynamo is still under debate (shear--current versus incoherent $\\alpha$--shear effects). This is discussed in detail elsewhere (e.g.\\ Brandenburg et al.\\ \\cite{BRRK08}; K\\\"apyl\\\"a et al.\\ \\cite{KKB09b}). In forced turbulence simulations with fully periodic boundaries, that do not allow magnetic helicity fluxes, large-scale magnetic fields show slow saturation on a diffusive timescale (Brandenburg \\cite{B01}). Earlier results from convection also suggest that no appreciable large-scale magnetic fields are seen if magnetic he\\-li\\-ci\\-ty fluxes are suppressed (Paper~I). In most of the simulations of Paper~I the isocontours of shear were vertical and thus intersecting the boundaries. However, the boundary conditions (bc) imposed on the magnetic field on the vertical borders would either allow (open vertical-field bc) or inhibit (closed perfect-conductor bc) a net flux out of the domain. It was shown that when the flux is suppressed, only weak large-scale magnetic fields occur in the saturated state at large $\\Rm$. Conversely, allowing a flux produced a significant large-scale field. Furthermore, using a similar setup as in Tobias et al.\\ (\\cite{TCB08}), where the isocontours of shear are horizontal, a similar effect was found. Thus it would appear that the shear-induced flux plays a critical role in exciting a large-scale dynamo in turbulent convection. This has given rise to the impression that the dynamos seen in Paper~I are solely due to the magnetic helicity flux and thus inherently nonlinear (Hughes \\& Proctor \\cite{HP09}). In the present paper we argue that this interpretation is incorrect and that the lack of appreciable large-scale magnetic fields in some of our earlier cases with vanishing helicity flux is due to the fact that the critical dynamo number for the excitation of the dynamo is larger in that case. We also note that, more recently, large-scale dynamos have been found in turbulent convection without shear (K\\\"apyl\\\"a et al.\\ \\cite{KKB09a}), irrespective of the presence of magnetic helicity fluxes. The origin of the dynamos reported in Paper~I was interpreted in the context of turbulent dynamo theory in K\\\"apyl\\\"a et al.\\ (\\cite{KKB09b}). An important feature of turbulent dynamos is that according to standard mean-field theory neither the turbulent transport coefficients nor the growth rate of the field should depend on the microscopic resistivity. The former prediction was confirmed by K\\\"apyl\\\"a et al.\\ (\\cite{KKB09b}), but the latter has not yet been studied in detail in direct simulations. This is one of the goals of the present paper. Another aim is to study the saturation behaviour of the large-scale dynamo with both open and closed boundary conditions. Firstly, in homogeneous systems, i.e.\\ where the turbulent transport coefficients do not vary in space, full saturation is expected to occur on a slow diffusive time scale with closed boundaries whereas with open boundaries the saturation is expected to happen on a dynamical time scale if a helicity flux is present (Brandenburg \\cite{B01}). However, in inhomogeneous systems the situation is likely to be more complex (e.g.\\ Mitra et al.\\ \\cite{MCCTB10}). Furthermore, simulations of helically forced isotropic turbulence have shown that, in the absence of shear, the saturation level of the magnetic field scales as $\\meanv{B}^2/\\Beq^2\\propto\\Rm^{-1}$ for open boundaries (Fig.~2 of Brandenburg \\& Subramanian \\cite{BS05a}). When shear is present, the shear-induced magnetic helicity flux should alleviate the $\\Rm$-dependence of the saturation level. Preliminary results in Paper~I indicate that this might indeed be true, but the results are not fully conclusive because other parameters, such as the Rayleigh number and the fluid Reynolds number changed when $\\Rm$ was varied. In the present paper we perform a more detailed study where only $\\Rm$ is varied. ", "conclusions": "We have studied the effects of magnetic boundary conditions on the excitation and saturation of large-scale dynamos driven by turbulent convection, shear, and rotation by means of numerical simulations. We find that the critical magnetic Reynolds number is greater ($\\Rm\\approx5$) in the case of closed boundaries (perfect-conductor) compared to simulations with open (vertical-field) boundaries where the large-scale dynamo is excited already for $\\Rm\\approx1$. A similar result is obtained from mean-field models operating in the same parameter regime. Furthermore, the relevant dynamo number is roughly three times greater with closed boundaries. These effects are likely to explain the weak large-scale dynamo action seen in the perfect-conductor runs in Paper~I. The measured growth rate of the magnetic field is independent of the microscopic resistivity when $\\Rm$ is sufficiently above the critical value. This is manifested by the approximately constant growth rate in the intermediate $\\Rm$ range in Fig.~\\ref{fig:pgr}. For $\\Rm\\ga30$, the small-scale dynamo is excited, and for $\\Rm\\ga60$ it becomes dominant. The growth rate of the magnetic field is then consistent with $\\lambda\\propto\\Rm^{1/2}$ scaling, which is in accordance with the results of Schekochihin et al.\\ (\\cite{Scheko04}) and Haugen et al.\\ (\\cite{{HBD04}}). The fact that the qualitative behaviour of the growth rate of the magnetic field is similar for both boundary conditions suggests that the origin of the large-scale dynamo is not likely to be a process that is essentially nonlinear, as suggested by Hughes \\& Proctor (\\cite{HP09}), but that it can be understood within the framework of classical kinematic mean-field theory. In the saturated state the energy of the total magnetic field remains independent of $\\Rm$ for open boundaries and decreases as $\\Rm^{-1}$ for closed boundaries. The latter result is consistent with catastrophic quenching while the former result suggests that magnetic helicity fluxes are efficiently driven out of the system. On the other hand, the energy of the mean field, taken here to be represented by horizontal averages, decreases approximately as $\\Rm^{-0.25}$ for open and as $\\Rm^{-1.6}$ for closed boundaries. It is not yet clear why the mean fields tend to show a steeper decline than the total field. It is possible that this declining trend levels off at a higher $\\Rm$ and that the magnetic Reynolds numbers in our simulations are still not large enough (cf.\\ Brandenburg et al.\\ \\cite{BCC09}). A similarly weak $\\Rm$-dependence has been observed in the cycle period of $\\alpha$-shear dynamos with isotropically forced turbulence (K\\\"apyl\\\"a \\& Brandenburg \\cite{KB09}). We find that, for intermediate values of $\\Rm$, the large-scale magnetic field saturates on a resistive time scale when closed boundaries are used. However, with the largest $\\Rm$ no clear signs of slow saturation are observed. Earlier results using perfect-conductor boundaries have shown that the mean field can evolve in steps (Brandenburg \\& Dobler \\cite{BD02}; see also Brandenburg et al.\\ \\cite{BKMMT07}) which are associated with a change of the large-scale magnetic field configuration. We have not seen such a behaviour in our current simulations but the existence of such events at a later stage cannot be ruled out. Another possible explanation is that there are magnetic helicity fluxes occurring inside the domain which arise from the spatial gradients of magnetic helicity (e.g.\\ Covas et al.\\ \\cite{CTTB98}; Kleeorin et al.\\ \\cite{KMRS00}; Mitra et al.\\ \\cite{MCCTB10}). However, a quantitative study of these effects requires more detailed knowledge of the helicity fluxes and possibly an anisotropic formulation of the magnetic $\\alpha$ effect. These issues merit further investigation and are beyond the scope of the present paper." }, "0911/0911.1369_arXiv.txt": { "abstract": "We investigate a new theory of the origin of the irregular satellites of the giant planets: capture of one member of a $\\sim$100-km binary asteroid after tidal disruption. The energy loss from disruption is sufficient for capture, but it cannot deliver the bodies directly to the observed orbits of the irregular satellites. Instead, the long-lived capture orbits subsequently evolve inward due to interactions with a tenuous circumplanetary gas disk. We focus on the capture by Jupiter, which, due to its large mass, provides the most stringent test of our model. We investigate the possible fates of disrupted bodies, the differences between prograde and retrograde captures, and the effects of Callisto on captured objects. We make an impulse approximation and discuss how it allows us to generalize capture results from equal-mass binaries to binaries with arbitrary mass ratios. We find that at Jupiter, binaries offer an increase of a factor of $\\sim$10 in the capture rate of 100-km objects as compared to single bodies, for objects separated by tens of radii that approach the planet on relatively low-energy trajectories. These bodies are at risk of collision with Callisto, but may be preserved by gas drag if their pericenters are raised quickly enough. We conclude that our mechanism is as capable of producing large irregular satellites as previous suggestions, and it avoids several problems faced by alternative models. ", "introduction": "\\label{intro} \\subsection{Previously suggested capture models} \\label{models} With discoveries accelerating in the last decade, we now know of over 150 satellites orbiting the giant planets. About one-third of these are classified as regular, with nearly circular and planar orbits. It is thought that these satellites are formed by accretion in circumplanetary disks. The majority of the satellites, however, are irregular and follow distant, highly eccentric and inclined paths. It is widely believed that irregular satellites originated in heliocentric orbits and were later captured by their planets, but the details of how this occurred are still uncertain. At least seven different models have been proposed, involving dissipative forces, collisions, resonances, and three-body effects. Each model has its own strengths and weaknesses. In one long-standing theory, planetesimals are slowed as they punch through the gas disk surrounding a young, growing planet (Pollack \\etall, 1979). For this mechanism to be efficient, the gas must be sufficiently dense to capture the planetesimals in one pass. This is problematic, however, because if the gas disk does not rarefy substantially in $\\sim$100-1000 years, the orbits of the new satellites will decay inward, leading to collisions with the planet and its regular satellites. Furthermore, the atmospheres of Uranus and Neptune have only a few Earth-masses of hydrogen and helium at present, so their gas disks could not have been as extensive or long-lived as those of Jupiter and Saturn. A likely outcome of this model, then, is that satellite capture should have been different at Jupiter and Saturn than at Uranus and Neptune; however, current observational estimates suggest equal efficiencies (Jewitt \\& Sheppard, 2005). With a model similar to that of Pollack \\etall, {\\'C}uk \\& Burns (2004a) found that Jupiter's largest irregular satellite, Himalia, would evolve inward to its current orbit in $10^4-10^6$ years. This tenuous gas, however, may make capture difficult. In another model, planetesimals are captured when the mass of the planet increases (Heppenheimer \\& Porco, 1977). This mass growth causes the planet's escape velocity to increase, rendering a previously free planetesimal bound to the planet. For this method to be effective, the planet's mass must increase substantially on $\\sim$100-1000-year timescales. However, in most planet formation models (e.g. Pollack \\etall, 1996), giant planet growth is hypothesized to take place on timescales many orders of magnitude longer than required by this capture scenario. Furthermore, Uranus and Neptune's gas deficiency implies that their growth was of very short duration. Thus, our current understanding of planetary formation makes this model improbable. The observation that the four giant planets contain approximately the same number of irregular satellites (accounting for observational biases; Jewitt \\& Sheppard, 2005) has led to a renewal of interest in capture theories that do not depend strongly on the planet's formation process. In one such scenario, a planetesimal collides with a current satellite or another planetesimal in the vicinity of the planet, resulting in its capture (Colombo \\& Franklin, 1971). Though collisions were certainly more common in the early Solar System than they are today, if they resulted in enough energy loss to permit capture, they would likely also have catastrophically disrupted the bodies. Nevertheless, the fragments might then have become independent satellites. A fourth suggestion involves the possible instability in the orbits of the outer planets early in the Solar System (e.g., by a 2:1 resonance crossing between Jupiter and Saturn). Outlining the theory of the Nice model of Solar System evolution, Tsiganis \\etal (2005) have shown that such an event could cause Uranus and Neptune to have many close approaches with each other and with Jupiter and Saturn. During these encounters, the influence of the massive interloping planet can cause planetesimals to be stabilized as satellites (Nesvorn{\\'y} \\etall, 2007). This method is promising but has an important disadvantage in that Jupiter (and Saturn, to a lesser extent) sustains very few close encounters relative to the ice giants. Thus the gas giants are inefficient at capturing satellites in this way (Nesvorn{\\'y} \\etall, 2007). Astakhov \\etal (2003) examined low-energy orbits that linger near Jupiter and Saturn. While these bodies are not permanently captured, the authors found that some of them were stable for thousands of years, long enough to allow a weak dissipative force such as gas drag to complete the capture process. However, the overall percentage of temporary captures that do not escape is small, and many of these bodies are threatened by collision with the planets' large outer satellites (e.g., Callisto and Titan). Agnor \\& Hamilton (2006a) examined the capture of Triton from an exchange reaction between a binary pair and Neptune. Their motivation stemmed from the newly-discovered abundance of binaries in small-body populations. Currently, it is estimated that binaries account for $\\sim$30\\% of Kuiper belt objects (KBOs) with inclinations $<$ 5$\\deg$, $\\sim$5\\% of the rest of the KBOs (Noll \\etall, 2008), and $\\sim$2\\% of large main belt asteroids (diameters $>$ 20 km; percentage increases for smaller objects; Merline \\etall, 2007). In Agnor \\& Hamilton's capture model, a binary is tidally disrupted and one of its members, Triton, is captured as a satellite. This process is most effective for large satellites like Triton, with radius 1350 km. However, the largest of the other irregular satellites are more than 10 times smaller than Triton: Himalia at Jupiter is $\\sim$85 km in radius, Saturn's largest irregular, Phoebe, is $\\sim$110 km, Uranus's Sycorax is $\\sim$80 km, and Neptune's Halimede and Neso are only $\\sim$30 km each. Capturing these satellites via binary exchange reactions would be significantly more difficult, as we will discuss further below. Finally, Vokrouhlick{\\'y} \\etal (2008) examined binary exchange reactions during the first 100 Myr after an assumed Jupiter/Saturn 2:1 resonance crossing, using results of the Nice model (Tsiganis \\etall, 2005) to guide their initial conditions. Because planetesimal speeds relative to the planets are high after the scattering phase of the Nice model, they found that captures from binaries during that time do not match current orbital parameters and occur too infrequently to account for today's populations. \\subsection{Our model: Capture from 100-km binaries} \\label{ourmodel} All of the above models have promising aspects coupled with important limitations. In this work, we seek to combine the best features of several models into a viable capture scenario. In particular, we examine binaries (as in Agnor \\& Hamilton, 2006a and Vokrouhlick{\\'y} \\etall, 2008) as a way to augment capture from low-velocity orbits resulting from three-body interactions like those studied by Astakhov \\etal (2003). While Vokrouhlick{\\'y} \\etal (2008) studied exchange reactions in the context of an assumed initial planetesimal population, we focus on assessing the viability of the mechanism itself. Our goal is to determine how various parameters of the model affect its plausibility. We examine its viability at Jupiter, as a number of the above models suggest that capturing at the largest gas giant is especially difficult. As the largest of the existing irregular satellites are $\\sim$80-110 km, capture of objects in this size-range is particularly interesting. Since it is likely that the irregular satellite population contains collisional families (Nesvorn{\\'y} \\etall, 2003; Sheppard \\& Jewitt, 2003), it may be the case that only the largest objects were captured, while the smaller satellites formed later, via collisions. For this reason, we focus our investigation on capturing the $\\sim$100-km progenitors. In order to stabilize and shrink the resulting capture orbits, a dissipation source is required; we suggest a tenuous version of the gas drag originally proposed by Pollack \\etal (1979). Two of Jupiter's irregular satellites, Pasiphae and Sinope, as well as Saturn's satellite, Siarnaq, and Uranus' Stephano, are found in resonances that seem to require just such a weak dissipative force (Whipple \\& Shelus, 1993; Saha \\& Tremaine, 1993; {\\'C}uk \\& Burns, 2004b). Furthermore, a tenuous circumplanetary disk is consistent with current theories of late-stage planetary formation. Jupiter's massive gaseous envelope of hydrogen and helium necessitates that it formed in the Solar System's circumstellar gas disk. Before the end of its accretion, Jupiter was likely able to open a gap in the local density distribution of the gas (for a review, see e.g. Papaloizou \\etall, 2008). After gap opening, gas continues to leak into the planet's Hill sphere through the $L_1$ and $L_2$ points, but at a rate much reduced in comparison to the previous epochs. A tenuous circumplanetary gas disk results (e.g., Lubow \\etall, 1999; D'Angelo \\etall, 2003; Bate \\etall, 2003), from which material may condense and regular satellites may accrete near the planet (e.g., Canup \\& Ward, 2002; Mosquiera \\& Estrada, 2003). In {\\'C}uk \\& Burns' study (2004a) of the Himalia progenitor's orbital evolution, they considered circumjovian nebular conditions consistent with hydrodynamical simulations of Jupiter's gap opening in a circumstellar gas disk (e.g., Lubow \\etall, 1999) and found that the post-capture timescale for evolving this progenitor to its present orbit to be roughly in the range of $10^4-10^6$ years. This is similar to the timescale in which extrasolar circumstellar disks transition from optically thick to thin ($\\sim10^5$ years; Skrutskie \\etall, 1990; Silverstone \\etall, 2006; Cieza \\etall, 2007). The similarity of timescales suggests that satellites captured at the onset of disk dispersal have a good chance of experiencing stabilizing orbital evolution while also avoiding collision with the planet. The timescale for binary capture is very short compared to evolution timescales from a tenuous gas disk. Therefore, we focus our study first on characterizing the effectiveness of binary capture in the absence of gas. In the following sections, we critically evaluate our model for capturing irregular satellites from low-mass ($\\sim$100-km) binaries. We begin with a closer examination of the three-body capture process and then explore parameter space with a large suite of numerical simulations. We then discuss the ability of gas drag to stabilize post-capture orbits in Section~\\ref{survivability}. ", "conclusions": "The new model discussed here, capture from low-mass binaries with subsequent orbital evolution, has both significant advantages and disadvantages in comparison to other suggested models. One important advantage is that capture is viable at Jupiter and Saturn, unlike three of the models discussed in Section~\\ref{intro}. For example, Agnor \\& Hamilton's (2006) direct three-body capture model works only for very large bodies in the gas giants' high-approach-speed environments. Also, the theory of Nesvorn{\\'y} \\etal (2003) requires close approaches among the giant planets, of which Jupiter has very few in the Nice model scenario. Saturn encounters its outer neighbors more frequently than Jupiter, but it still suffers from few close approaches overall in this model. In Vokrouhlick{\\'y} \\etal (2007), the capture statistics from planet-binary encounters are low for all planets, but especially Jupiter and Saturn. This is primarily because of the high relative velocities assumed in their model. Though these latter two capture mechanisms in their current forms cannot explain the gas giants' irregular satellites, they are worth further study, perhaps in the context of altered versions of the Nice model or other early Solar System models. Our model also has an important advantage over that of Pollack \\etall, 1979, in that our capture scenario allows for a much weaker gas, since the gas here is needed for shrinking the orbits, not capturing satellites. Also, the gas in our model can persist for much longer than that in Pollack \\etall, as a weaker gas does not have the problem of quickly destroying the captured bodies. Furthermore, various groups (e.g., Canup \\& Ward, 2002) have proposed that the Galilean satellites formed from a tenuous gas; if true, the gas was most likely even thinner when the irregular satellites were captured. Thus it is important that our capture mechanism does not rely on a dense gas. To directly compare our mechanism with that of Astakhov \\etal (2003), we need to consider both relative capture rates (and their survivability) and the prevalence of binaries vs. single bodies. In our simulations, we find that binary capture can provide a significant advantage over capturing from populations of single bodies for binaries with particular characteristics: high enough masses ($\\gtrsim$ 100 km), optimal separations ($\\sim$10-20 $R_B$), and low incoming energies (corresponding to Jacobi constants $\\gtrsim$ 3.02 and mostly prograde encounters). However, like Astakhov \\etal (2003), we also find that the probability of captures (from either binaries or single bodies) of avoiding collisions with Callisto is low for readily captured progenitors. In Section~\\ref{survivability}, we discussed that this problem can be alleviated by altering the capture orbits with the surrounding gas or by capturing binaries with larger-than-optimal separations that do not lead to Callisto-crossing orbits. So, how common were easily-captured binaries early in Solar System history? This question is difficult to assess. Observational surveys of the current population of the cold, classical Kuiper belt (i.e. objects with modest inclinations and eccentricities) find a 30\\% binary fraction among bodies larger than 100 km (Noll \\etall, 2008), many with nearly equal mass components. Also, recent studies of planetesimal formation have suggested that large, $\\gtrsim$ 100-km bodies may form quickly in the gaseous proto-planetary disk, providing the building blocks of subsequent planet formation (Johansen \\etall, 2007; Cuzzi \\etall, 2008). Binary formation is likely to be contemporaneous with the formation of these bodies (Nesvorn{\\'y}, 2008). Further, Morbidelli \\etal (2009) have shown that the size-frequency distribution of asteroids in the main belt is consistent with large initial planetesimals, at least $\\sim$100-km in size. Together these results indicate that the $\\sim$100-km binary objects considered here may have been quite common as the very last portions of the Solar System's gas disk were being depleted. Finally, the known irregular satellite population numbers $<$ 100, and many of these are probably members of families -- thus, we need only produce at most a few dozen captures. While accounting for the origin of such a small population is difficult, our simulations show that it is likely binaries played a role. We conclude by offering our model as a new idea that alleviates many but not all of the problems faced by previous models, but acknowledge that the without a detailed understanding of the initial population including binary statistics, a firm conclusion is not possible. \\newpage" }, "0911/0911.5476_arXiv.txt": { "abstract": "\\noindent We present new brightness and magnetic images of the weak-line T~Tauri star V410~Tau, made using data from the NARVAL spectropolarimeter at T\\'{e}lescope Bernard Lyot (TBL). The brightness image shows a large polar spot and significant spot coverage at lower latitudes. The magnetic maps show a field that is predominantly dipolar and non-axisymmetric with a strong azimuthal component. The field is 50\\% poloidal and 50\\% toroidal, and there is very little differential rotation apparent from the magnetic images. A photometric monitoring campaign on this star has previously revealed V--band variability of up to 0.6 magnitudes but in 2009 the lightcurve is much flatter. The Doppler image presented here is consistent with this low variability. Calculating the flux predicted by the mapped spot distribution gives an peak-to-peak variability of 0.04 magnitudes. The reduction in the amplitude of the lightcurve, compared with previous observations, appears to be related to a change in the distribution of the spots, rather than the number or area. This paper is the first from a Zeeman-Doppler imaging campaign being carried out on V410~Tau between 2009--2012 at TBL. During this time it is expected that the lightcurve will return to a high amplitude state, allowing us to ascertain whether the photometric changes are accompanied by a change in the magnetic field topology. ", "introduction": "T~Tauri stars (TTS) are pre-main sequence late-type stars. As they contract towards the main sequence the increasing density in their interiors will lead to the development of a radiative core, if the star is sufficiently massive. A point of uncertainty concerns the evolution of their dynamos. In fully convective stars, or stars with small radiative cores, we expect the dynamo to look fundamentally different to that of the Sun. We presume that the solar dynamo is concentrated in the boundary between the convective and radiative layers, known as the tachocline. As the radiative core in the young star grows the dynamo should change from being distributed throughout the convective zone to being confined to the tachocline. Observations of young stars with a range of ages, masses and rotation rates will help to illuminate this process. V410~Tau is one of the most well-observed TTS. It is a weak-line T~Tauri star (wTTS) which has already dissipated its disk, hence it does not display signatures of accretion characteristic of classical TTS (cTTS). It has a \\vsini\\ of $74\\pm3$\\kms\\ and a period of 1.872~days \\citep{stelzer03}, and is therefore an example of the young, rapidly rotating stars which are characterised by strong, large scale magnetic fields. Other, more evolved, examples include the 10~Myr old TWA~6 \\citep{skelly1}, Speedy~Mic \\citep[20 Myr,][]{barnes01,barnes05c, duns06} and AB~Dor \\citep[50~Myr,][]{cc02b,donati03b,jeffers07}. The strong fields in these stars have either been directly measured using Zeeman-Doppler imaging \\citep[ZDI,][]{semel89,jf97,tausco06}, or their presence indicated indirectly e.g. by the large spot coverage on the surface. In the case of V410~Tau this field has revealed itself in the presence of cool spots covering a large fraction of the photosphere. In fact, V410~Tau has exhibited one of the highest optical variabilities observed on a young star, with $\\rm \\Delta V$ reaching 0.6 magnitudes \\citep{stelzer03,grankin08}. V410~Tau is the target of an ongoing photometric monitoring campaign which began in 1981 \\citep{vrba88, herbst89,petrov94, grankin08}. For most of this time the lightcurve has been smooth, repeating, with a clearly defined period and an amplitude between 0.2 and 0.6 magnitudes. Two significant decreases in the amplitude of the lightcurve, where $\\Delta \\rm V$ fell below 0.2, have been observed during this campaign. The first of these was between 1981--1983 and the second began in 2005, reaching a minimum in 2007. Additional observations of V410~Tau dating from 1905--1987 are available on photographic plates in the archives of the Sternberg Astronomy Institute. \\cite{sokoloff08} have digitised this data and used it to investigate the variability over a longer period of time. This revealed a third decrease in lightcurve amplitude between 1963--1970, this time in the $B$--band. These changes in the lightcurves suggest a rearrangement of the cool spots. Given that starspots are a magnetic phenomenon such a change is likely to be caused by a change in the magnetic topology. The fact that this change has been observed three times with intervals of approximately 20~years may be evidence of a stellar cycle, similar to the 22-year solar cycle. The existence of magnetic cycles in active stars has been extensively investigated through changes in their starspot coverage \\cite[see, e.g., ][and references therein]{berd05, strass09}. Possible short-period cycles (of between 5--7 years) have been found on V410~Tau by \\cite{stelzer03} and \\cite{olah09}, using photometric data spanning several decades. Doppler imaging \\citep{vogt87,strass02} can provide more detail on the changes in spot distribution. Maps of the spot distribution on V410~Tau have been made by \\cite{v41094b, hatzes95b} and \\cite{v41096}. The images have shown decentred polar spots and low latitude spots. Often the lower latitudes features are found to be confined within particular longitude ranges \\citep[e.g.][]{v41096}. The flattening of the lightcurve suggests that a change in the distribution or number of photospheric spots has occurred, i.e., either there are fewer spots than previously or there is a more even longitudinal distribution of spots. \\cite{grankin08} found a stronger correlation between the non-uniformity of the spot distribution and the amplitude of the lightcurve than the total spot area. This suggests that the reduction in lightcurve amplitude is related to a rearrangement in the spot distribution, rather than a change in the number or area of the spots. To date, no magnetic images of V410~Tau have been published. This has motivated a long--term campaign, being carried out at T\\'{e}lescope Bernard Lyot (TBL), which aims to produce several spot and magnetic maps of V410~Tau between 2009--2012. We expect that during this time the lightcurve will return to the state observed between 1988--2004. In this paper we present the first set of images. Sec.~2 discusses the evolutionary status of V410~Tau. Sec.~3 describes the observations and data reduction; in Sec.~4 we describe the image production and present the brightness and magnetic images, and a new measurement of the differential rotation. Sec.~5 is a discussion of the implication of the results for the dynamo theory in pre-main sequence stars. ", "conclusions": "The Doppler image of the weak-line T~Tauri star V410~Tau in January 2009 shows a large polar spot and a several lower latitude spots spread around the latitude band between $0 - 30^{\\circ}$. This spot distribution can explain the relatively flat V--band lightcurve currently being observed on the star as, unlike during previous Doppler imaging campaigns, the spots are not confined to a particular longitude range. The polar spot on the image is centred on the pole, unlike on previous images where it was shifted from the pole. The \\ha\\ profile is complex and highly variable. Fast-moving absorption transients in the line may indicate the presence of high-lying circumstellar material, such as slingshot prominences. The magnetic field is predominately dipolar and non-axisymmetric. It has a complex radial field as well as a strong azimuthal field with large monolithic regions. The field resembles zero-age main sequence stars such as AB~Dor, but unlike that star it has very weak differential rotation. The rotation rate and depth of convection zone appear to be the most important parameters in determining the strength and configuration of the magnetic field Future Zeeman Doppler images planned as part of a campaign on T\\'{e}lescope Bernard Lyot (TBL) will reveal whether changes in the amplitude of the optical lightcurve are accompanied by significant changes in the spot distribution or magnetic field topology." }, "0911/0911.0779_arXiv.txt": { "abstract": "{\\bf star formation; cluster formation} Stellar clusters are born in cold and dusty molecular clouds and the youngest clusters are embedded to various degrees in dusty dark molecular material. Such embedded clusters can be considered protocluster systems. The most deeply buried examples are so heavily obscured by dust that they are only visible at infrared wavelengths. These embedded protoclusters constitute the nearest laboratories for direct astronomical investigation of the physical processes of cluster formation and early evolution. I review the present state of empirical knowledge concerning embedded cluster systems and discuss the implications for understanding their formation and subsequent evolution to produce bound stellar clusters. ", "introduction": "The question of the origin of stellar clusters is an old one. As early as 1785, William Herschel first considered this problem in the pages of these {\\it Transactions} when he speculated on the origin of the clusters Messier 80 and Messier 4 in Ophiuchus. More than two centuries later, despite profound advances in astronomical science, we find that the question of the physical origin of stellar clusters remains largely a mystery. This is, in part, due to the fact that cluster formation is a complex physical process that is intimately linked to the process of star formation, for which there is as yet no complete theory. The development of a theoretical understanding of star and cluster formation therefore depends on first acquiring detailed empirical knowledge of these phenomena. This can be a very difficult task. Consider, for example, the globular clusters, the most massive stellar clusters in our own Galaxy, the Milky Way. These stellar systems are more than 12 billion years old and are no longer formed in the Milky Way. Consequently, direct empirical study of their formation process is not possible. The situation is considerably better for Galactic open clusters. These typically span ages from 1 Myr to 1 Gyr and thus must be continually forming in the Milky Way, making direct observational study of their formation process possible, at least in principle. However, such studies have been seriously hindered by the fact that open clusters are born in molecular clouds and, during their formation and early evolution, are completely embedded in molecular gas and dust. They are thus obscured from view at optical wavelengths, where the traditional astronomical techniques are most effective. Fortunately, molecular clouds are considerably less opaque at infrared wavelengths and the development and deployment of infrared imaging cameras on optical- and infrared-optimized telescopes during the past two decades has provided astronomers with the ability to detect, survey and systematically study the extremely young embedded stellar clusters within nearby molecular clouds. Such studies indicated that embedded clusters are quite numerous and that they account for a significant fraction, if not the majority, of all star formation presently taking place in the Galaxy. Similarly, as discussed in accompanying articles by de Grijs (2010) and Larsen (2010), young clusters also appear to account for a significant fraction of star formation in other galaxies, such as in the very active and luminous starburst galaxies and in the vigorous star-formation episodes which accompany galaxy mergers and close interactions. In our Galaxy, the embedded-cluster phase lasts only 2--4 Myr and the vast majority of embedded clusters which form in molecular clouds dissolve within 10 Myr or less of their birth. High cluster infant mortality has also been inferred for clusters in other galaxies (see, e.g., de Grijs 2010; Larsen 2010). This early mortality of embedded clusters is likely a result of the low star-formation efficiency that characterizes the massive molecular-cloud cores within which the clusters form. The physical reason for the low formation efficiencies is not well understood and the origin of the massive cores themselves is one of the many mysteries of modern astrophysical research. The embedded clusters are the primary laboratory for research into the question of the physical origin of stellar clusters. At the present time, the answer to this question is shrouded by the dusty veils of the dark molecular clouds in which stars and clusters are born. The ultimate goal of a theoretical understanding of the physical process of cluster formation is far from being realized. The first step towards achieving this goal is to construct a solid empirical foundation upon which a physical theory can eventually be built. This then requires a detailed understanding of the physical nature of embedded clusters and the environments in which they form. In this article, I will review the empirical knowledge that is setting the stage for the eventual development of a theory of cluster formation. I will describe existing knowledge concerning many of the basic physical properties of the embedded-cluster population, including the cluster mass function, cluster birthrate, structural properties, etc., as well as the nature of the gaseous cloud cores in which they form. I will (at the end) briefly speculate on the implications of these results for understanding the origin of stellar clusters. A more theoretical discussion of possible physical mechanisms for forming such clusters is contained in the accompanying article by Clarke (2010). \\begin{figure} \\vspace{-3.5cm} \\begin{center} \\includegraphics[scale=0.55, angle=90]{lada_rs_f1.ps} \\end{center} \\vskip -1.0in \\caption{{\\it (left)} Optical and {\\it (right)} infrared view of an embedded Trapezium cluster associated with the Great Orion Nebula. North is on the left, west at the top. (From Lada \\& Lada 2003.)} \\end{figure} ", "conclusions": "More than two centuries after Herschel (1785) first considered the question, we now understand that clusters are formed in cold, dark molecular clouds. Stellar clusters begin their lives deeply buried in dense molecular gas and dust as embedded infrared protoclusters. Over the past two decades, advances in detector and telescope technology have led to steady advances in our knowledge of embedded protoclusters. These extremely young stellar systems appear to account for a significant fraction of all star formation presently taking place in the Milky Way and, thus, they are tracers of the current epoch of star formation in the Galaxy. Although considerable progress has been made towards understanding the basic physical properties of embedded clusters, a physical theory of cluster formation eludes us. In the Milky Way, the primary mode of cluster formation at the present epoch appears to be one in which relatively compact, massive dense cores in GMCs undergo a process of Jeans-like fragmentation that transforms cold, dense interstellar material into stellar form. Triggering of clouds by shocks or other mechanisms that lead to an increase in external pressure may represent an additional mode of Galactic cluster formation. Although its relative contribution to present-day cluster formation in the Milky Way is unclear, this latter mechanism may be significant in interacting galaxies and nuclear starbursts, and perhaps even in the early history of the Milky Way itself. It is now clear that to develop a predictive theory of cluster formation will require both a better understanding of the process of star formation and a comprehensive understanding of the physical mechanisms that organize the dense material of a GMC into massive, compact cores which, for a typical GMC in the Milky Way, occupy only a small fraction ($< 1$\\%) of its volume and account for only a small fraction (1--10\\%) of its total mass." }, "0911/0911.2310_arXiv.txt": { "abstract": "We investigate the global dynamics of the universe within the framework of the Interacting Dark Matter (IDM) scenario. Considering that the dark matter obeys the collisional Boltzmann equation, we can obtain analytical solutions of the global density evolution, which can accommodate an accelerated expansion, equivalent to either the {\\em quintessence} or the standard $\\Lambda$ models. This is possible if there is a disequilibrium between the DM particle creation and annihilation processes with the former process dominating, which creates an effective source term with negative pressure. Comparing the predicted Hubble expansion of one of the IDM models (the simplest) with observational data, we find that the effective annihilation term is quite small, as suggested by various experiments. ", "introduction": "The detailed analysis of the available high quality cosmological observations (\\cite{Spergel07,essence,Kowal08,komatsu08} and references therein) have converged during the last decade towards a cosmic expansion history that involves a spatial flat geometry and a recent accelerating expansion of the universe. This expansion has been attributed to an energy component (the so called dark energy) with negative pressure which dominates the universe at late times and causes the observed accelerating expansion. The nature of the dark energy is still a mystery and it is one of the most fundamental current problems in physics and cosmology. Indeed, due to the absence of a physically well-established fundamental theory, there have been many theoretical speculations regarding the nature of the above exotic dark energy (DE) among which a cosmological constant, scalar or vector fields (see \\cite{Weinberg89,Wetterich:1994bg, Caldwell98,Peebles03,Brax:1999gp,Brookfield:2005td, Boehmer:2007qa} and references therein). Most of the recent papers in this kind of studies are based on the assumption that the DE evolves independently of the dark matter (DM). The unknown nature of both DM and DE implies that we can not preclude future surprises regarding the interactions in the dark sector. This is very important because interactions between the DM and {\\em quintessence} could provide possible solutions to the cosmological coincidence problem. Recently, several papers have been published in this area \\cite{Amm:1999, Bin:2006} proposing that the DE and DM could be coupled, assuming also that there is only one type of non-interacting DM. However, there are other possibilities. (a) It is plausible that the dark matter is self-interacting (IDM) \\cite{SperStei}, a possibility that has been proposed to solve discrepancies between theoretical predictions and astrophysical observations, among which the gamma-ray and microwave emission from the center of our galaxy (eg. \\cite{92, 90}, \\cite{hoop, regis} and references therein). It has also been shown that some dark matter interactions could provide an accelerated expansion phase of the Universe \\cite{zim, balakin, Lima2008}. (b) The DM could potentially contain more than one particle species, for example a mixture of cold and warm or hot dark matter \\cite{Farrar:2003uw,Gubser:2004uh}, with or without inter-component interactions. In this work we are not concerned with the viability of the different such possibilities, nor with the properties of interacting DM models. The aim of this work is to investigate only whether there are repercussions of DM self-interactions for the global dynamics of the universe and specifically whether such models can yield an accelerated phase of the cosmic expansion, without the need of the {\\em dark energy}. We note that we do not ``design'' the fluid interactions to produce the desired accelerated cosmic evolution, as in some previous works (eg., \\cite{balakin}), but investigate the circumstances under which the analytical solution space of the collisional Boltzmann equation, in the expanding universe, allows for a late accelerated phase of the universe. ", "conclusions": "In this work we investigate analytically the evolution of the global density of the universe in the context of an interacting DM scenario by solving analytically the collisional Boltzmann equation in an expanding Universe. The possible disequilibrium between the DM particle creation and annihilation processes, regardless of its cause and in which the particle creation term dominates, creates an effective source term with negative pressure which (acting as {\\em dark energy}) provides an accelerated expansion phase of the scale factor. Finally, comparing the observed Hubble function of a few high-redshift elliptical galaxies with that predicted by our simplest IDM model ($m=0$), we find that the effective annihilation term is quite small." }, "0911/0911.3316_arXiv.txt": { "abstract": "{Recent investigations of the white dwarf (WD) + He star channel of Type Ia supernovae (SNe Ia) imply that this channel can produce SNe Ia with short delay times. The companion stars in this channel would survive and potentially be identifiable.} {In this {\\em Letter}, we study the properties of the companion stars of this channel at the moment of SN explosion, which can be verified by future observations.} {According to SN Ia production regions of the WD + He star channel and three formation channels of WD + He star systems, we performed a detailed binary population synthesis study to obtain the properties of the surviving companions.} {We obtained the distributions of many properties of the companion stars of this channel at the moment of SN explosion. We find that the surviving companion stars have a high spatial velocity ($>$400\\,km/s) after SN explosion, which could be an alternative origin for hypervelocity stars (HVSs), especially for HVSs such as US 708.} {} ", "introduction": "\\label{1. Introduction} Type Ia supernovae (SNe Ia) play an important role in the study of cosmic evolution, especially in cosmology. They have been applied successfully in determining cosmological parameters (e.g., $\\Omega$ and $\\Lambda$; Riess et al. 1998; Perlmutter et al. 1999). It is generally believed that SNe Ia are thermonuclear explosions of carbon--oxygen white dwarfs (CO WDs) in binaries (for the review see Nomoto et al. 1997). However, there is still no agreement on the nature of their progenitors (Hillebrandt \\& Niemeyer 2000; Podsiadlowski et al. 2008; Wang et al. 2008), and no SN Ia progenitor system has been conclusively identified from before the explosion. Over the past few decades, two families of SN Ia progenitor models have been proposed, i.e., the double-degenerate (DD) and single-degenerate (SD) models. Of these two models, the SD model is widely accepted at present (Nomoto et al. 1984). It is suggested that the DD model, which involves the merger of two CO WDs (Iben \\& Tutukov 1984; Webbink 1984; Han 1998), likely leads to an accretion-induced collapse rather than to an SN Ia (Nomoto \\& Iben 1985). For the SD model, the companion is probably a main-sequence (MS) star, a slightly evolved subgiant star (WD + MS channel), or a red-giant star (WD + RG channel) (e.g., Hachisu et al. 1996, 1999a,b; Li $\\&$ van den Heuvel 1997; Langer et al. 2000; Han $\\&$ Podsiadlowski 2004, 2006; Chen $\\&$ Li 2007, 2009; Meng et al. 2009; L\\\"{u} et al. 2009; Wang, Li \\& Han 2009). An explosion following the merger of two WDs would leave no remnant, while the companion star in the SD model would survive and be potentially identifiable (Podsiadlowski 2003). There has been no conclusive proof yet that any individual object is the surviving companion star of an SN Ia. It will be a promising method to test SN Ia progenitor models by identifying their surviving companion stars. Yoon $\\&$ Langer (2003) followed the evolution of a CO WD + He star system with a $1.0\\,M_{\\odot}$ CO WD and a $1.6\\,M_{\\odot}$ He star in a 0.124\\,d orbit. In this binary, the WD accretes He from the He star and grows in mass to the Chandrasekhar (Ch) mass. SNe Ia from this binary channel can neatly avoid H lines. Recently, Wang et al. (2009a) systematically studied the WD + He star channel of SNe Ia. In the study, they carried out binary evolution calculations of this channel for about 2600 close WD binaries, in which a CO WD accretes material from an He MS star or an He subgiant to increase its mass to the Ch mass. The study shows the parameter spaces for the progenitors of SNe Ia. By using a detailed binary population synthesis (BPS) approach, Wang et al. (2009b) find that the Galactic SN Ia birthrate from this channel is $\\sim$$0.3\\times 10^{-3}\\ {\\rm yr}^{-1}$ and that this channel can produce SNe Ia with short delay times ($\\sim$45$-$140\\,Myr) from the star formation to SN explosion. The companion star in this channel would survive and show distinguishing properties. In recent years hypervelocity stars (HVSs) have been observed in the halo of the Galaxy. HVSs are stars with a velocity so great that they are able to escape the gravitational pull of the Galaxy. However, the formation of HVSs is still unclear (for a recent review see Tutukov \\& Fedorova 2009). It has been suggested that such HVSs can be formed by the tidal disruption of a binary through interaction with the super-massive black hole (SMBH) at the Galactic center (GC) (Hills 1988; Yu \\& Tremaine 2003). The first three HVSs have only recently been discovered serendipitously (e.g., Brown et al. 2005; Hirsch et al. 2005; Edelmann et al. 2005). Up to now, about 17 HVSs have been discovered in the Galaxy (Brown et al. 2009; Tillich et al. 2009), most of which are B-type stars, probably with masses ranging from 3 to 5\\,$M_\\odot$ (Brown et al. 2005, 2009; Edelmann et al. 2005). One HVS, HE 0437-5439, is known to be an apparently normal early B-type star. Edelmann et al. (2005) suggests that the star could have originated in the Large Magellanic Cloud, because it is much closer to this galaxy (18 kpc) than to the GC (see also Przybilla et al. 2008). At present, only one HVS, US 708, is a subdwarf O (sdO) star, and Hirsch et al. (2005) speculatS that US 708 is formed by the merger of two He WDs in a close binary induced by the interaction with the SMBH in the GC and then escaped. Recently, Perets (2009) has suggested that US 708 may have been ejected as a binary from a triple disruption by the SMBH, which later on evolved and merged to form an sdO star. Because of the existence of the short orbital periods ($\\sim$1\\,h) for the WD + He star systems, Justham et al. (2009) argues that the WD + He star channel of SNe Ia may provide a natural explanation for stars like US 708. Han (2008a) obtained the distributions of many properties of the surviving companions from the WD + MS channel of SNe Ia. The properties can be verified by future observations. The purpose of this {\\em Letter} is to investigate the properties of the surviving companions of the WD + He star channel and to explore whether HVSs such as US 708 could have been released from the binaries that produced SNe Ia. In Section 2, we describe the BPS approach and the simulation results for the properties of the surviving companions. Finally, a discussion is given in Section 3. ", "conclusions": "\\label{3. Discussion} \\begin{figure} \\includegraphics[width=5.6cm,angle=270]{f1.ps} \\caption{The distribution of properties of companion stars in the plane of ($V_{\\rm orb}$, $M_2^{\\rm SN}$) at the current epoch, where $V_{\\rm orb}$ is the orbital velocity and $M_2^{\\rm SN}$ the mass at the moment of SN explosion.} \\end{figure} \\begin{figure} \\includegraphics[width=5.6cm,angle=270]{f2.ps} \\caption{Similar to Fig. 1, but in the plane of ($\\log T_{\\rm eff}$, $\\log g$), where $T_{\\rm eff}$ is the effective temperature of companion stars at the moment of SN explosion and $\\log g$ the surface gravity. The companion stars are out of thermal equilibrium, e.g., a 1.243$\\,M_\\odot$ He star with 0.274$\\,R_\\odot$ and ($\\log T_{\\rm eff}$, $\\log g$)=(4.70, 5.65) will be a 0.267$\\,R_\\odot$ He star and with ($\\log T_{\\rm eff}$, $\\log g$)=(4.75, 5.67) after the He star back to the thermal equilibrium.} \\end{figure} \\begin{figure} \\includegraphics[width=8.6cm,angle=0]{f3.ps} \\caption{The distribution of spatial velocity for surviving companion stars of SNe Ia. The dotted line denotes the results of ejecta velocity 11000\\,km/s, while the solid line shows ejecta velocity 13500\\,km/s.} \\end{figure} \\begin{figure} \\includegraphics[width=5.6cm,angle=270]{f4.ps} \\caption{Similar to Fig. 1, but in the plane of ($\\log P^{\\rm SN}$, $M_2^{\\rm SN}$), where $P^{\\rm SN}$is the orbital period at the moment of SN explosion.} \\end{figure} \\begin{figure} \\includegraphics[width=5.6cm,angle=270]{f5.ps} \\caption{Similar to Fig. 1, but in the plane of ($V_{\\rm rot}$, $M_2^{\\rm SN}$), where $V_{\\rm rot}$ is the equatorial rotational velocity of companion stars at the moment of SN explosion.} \\end{figure} Figures 1 and 2 show the distributions of the masses, the orbital velocities, the effective temperatures, and the surface gravities of companion stars at the moment of SN explosion. In Figure 1, the companion star has an orbital velocity of $\\sim$300$-$500\\,${\\rm km/s}$ for a corresponding mass of $\\sim$0.6$-$1.7\\,$M_\\odot$ at the moment of SN explosion. In the three formation channels of the WD + He star systems, SNe Ia mainly come from the He star channel. For this formation channel, the recorded properties at each step show that a primordial binary system with a primary mass $M_{\\rm 1i}\\sim 5.0-8.0\\,M_\\odot$, a secondary mass $M_{\\rm 2i} \\sim 2.0-6.5\\,M_\\odot$, and an orbital period $P_{\\rm i} \\sim 10-40\\,{\\rm d}$ would evolve to a close WD + He star system with a WD mass $M^{\\rm i}_{\\rm WD} \\sim 0.87-1.2\\,M_\\odot$, a He star mass $M^{\\rm i}_{\\rm 2} \\sim 1.0-2.6\\,M_\\odot$, and an orbital period $P^{\\rm i} \\sim 0.04-0.2\\,{\\rm d}$. Finally, the WD + He star system results in an SN Ia explosion and survives a companion star. However, Figures 1 and 2 are for the moment of SN explosion, which could be modified by the explosion. The SN ejecta will interact with its companion. The companion stars will be stripped of some mass and receive a kick velocity that is perpendicular to the orbital velocity. Adopting a similar method to Meng, Chen \\& Han (2007), we estimated the stripped mass for the companion stars of the WD + He star channel and find that the stripped mass is very low, e.g., a 1.243$\\,M_\\odot$ companion star with 0.274$\\,R_\\odot$ only loses a mass of 0.015$\\,M_\\odot$ in our simulation. We also roughly obtain the kick velocity based on the momentum conservation equation.\\footnote{ However, in reality, the collision between the SN ejecta and its companion cannot be elastic. If we adopt the inelastic collision in our calculations, the kick velocity will only decrease by $\\sim$3\\%$-$5\\%. Thus, the inelastic collision is no significant influence on the kick velocity.} The kick velocity mainly depends on the ratio of separation to the radius of companions at the moment of SN explosion, $A/R_{\\rm 2}^{\\rm SN}$, and the leading head velocity of SN ejecta, which is assumed to be in the range of 11000 to 13500\\,km/s. The velocity of 11000\\,km/s is from the SN ejecta kinetic energy $1.0\\times10^{51}$\\,erg corresponding to the lower limit of normal SN Ia kinetic energy, while the velocity of 13500\\,km/s is from the SN ejecta kinetic energy $1.5\\times10^{51}$\\,erg corresponding to the upper limit of the kinetic energy (Gamezo et al. 2003). We can obtain the spatial velocity by the formula $V_{\\rm 2}^{\\rm SN}=\\sqrt{V_{\\rm kick}^{2}+V_{\\rm orb}^{2}}$, where $V_{\\rm kick}$ and $V_{\\rm orb}$ are the kick velocity and the orbital velocity of the companion star at the moment of SN explosion, respectively. In Figure 3, we show the distribution of the spatial velocity for the surviving companion stars of the WD + He star channel. We see that the surviving companion stars have high spatial velocities ($>$400\\,km/s) that almost entirely exceed the gravitational pull of the Galaxy nearby the sun. Thus, the surviving companion stars from the WD + He star channel could be an alternative origin for HVSs. US 708 is an extremely He-rich sdO star in the Galaxy halo, with a heliocentric radial velocity of +$708\\pm15$\\,${\\rm km/s}$ (Hirsch et al. 2005). We note that the local velocity relative to the Galatic center may lead to a higher observation velocity for the surviving companion stars, but this may also lead to a lower observation velocity. Considering the local velocity near the sun ($\\sim$220\\,km/s), we find that $\\sim$30\\% of the surviving companion stars may be observed to have velocity $V>700\\,{\\rm km/s}$ for a given SN ejecta velocity 13500\\,km/s. In addition, the asymmetric explosion of SNe Ia may also enhance the velocity of the surviving companions. Thus, a surviving companion star in the WD + He star channel may have a high velocity like US 708 (see also Justham et al. 2008). The companion stars are out of thermal equilibrium at the moment of SN explosion. For He stars, the equilibrium radii are lower than at the moment of SN explosion. Thus, the surface gravity at equilibrium should be greater than the one in Figure 2; e.g., a 1.243$\\,M_\\odot$ He star with 0.274$\\,R_\\odot$ and ($\\log T_{\\rm eff}$, $\\log g$)=(4.70, 5.65) will be a 0.267$\\,R_\\odot$ He star and with ($\\log T_{\\rm eff}$, $\\log g$)=(4.75, 5.67) after the He star is back to the thermal equilibrium. It is well known that a shock will develop after impact by the ejecta. A large part of the material in the companion's envelope is heated by the shock, and some of the material is vaporized from the surface of the companion stars. Following a similar method to Chen \\& Li (2007), we also estimated the vaporized mass, and find that the mass loss from the companion stars is not significant ($<$5\\%). Such a surviving companion star may be significantly overluminous or underluminous depending on the amount of heating (e.g., Podsiadlowski 2003). Figure 2 could be a starting point for further studies of this kind. Figure 4 shows the distributions of orbital periods and secondary masses of the WD + He star systems at the moment of SN explosion. The orbital periods and secondary masses of the WD + He star systems at this moment are basic input parameters when one simulates the interaction between SN ejecta and its companion. It is suggested that, for hot stars with radiative envelopes (such as He stars), tidal forces may be inefficient for synchronization (e.g., Zahn 1977). However, the recent study by Charpinet et al. (2008) supports efficient tidal synchronization for hot subdwarf stars. The work by Toledano et al. (2007) also indicates, on the other hand, that even stars with radiative envelopes may have efficient tidal interaction on a time scale comparable to convective envelopes. Thus, we make an assumption that the companion stars co-rotate with their orbits. In Figure 5, we show the distributions of equatorial rotational velocities of the companions stars. We see that the surviving companion stars are fast rotators, so their spectral lines should be broadened noticeably. The simulation in this {\\em Letter} was made with $\\alpha_{\\rm ce}\\lambda =0.5$. If we adopt a higher value for $\\alpha_{\\rm ce}\\lambda$, e.g., 1.5, the birthrate of SNe Ia would be a little bit higher and the delay time from the star formation to SN explosion longer. This is because binaries emerging from CE ejections tend to have longer orbital periods for a large $\\alpha_{\\rm ce}\\lambda$ and are more likely to be located in the SN Ia production region (Fig. 8 of Wang et al. 2009a). Due to the lack of WD binaries with short orbital periods ($\\log P^{\\rm i} < -1.2$) for a large $\\alpha_{\\rm ce}\\lambda$, the companions with orbital velocity $V_{\\rm orb}>430\\,{\\rm km/s}$ would be noticeably absent in Figure 1. The distributions are results of the current epoch for a constant SFR. For a single starburst, most of the SN explosions occur between $\\sim$45\\,Myr and $\\sim$140\\,Myr after the starburst; i.e., SNe Ia from the WD + He star channel will be absent in old galaxies. The Galactic SN Ia birthrate from this channel is $\\sim$$0.3\\times 10^{-3}\\ {\\rm yr}^{-1}$ (Wang et al. 2009b). By multiplying the birthrate with a typical MS lifetime of He stars, $\\sim$$10^7$\\,yr, we estimated the current number of this type of HVSs in the Galaxy to be $\\sim$$10^3$. In future investigations, we will employ the Large sky Area Multi-Object fiber Spectral Telescope (LAMOST) to search the HVSs originating from the surviving companion stars of SNe Ia." }, "0911/0911.1639_arXiv.txt": { "abstract": "I show that Eddington accretion episodes in AGN are likely to produce winds with velocities $v \\sim 0.1c$ and ionization parameters up to $\\xi \\sim 10^4$~(cgs), implying the presence of resonance lines of helium-- and hydrogenlike iron. These properties are direct consequences of momentum and mass conservation respectively, and agree with recent X--ray observations of fast outflows from AGN. Because the wind is significantly subluminal, it can persist long after the AGN is observed to have become sub--Eddington. The wind creates a strong cooling shock as it interacts with the interstellar medium of the host galaxy, and this cooling region may be observable in an inverse Compton continuum and lower--excitation emission lines associated with lower velocities. The shell of matter swept up by the (`momentum--driven') shocked wind must propagate beyond the black hole's sphere of influence on a timescale $\\la 3\\times 10^5$~yr. Outside this radius the shell stalls unless the black hole mass has reached the value $M_{\\sigma}$ implied by the $M - \\sigma$ relation. If the wind shock did not cool, as suggested here, the resulting (`energy--driven') outflow would imply a far smaller SMBH mass than actually observed. In galaxies with large bulges the black hole may grow somewhat beyond this value, suggesting that the observed $M -\\sigma$ relation may curve upwards at large $M$. Minor accretion events with small gas fractions can produce galaxy--wide outflows with velocities significantly exceeding $\\sigma$, including fossil outflows in galaxies where there is little current AGN activity. Some rare cases may reveal the energy--driven outflows which sweep gas out of the galaxy and establish the black hole--bulge mass relation. However these require the quasar to be at the Eddington luminosity. ", "introduction": "Outflows and winds are widely observed in many types of galaxies, from active to normal (e.g. Chartas et al., 2003; Pounds et al., 2003a, 2006; O'Brien et al., 2005, Holt et al., 2008; Krongold et al., 2007; Tremonti et al., 2007). There is often a strong presumption that the central supermassive black hole (SMBH) is implicated. Although other types of driving exist, e.g. by supernovae or starbursts, in many cases these phenomena are themselves associated with accretion episodes on to the SMBH. From a theoretical viewpoint, outflows driven by black holes offer a simple way of establishing relations between the SMBH and its host galaxy, and hence potential explanations for the $M - \\sigma$ and $M - M_{\\rm bulge}$ relations (Ferrarese \\& Merritt, 2000; Gebhardt et al. 2000; H\\\"aring \\& Rix 2004). Such outflows are plausible, as AGN must feed at high rates to grow the observed SMBH masses, given the short duty cycle implied by the rarity of AGN among all galaxies. There is no obvious reason why these rates should respect the hole's Eddington limit, and outflows are a natural consequence. However it is unclear just how, if at all, the various types of observed outflows mentioned in the first paragraph fit together in terms of these ideas. This paper aims to clarify this. ", "conclusions": "We have shown that Eddington accretion episodes in AGN produce winds with velocities $v \\sim 0.1c$ and ionization parameters $\\xi \\sim 10^4$~(cgs) requiring the presence of resonance lines of helium-- and hydrogenlike iron. These properties follow from momentum and mass conservation, and agree with recent X--ray observations of high--speed outflows from AGN. Because the winds have speeds $\\sim 0.1c$ they can persist long after the AGN have become sub--Eddington. The wind creates a strong cooling shock as it impacts the interstellar medium of the host galaxy. This cooling region may be observable, as it produces an inverse Compton continuum and lower--excitation emission lines associated with lower velocities. The shell of matter swept up by the ('momentum--driven') shocked wind emerges from the black hole's sphere of influence on a timescale $\\la 3\\times 10^5$~yr. The shell then stalls unless the black hole mass has reached the value $M_{\\sigma}$ implied by the $M - \\sigma$ relation. If the wind shock did not cool, as suggested here, the resulting (`energy--driven') outflow would imply a far smaller SMBH mass than actually observed. In galaxies with large bulges the black hole must grow somewhat beyond this value, suggesting that the observed $M -\\sigma$ relation may curve upwards at large $M$. Galaxy--wide outflows with velocities significantly exceeding $\\sigma$ probably result from minor accretion episodes with low gas fractions. These can appear as fossil outflows in galaxies where there is little current AGN activity. In rare cases it may be possible to observe the energy--driven outflows which sweep gas out of the galaxy and establish the black hole--bulge mass relation. However these require the quasar to be at the Eddington luminosity." }, "0911/0911.3120_arXiv.txt": { "abstract": "{The historic plates of the {\\it Carte du Ciel}, an international cooperative project launched in 1887, offer valuable first-epoch material for the determination of absolute proper motions. } {We present the CdC-SF, an astrometric catalogue of positions and proper motions derived from the \\textit{Carte du Ciel} plates of the San Fernando zone, photographic material with a mean epoch of 1901.4 and a limiting magnitude of V$\\sim$16, covering the declination range of $-10^{\\circ} \\leq \\delta \\leq -2^{\\circ}$. } {Digitization has been made using a conventional flatbed scanner. Special techniques have been developed to handle the combination of plate material and the large distortion introduced by the scanner. The equatorial coordinates are on the ICRS defined by Tycho-2, and proper motions are derived using UCAC2 as second-epoch positions. } { The result is a catalogue with positions and proper motions for 560000 stars, covering 1080 degrees$^2$. The mean positional uncertainty is 0$\\farcs$20 (0$\\farcs$12 for well-measured stars) and the proper-motion uncertainty is 2.0 mas/yr (1.2 mas/yr for well-measured stars). } {The proper motion catalogue CdC-SF is effectively a deeper extension of Hipparcos, in terms of proper motions, to a magnitude of 15. } ", "introduction": "The Hipparcos Catalogue (Perryman et al. 1997) contains the positions, parallaxes and proper motions of 118218 stars with great precision (proper motions precision is 1~mas/year), complete up to a magnitude of $V=7.3$. Due to its precision and homogeneity, this catalogue constitutes an important source of information and has provided a useful tool in numerous kinematic studies. However, its relatively bright magnitude limit restricts it to very nearby stars, with the number of stars per unit area also being lower than what is often desired in kinematic analyses of the Galaxy. For this reason, various efforts have been made to develop astrometric catalogues that are more dense and reach deeper magnitudes, becoming in effect an extension of the Hipparcos catalogue. It is necessary that such catalogues contain proper motions besides positions in order to provide the essential information with the quality and completeness needed. The Tycho-2 Catalogue (H{\\o}g et al. ~ 2000) can be considered an extension of Hipparcos in the way that it contains positions and proper motions for stars brighter than V$\\sim$12 with errors of 2.5~mas/year. At the present time the Hipparcos and Tycho-2 catalogues are the largest high-quality proper motion surveys. Other proper motion catalogues available at the present time are UCAC2 catalogues (Zacharias et al. 2004, 2--7 mas/yr depending on the magnitude and current version has systematic errors of the order of 3~mas/yr), NPM (Hanson et al. 2004, 5~mas/yr, $81.5\\times10^{24}$~cm$^{-2}$). On the basis of these techniques we show that SMBH growth was typically heavily obscured ($N_{\\rm H}\\ge10^{23}$~cm$^{-2}$) at $z\\approx$~2, and find that the growth of the SMBH and spheroid was closely connected, even in the most rapidly evolving systems. ", "introduction": "% There is no doubt that Active Galactic Nuclei (AGN) are an important component in the formation and evolution of galaxies. The seminal discovery that massive galaxies in the local Universe harbor a super-massive black hole (SMBH) indicates that they must have hosted AGN activity over the past $\\approx$~13~Gyrs (e.g.,\\ Soltan 1982; Kormendy \\& Richstone 1995). Furthermore, the close relationship between the mass of the SMBH and the spheroid in local galaxies points towards a close connection between the growth of the SMBH and the host galaxy (e.g.,\\ Magorrian et~al. 1998; Ferrarese \\& Merritt 2000). Finally, the most successful galaxy formation and large-scale structure models require AGN outflows to suppress star formation and reproduce the properties of massive galaxies in the local Universe (e.g.,\\ Croton et~al. 2006; Bower et~al. 2006). Stellar population synthesis modelling of nearby galaxies and studies of the cosmic history of star formation indicate that the bulk of the stellar build up of massive galaxies must have occurred at high redshift ($z\\approx$~2; i.e.,\\ when the Universe was $\\approx$~25\\% the current age; e.g.,\\ Heavens et~al. 2004; Hopkins et~al. 2006). Under the assumption that the growth of the SMBH is concordant with that of the galaxy spheroid, the dominant growth phase of SMBH growth in massive galaxies must also have occurred at high redshift. Currently the most {\\it efficient} identification of AGNs (and therefore SMBH growth) is made with deep X-ray surveys (e.g.,\\ Brandt \\& Hasinger 2005). For example, the deepest X-ray surveys (e.g.,\\ Alexander et~al. 2003; Luo et~al. 2008) are able to detect even moderately luminous AGN activity at $z\\approx$~2 ($L_{\\rm X}\\approx10^{42}$--$10^{43}$~erg~s$^{-1}$; i.e.,\\ $>10$ times below the canonical threshold assumed for quasars). Furthermore, the high rest-frame energies probed by these surveys at $z\\approx$~2 ($\\approx$~1.5--24~keV) means that the selection of AGNs is {\\it almost} obscuration independent. This latter point is important since the dominant growth phase of massive galaxies appears to have been heavily obscured by dust and gas (e.g.,\\ Chapman et~al. 2005; Le Floc'h et~al. 2005). In this brief contribution we hightlight some of our recent work aimed at identifying obscured AGN activity in distant rapidly evolving galaxies at $z\\approx$~2; see \\S2 \\& \\S3. We discuss these results within the context of the SMBH--sperhoid growth phase of today's massive galaxies ($M_{\\rm GAL}\\approx10^{11}$~$M_{\\odot}$); see \\S4. ", "conclusions": "" }, "0911/0911.1063_arXiv.txt": { "abstract": "We discuss a possible solution to the cosmological constant problem based on the hypothesis of a fixed point for higher dimensional quantum gravity coupled to a scalar. At the fixed point, which is reached only for an infinite value of the scalar field, dilatation symmetry becomes exact. For this limit we concentrate on the absence of a scalar potential since a polynomial potential is not consistent with dilatation symmetry in higher dimensions. We find generic solutions of the higher dimensional field equations for which the effective four-dimensional cosmological constant vanishes, independently of the parameters of the higher dimensional effective action. Under rather general circumstances these are the only quasistatic stable extrema of the effective action which lead to a finite four-dimensional Planck mass. We discuss the associated higher dimensional self-tuning mechanism for the cosmological constant. If cosmological runaway solutions approach the fixed point as time goes to infinity, the effective dark energy vanishes asymptotically. In the present cosmological epoch the fixed point is not yet reached completely, resulting in a tiny amount of dark energy, comparable to dark matter. We discuss explicitly higher dimensional geometries which realize such asymptotic solutions for time going to infinity. ", "introduction": "\\label{intro} Anomalous dilatation symmetry may be a key ingredient for a dynamical solution of the cosmological constant problem \\cite{CWQ}, \\cite{PSW}. In the asymptotic limit of time $t$ going to infinity, a cosmological runaway solution of the field equations can approach a fixed point. At the fixed point all memory of explicit mass or length scales is lost and the quantum effective action becomes dilatation symmetric. In other words, the dilatation anomaly vanishes when it is evaluated for the field configurations corresponding to the fixed point \\cite{CWCC}. In consequence, a scalar field becomes massless in the asymptotic limit for $t\\to\\infty$, corresponding to the Goldstone boson of spontaneously broken dilatation symmetry. For the approach to the fixed point at finite $t$ the anomaly is not yet zero, and correspondingly the scalar ``pseudo-Goldstone boson'' still has a small mass, that vanishes only asymptotically. These ideas are realized in practice in quintessence cosmologies, where the ``cosmon''-field plays the role of the pseudo-Goldstone boson of spontaneously broken anomalous dilatation symmetry \\cite{CWQ}, \\cite{CWCC}. The cosmon mass is varying with time and of the order of the Hubble parameter \\cite{CWAA}. Before discussing cosmological solutions, it is crucial to understand the asymptotic solution towards which the cosmological runaway solution converges asymptotically. In our scenario, this asymptotic solution should correspond to flat Minkowski space and therefore to an ``asymptotically vanishing cosmological constant''. However, a vanishing cosmological constant is not enforced by dilatation symmetry - asymptotic dilatation symmetry could, in principle, also be realized with a non-zero effective cosmological constant. In two recent papers \\cite{CWCC}, \\cite{CWNL} we have argued that a higher dimensional setting sheds new light on the question why the effective four-dimensional cosmological constant $\\Lambda$ vanishes asymptotically. Actually, such an asymptotic vanishing happens for a large generic class of cosmological runaway solutions without any tuning of parameters. The remaining problem concerns the strict limits on the possible variation of fundamental couplings, which are not obeyed for some classes of such solutions. In ref. \\cite{CWCC} we have discussed several scenarios how to reconcile asymptotically static couplings (with an interesting possible exception in the neutrino sector \\cite{CWN}) and an asymptotically vanishing cosmological constant. In the present paper we investigate higher dimensional models for which the quantum effective action exhibits an exact dilatation symmetry for a suitable fixed point. This fixed point should become relevant for the asymptotic solution. A dilatation symmetric effective action is also the starting point for approaches where dilatation symmetry is realized as an exact quantum symmetry \\cite{SD}. Thus the basic object of our investigation is the quantum effective action $\\Gamma$ where all quantum fluctuations are already included. The field equations derived from $\\Gamma$ are exact without any further quantum corrections. Our key finding is that in the presence of higher dimensional dilatation symmetry these field equations have generic solutions for which the four-dimensional cosmological constant vanishes, $\\Lambda=0$. They correspond to stable extrema of $\\Gamma$. We do not postulate here that the effective action of a fundamental theory is dilatation symmetric. In general, it is not, and dilatation anomalies are present. For example, a dilatation anomaly arises if running couplings induce an explicite scale, or if a higher dimensional cosmological constant is present. We only make the hypothesis that $\\Gamma$ has a fixed point for certain asymptotic field values to be specified below. Only in this asymptotic limit the dilatation anomaly vanishes - only the ``fixed point effective action'' is dilatation symmetric. For a cosmological runaway solution the values of fields are not static but continue to change for all times. We investigate solutions where the fields move towards the fixed point region for $t\\to\\infty$. For $t\\to\\infty$ the field equations derived from the fixed point effective action become accurate. As the runaway solution approaches a fixed point the anomalous parts of the effective action vanish. Since such a fixed point is generally reached only asymptotically for $t\\to\\infty$, it is only in this limit that the field equations exhibit an exact dilatation symmetry. Our discussion of a dilatation symmetric effective action will therefore be insufficient for the cosmological solutions. The solutions of the field equations derived from the dilatation symmetric effective action account only for the possible asymptotic states for $t\\to\\infty$. Nevertheless, it is a necessary condition in our scenario that the dilatation symmetric effective action has an extremum which describes an acceptable asymptotic state. This solution should have flat four-dimensional space. Furthermore, a simple solution to the problem of time-varying couplings would be a static ratio between the characteristic length scale of internal geometry and the effective four-dimensional Planck length. The latter should lead after dimensional reduction to finite non-zero values of the gauge couplings and other dimensionless couplings and mass ratios in the resulting model of particle physics. Inversely, if a satisfactory asymptotic solution exists, the chances that a runaway solution will approach it for $t\\to\\infty$ are quite high. Dilatation symmetry is easily realized in models that contain besides the metric also a scalar field $\\xi$. A striking difference between dilatation symmetry in higher dimensions, $d>6$, as compared to four dimensions is the absence of a polynomial potential for a scalar field with a canonical kinetic term \\cite{CWCC}. Indeed, for $d=4~mod~4$ (or $d=2~mod~4$ with a symmetry $\\xi\\to-\\xi$) the most general effective action , which is polynominal in $\\xi$ and consistent with general coordinate transformations (diffeomorphism-symmetry) and global dilatation symmetry, reads \\begin{equation}\\label{AA1a} \\Gamma=\\int_{\\hat x}\\hat g^{1/2}\\Big\\{-\\frac12\\xi^2\\hat R +\\frac\\zeta2\\partial^{\\hat\\mu}\\xi\\partial_{\\hat\\mu}\\xi +F(\\hat R_{\\hat\\mu\\hat\\nu\\hat\\rho\\hat\\sigma})\\Big\\}. \\end{equation} The normalization of $\\xi$ is chosen such that it has a canonical kinetic term, up to the free dimensionless parameter $\\zeta$. While for $d=4$ a term $\\lambda\\xi^4$ is dilatation invariant (and for $d=6$ a term $\\gamma\\xi^3$), no polynomial $\\xi^n$ with integer $n$ is dilatation invariant for $d>6$. We may first investigate the case that $F$ contains only polynomials of the curvature tensor $\\hat R_{\\hat\\mu\\hat\\nu\\hat\\rho\\hat\\sigma}$ or derivatives thereof. A typical example for $F$ is \\begin{equation}\\label{AA2a} F=\\tau\\hat R^{\\frac d2}, \\end{equation} and we will discuss more general polynomial forms of $F$ in the next section (cf. ref. \\cite{CWCC}, \\cite{CWNL}.) For odd dimensions no invariant polynomial in the curvature tensor exists and therefore $F=0$. (We disregard invariants involving the $\\epsilon$-tensor.) We will often concentrate on the simplest form of such a fixed point effective action by taking $F=0$ in eq. \\eqref{AA1a}. Nevertheless, we also discuss more general forms. The polynomial form of $F$ is actually not essential for our findings and we will extend our discussion to the most general dilatation symmetric term $F$ that only involves the metric. In contrast, the absence of a dilatation symmetric (non-polynomial) potential $V(\\xi)$ at the fixed point remains important for certain aspects of our discussion. We only briefly comment in sect. \\ref{quasi} on possible modifications if a non-polynominal potential for $\\xi$ would be present at the fixed point. It seems likely that this will not change the existence of extrema of $\\Gamma$ with $\\Lambda=0$. The key argument for the generic existence of solutions with $\\Lambda=0$ is rather simple and directly related to higher dimensional dilatation symmetry. Let us consider the most general form of a dilatation symmetric quantum effective action \\begin{equation}\\label{2A} \\Gamma=\\int_{\\hat x}\\hat g^{1/2}{\\cal L}. \\end{equation} Dilatation transformations correspond to a rescaling of the metric by a constant factor $\\alpha^2$, and an associated rescaling of $\\xi$, \\begin{eqnarray}\\label{2B} \\hat g_{\\hat\\mu\\hat\\nu} \\to \\alpha^2\\hat g_{\\hat\\mu\\hat\\nu}~,~\\hat g^{1/2}\\to\\alpha^d\\hat g^{1/2},\\nonumber\\\\ \\xi\\to\\alpha^{-\\frac{d-2}{2}}\\xi~,~{\\cal L}\\to\\alpha^{-d}{\\cal L}. \\end{eqnarray} A dilatation symmetric effective action remains invariant under these rescalings. For general field values we may define $\\Gamma_\\kappa[\\hat g_{\\hat\\mu\\hat\\nu},\\xi]=\\Gamma[\\kappa^{-2}\\hat g_{\\hat\\mu\\hat\\nu}, \\kappa^{\\frac{d-2}{2}}\\xi]$ and consider the asymptotic limit $\\kappa\\to\\infty$. Our fixed point hypothesis states then that $\\Gamma_\\kappa$ becomes dilatation symmetric for $\\kappa\\to\\infty$. The special role of dilatation symmetry for the problem of the cosmological constant is visible already for the most general form of a dilatation symmetric effective action. We are interested in configurations with a block diagonal metric \\begin{equation}\\label{2c} \\hat g_{\\hat\\mu\\hat\\nu} (x,y)= \\left(\\begin{array}{ccc} \\sigma(y)g^{(4)}_{\\mu\\nu} (x) &,&0\\\\ 0&,&g^{(D)}_{\\alpha\\beta}(y) \\end{array}\\right). \\end{equation} Here $x^\\mu$ denotes the four-dimensional coordinates and $y^\\alpha$ are coordinates of $D$-dimensional internal space, with corresponding metrics $g^{(4)}_{\\mu\\nu}$ and $g^{(D)}_{\\alpha\\beta},d=D+4$. The function $\\sigma(y)$ accounts for a possible warping \\cite{RSW,7A,RDW,RS}. The configuration for the metric is supplemented by a configuration for the scalar field $\\xi(y)$. With $\\hat g^{1/2}=(g^{(4)})^{1/2}\\sigma^2(g^{(D)})^{1/2}$ we define \\begin{equation}\\label{2D} W(x)=\\int_y(g^{(D)}{(y)})^{1/2}\\sigma^2(y){\\cal L}(x,y), \\end{equation} and write \\begin{equation}\\label{2E} \\Gamma=\\int_x(g^{(4)})^{1/2} W. \\end{equation} We observe the scaling under dilatations $(g^{(4)}_{\\mu\\nu}\\to\\alpha^2 g^{(4)}_{\\mu\\nu}~,~g^{(D)}_{\\alpha\\beta}\\to\\alpha^2 g^{(D)}_{\\alpha\\beta})$ \\begin{equation}\\label{2F} W\\to\\alpha^{-4}W. \\end{equation} The possible extrema of $\\Gamma$ can be classified into two categories, according to the existence of an extremum of $W(x)$ or not. Consider first the case where an extremum of $W(x)$ exists for an appropriate field configuration $\\hat g_{\\hat\\mu\\hat\\nu},\\xi$. This means that a (infinitesimally) close neighboring field configuration does not change the function $W(x)$. We may denote the value of $W(x)$ at its extremum by $W_0(x)$. The scaling property \\eqref{2F} immediately implies $W_0(x)=0$. Indeed, using in eq. \\eqref{2B} $\\alpha=1+\\epsilon$, with infinitesimal $\\epsilon$, defines a neighboring configuration. The combination of scaling and extremum condition, $\\partial_\\epsilon(1+\\epsilon)^{-4}W_0=0$, can be obeyed only for $W_0=0$. With eq. \\eqref{2E} and $W_0=0$ an extremum of $W$ is also an extremum of the effective action $\\Gamma$. Thus, whenever an extremum of $W$ exists, this defines an extremum of $\\Gamma$ obeying the field equations, with a vanishing value of $\\Gamma$ at the extremum, $\\Gamma_0=0$. For all solutions which admit dimensional reduction to an effective four dimensional ``local'' theory of gravity we will show that $\\Gamma_0=0$ implies that the effective four-dimensional cosmological constant $\\Lambda$ vanishes. The existence of extrema of $W$ is a rather generic feature and we find that within our setting stable extrema in this ``flat phase'' always exist. Only the question remains if this flat phase includes solutions with interesting particle physics. There may exist additional extrema of $\\Gamma$ which are not extrema of $W$ and for which $\\Lambda$ does not vanish. Four dimensional theories with a non-vanishing cosmological constant are possible only for this possible second category of extrema of $\\Gamma$ which are not extrema of $W(x)$. In the absence of a non-polynominal scalar potential we find that all such possible solutions in the non-flat phase are unstable if $\\xi \\neq 0$. This singles out a vanishing four-dimensional cosmological constant. The existence of extrema of $W$ is a qualitative feature that does not depend on the precise values of the parameters characterizing ${\\cal L}$. As an example we may consider the polynomial effective action \\eqref{AA1a}. It is obvious that $\\hat R_{\\hat\\mu\\hat\\nu\\hat\\rho\\hat\\sigma}=0~,~\\xi=\\xi_0=$const. corresponds to such an extremum, with ${\\cal L}_0=0$. The existence of this extremum extends to a very large class of non-polynomial ${\\cal L}$ as well. It is sufficient that the $\\xi$-dependent part of ${\\cal L}$ vanishes for flat space and constant $\\xi$, and that the remaining $\\xi$ independent part $F$ vanishes for $\\hat R_{\\hat\\mu\\hat\\nu\\hat\\rho\\hat\\sigma}=0$. Furthermore, if \\begin{equation}\\label{2G} W_F(x)=\\int_y(g^{(D)})^{1/2}\\sigma^2 F \\end{equation} admits an extremum with respect to variations of the metric, this must occur for $W_{F,0}(x)=0$ by the same scaling arguments as above. (No polynomial form of $F$ needs to be assumed here.) A dilatation symmetric theory involving only gravity (without the scalar field $\\xi$) will imply a vanishing cosmological constant if an extremum of $W_F$ exists which leads to acceptable four dimensional gravity (with nonzero and finite effective Planck mass). Let us next extend the setting by adding a scalar part ${\\cal L}_\\xi$ which is quadratic in $\\xi$, without being necessarily polynomial in the curvature tensor. For a given metric $\\hat g_{\\hat\\mu\\hat\\nu}$ we can find a partial extremum of $\\Gamma_\\xi=\\int_{\\hat x}\\hat g^{1/2}{\\cal L}_\\xi=\\int_x(g^{(4)})^{1/2} W_\\xi$ by solving the higher dimensional field equation for $\\xi$, consistent with an extremum condition in case of singular geometries which will be derived later. This results in $W_\\xi=0$, as expected for possible extrema of a purely quadratic polynomial. A necessary condition for an extremum of $W=W_\\xi+W_F$ remains therefore $W_F=0$. However, only the sum $W_\\xi+W_F$ has to be an extremum with respect to variations of the metric, and not $W_\\xi$ and $W_F$ separately. We discuss an example of this type of extrema in the appendix. We observe that for a quadratic ${\\cal L}_\\xi$ we could ``integrate out'' the scalar field in favor of non-local gravitational interactions. This demonstrates incidentally that our setting is not restricted to a local effective gravitational action. From these simple observations we conclude that in the presence of dilatation symmetry the vanishing of the cosmological constant $\\Lambda$ is very robust with respect to variations of the precise form of $\\Gamma$. The effective action may be characterized by many parameters, as for example the coefficients of different terms appearing in a polynomial approximation of $F$. A change of the values of these parameters will typically not change the value $\\Lambda=0$, as long as an extremum of $W$ continues to exist and is compatible with effective four dimensional gravity. This feature is a particular consequence of dilatation symmetry - it no longer holds in presence of dilatation anomalies. In view of the robustness of $\\Lambda=0$ we find many analogies to phases in many body theories. For the ``flat phase'' an extremum of $W$ exists and $\\Lambda=0$. Extrema of $\\Gamma$ in the ``non-flat phase'' are not extrema of $W$, and we will typically find $\\Lambda>0$. For effective four dimensional theories in the flat phase we will show that $W$ can be associated with the effective potential $V$ for four-dimensional scalar fields. The minimum of $V$ occurs then necessarily for $V_0=0$. Furthermore, $V$ has flat directions. One such flat direction corresponds to the dilatations \\eqref{2B} and the corresponding massless scalar field is the dilaton. There may be more flat directions since the extrema of $W(x)$ may occur for a whole class of different configurations $\\hat g_{\\hat\\mu\\hat\\nu},\\xi$. In this case we expect additional massless fields for the dilatation symmetric asymptotic cosmological solution for $t\\to\\infty$. For finite $t$ the presence of dilatation anomalies will induce mass terms for these scalars, which vanish only for $t\\to\\infty$. It is conceivable that such additional light scalar fields (beyond the cosmon) are interesting candidates for dark matter. Our investigation focuses on two issues. \\begin{itemize} \\item [(i)] The existence of extrema of $W$ and a first discussion of characteristic properties of solutions in the flat phase. \\item [(ii)] The dimensional reduction to effective four dimensional gravity and the establishment that $\\Lambda=0$ in the flat phase. \\end{itemize} A demonstration that an effective action which admits a flat phase does generically not allow other extrema with arbitrarily small $|\\Lambda|$ is given in ref. \\cite{CWNL}. In particular, there are no continuous families of extrema where $\\Lambda$ appears as a continuous parameter. This issue is important for warped geometries with singularities where the existence of families of solutions of the higher dimensional field equations with continuous $\\Lambda$ is known \\cite{RSW}, \\cite{7A}, \\cite{RDW}. In this case the extremum conditions for $\\Gamma$ go beyond the higher dimensional field equations \\cite{CWCON}. They precisely select the solutions with $\\Lambda=0$ out of the continuous family of solutions. The particular properties of warped spaces within our general dilatation symmetry setting will be discussed in an accompanying paper \\cite{P2}. In the course of our discussion we will explicitly address the issues of quantum corrections, ``tuning of the cosmological constant to zero'', and ``naturalness'' of the solutions with $\\Lambda=0$. Some of the general aspects are already discussed in \\cite{CWCC} and not repeated here, such that we concentrate here on our specific setting. We discuss the special role of higher dimensions for the ``self-tuning'' of the cosmological constant to zero. This self-tuning is particular to the case of dilatation symmetry and closely connected to the robustness of the existence of extrema of $W$ under a change of parameters in the effective action. We find rather satisfactory answers to the naturalness problem. Asymptotic dilatation symmetry in higher dimensional theories may indeed provide the key for a solution of the cosmological constant problem. Our paper is organized as follows. In sect. \\ref{dilatation} we present a first discussion of a polynomial effective action \\eqref{AA1a}. We establish explicitely the existence of extrema of $W$ belonging to the flat phase for a large class of effective actions, and demonstrate the vanishing of the cosmological constant if effective four dimensional gravity exists. We extend this discussion to the most general dilatation symmetric effective action and show that extrema in the flat phase with $\\Lambda=0$ exist whenever an appropriate $d-4$ dimensional functional $\\bar W$ admits an extremum. Section \\ref{quasi} addresses the existence of the two ``phases'' for solutions in an effective four dimensional framework and shows why $\\Lambda=0$ is singled out in the flat phase. We concentrate on ``quasistatic solutions'' for which the four dimensional fields are static and homogeneous, while the metric describes a geometry with maximally four dimensional symmetry, with positive, negative or vanishing $\\Lambda$. It also demonstrates that the non-flat phase with $\\Lambda>0$ can only be realized if certain conditions in parameter space are met. The non-flat phase appears to be less generic than the flat phase. We find that all possible extrema in the non-flat phase for $\\xi \\neq 0$ are unstable. In sect. \\ref{extended} we discuss an interesting ``extended scaling symmetry'' which becomes realized for the simplest fixed point with $F=0$, or if the contribution of the term $\\sim F$ can be neglected for the extremum condition. Extended scaling symmetry admits quasistatic solutions only in the flat phase. Sect. \\ref{adjustment} deals with the issue of an ``adjustment'' or tuning'' of the cosmological constant. The robustness of $\\Lambda=0$ with respect to changes of parameters in the dilatation symmetric effective action is associated to a mechanism of ``self-adjustment'' or ``self-tuning''. In this respect we highlight the differences between a four dimensional theory with a finite number of scalar fields and a higher dimensional setting with infinitely many effective four dimensional scalar fields. This is an important ingredient for the understanding of the robustness of the flat phase, which would be hard to implement with a finite number of degrees of freedom. Sect. \\ref{higherorder} investigates the dilatation symmetric effective action \\eqref{AA1a} with a polynomial form of $F$. We display the field equations and discuss simple solutions with a vanishing four-dimensional cosmological constant $\\Lambda$. An appendix is devoted to particular solutions with non-Ricci-flat internal space, which nevertheless results in $\\Lambda=0$. These solutions demonstrate explictly that non-abelian isometries of internal space are possible for extrema in the flat phase. The final part presents a simple estimate of the dilatation anomaly in sect. \\ref{dilatationanomaly}. In sect. \\ref{cosmologicalrunaway} we show that this leads to cosmological runaway solutions for which $\\Lambda$ vanishes asymptotically, while for finite $t$ a homogeneous dark energy component accounts for quintessence. We present our conclusions in sect. \\ref{conclusions}. ", "conclusions": "\\label{conclusions} In general, the quantum effective action $\\Gamma[\\hat g_{\\hat\\mu\\hat\\nu},\\xi]$ for gravity coupled to a scalar field $\\xi$ has a complicated form. We explore here the possibility of a simple scaling limit for large $\\xi$ and small $\\hat g_{\\hat\\mu\\hat\\nu}$. More precisely, we investigate the scaling with a multiplicative power of a dimensionless parameter $\\kappa$ and define \\begin{equation}\\label{Z1} \\Gamma_\\kappa[\\hat g_{\\hat\\mu\\hat\\nu},\\xi]=\\Gamma[\\kappa^{-2}\\hat g_{\\hat\\mu\\hat\\nu},\\kappa^{\\frac{d-2}{2}}\\xi]. \\end{equation} Our hypothesis is a simple limiting behavior of $\\Gamma_\\kappa$ for $\\kappa\\to\\infty$. In particular, we investigate a possible fixed point \\cite{CWQ}, \\cite{CWNL} \\begin{eqnarray}\\label{Z2} \\lim_{\\kappa\\to\\infty}\\Gamma_\\kappa=\\int_{\\hat x}\\hat g^{1/2} \\left\\{-\\frac12\\xi^2\\hat R+\\frac{\\zeta}{2} \\partial^{\\hat \\mu}\\xi\\partial_{\\hat \\mu}\\xi\\right\\}. \\end{eqnarray} As appropriate for a fixed point, eq. \\eqref{Z2} states that $\\Gamma_\\kappa$ becomes independent of $\\kappa$ for $\\kappa\\to\\infty$. In this limit the effective action is therefore invariant with respect to dilatations. If a fixed point \\eqref{Z2} exists, it may be used for a definition of quantum gravity, with a non-perturbative renormalization similar to the asymptotic safety scenario \\cite{25}, \\cite{26}. Quantum fluctuations are expected to modify the simple asymptotic form of the effective action by adding for finite $\\kappa$ additional terms to the r.h.s. of eq. \\eqref{Z2}. Such terms typically violate the dilatation symmetry and will then be treated as ``dilatation anomalies''. An example would be a cosmological constant \\begin{equation}\\label{Z3} \\Gamma^{(\\bar\\mu)}=\\int_{\\hat x} \\hat g^{1/2}\\bar\\mu^d. \\end{equation} Its contribution to $\\Gamma_\\kappa$ indeed vanishes for $\\kappa\\to\\infty$ \\begin{equation}\\label{Z4} \\Gamma^{(\\bar\\mu)}_\\kappa=\\kappa^{-d}\\int_{\\hat x}\\hat g^{1/2}\\bar\\mu^d. \\end{equation} One may imagine a non-linear functional flow equation for the $\\kappa$-dependence of $\\Gamma_\\kappa$ (at fixed $\\hat g_{\\hat\\mu\\hat\\nu},\\xi$ and besides the trivial linear one following from the definition \\eqref{Z1}), \\begin{equation}\\label{Z5} \\kappa\\partial_\\kappa\\Gamma_\\kappa[\\hat g_{\\hat\\mu\\hat\\nu},\\xi]= {\\cal F}[\\hat g_{\\hat\\mu\\hat\\nu},\\xi], \\end{equation} with a functional ${\\cal F}$ typically involving $\\Gamma_\\kappa$ and its functional derivatives. One could then define quantum gravity by a solution of this flow equation with the ``boundary condition'' or ``initial value'' \\eqref{Z2}. The fixed point would be an ``ultraviolet fixed point'' in the sense of $\\kappa\\to\\infty$. Effective actions which are close to this fixed point for finite $\\kappa$ may then be treated in the standard renormalization group formalism for relevant and marginal directions. The fixed point may become important for late time cosmological solutions. This happens if $\\xi(t)$ increases and $\\hat g^{1/2}(t)$ decreases for increasing time, such that the field equations, which obtain from the functional derivatives of $\\Gamma$ with respect to $\\hat g_{\\hat\\mu\\hat\\nu}$ and $\\xi$, have to be evaluated for values of the fields where the limit $\\Gamma_{\\kappa\\to\\infty}$ becomes relevant. In sect. \\ref{cosmologicalrunaway} we have discussed an explicit example for this type of ``runaway cosmology''. The approach to the fixed point leads to an interesting candidate for dynamical dark energy, where the potential energy of an appropriate scalar field vanishes only asymptotically for $t\\to\\infty$. The fixed point \\eqref{Z2} may exist for arbitrary dimension $d>2$, including $d=4$. In this paper we are interested in higher dimensional theories. Indeed, for $d>6$ no polynomial potential for $\\xi$ is consistent with dilatation symmetry. This may add to the plausibility of a fixed point effective action without a potential $V(\\xi)$. On the other hand, the vanishing of the higher dimensional cosmological constant is, in general, not sufficient to guarantee a vanishing effective four-dimensional cosmological constant after ``spontaneous compactification'' of the internal dimensions. The curvature of internal space or a non-trivial warping may generate such an effective cosmological constant. One of the important findings of the present paper is a statement about the most general ``compactified'' higher dimensional solutions of the field equations derived from the fixed point action \\eqref{Z2}: all stable solutions with maximal four-dimensional symmetry and finite effective four-dimensional gravitational constant must have a vanishing four-dimensional cosmological constant. This asymptotic vanishing of the cosmological constant can be understood by a self-adjustment mechanism for infinitely many degrees of freedom, as discussed in sect. \\ref{adjustment}. Consider now stable cosmological runaway solutions which drive the fields into a region where the fixed point effective action \\eqref{Z2} becomes relevant, and which allow for an effective four-dimensional gravity. Such solutions lead to a vanishing four-dimensional cosmological constant for $t\\to\\infty$, thus solving the ``cosmological constant problem''. Furthermore, since the Universe is very old (in units of the Planck time), but not infinitely old, a small amount of homogeneous energy density remains present in the effective four-dimensional Universe. This could give an explanation for the observed dark energy. Due to the huge age of the Universe (in Planck units) we expect that the present geometry of the Universe is very close to the asymptotic geometry, which is a quasi static solution of the field equations derived from eq. \\eqref{Z2}. It is easy to find candidates for such asymptotic solutions - the simplest being a direct product of four-dimensional flat space and an internal $d-4$-dimensional Ricci-flat space with finite volume, accompanied by a constant value of the scalar field $\\xi$. If internal space admits isometries, the gauge couplings of the resulting four-dimensional gauge interactions become time-independent for large $t$. This solution has a vanishing four-dimensional cosmological constant $\\Lambda$ and exists for all values of $\\zeta$ in eq. \\eqref{Z2}. On the other hand, we have shown that all possible stable quasistatic solutions of the field equations derived from the effective action \\eqref{Z2}, which are consistent with four dimensional gravity, must have a vanishing four-dimensional constant $\\Lambda=0$. Stable asymptotic solutions approaching de Sitter or anti-de Sitter space $(\\Lambda\\neq 0)$ are not possible. This singles out a vanishing cosmological constant for the asymptotic behavior for $t\\to\\infty$ and may be the basis for the solution of the ``cosmological constant problem''. Furthermore, we have extended our discussion to a more general form of a fixed point effective action. It includes dilatation symmetric higher curvature invariants for the metric. In the absence of a potential for $\\xi$ our conclusions remain essentially unchanged. Finally, we have presented a simple estimate of the dilatation anomaly and discussed the qualitative behavior of the corresponding runaway cosmology. It approaches a zero cosmological constant for infinite time, and has a dynamical dark energy component of similar size as dark matter. Our main conclusion is simple: if the quantum effective action for large $\\xi$ and small $\\hat g^{1/2}$ shows a limiting fixed point behavior \\eqref{Z2}, or a similar dilatation symmetric behavior without a potential term for $\\xi$, and if a stable cosmological runaway solution approaches this field region for large $t$, the cosmological constant problem can be solved without any fine tuning of parameters or initial conditions. \\LARGE" }, "0911/0911.1549_arXiv.txt": { "abstract": "The theory of radiative transfer provides the link between the physical conditions in an astrophysical object and the observable radiation which it emits. Thus accurately modelling radiative transfer is often a necessary part of testing theoretical models by comparison with observations. We describe a new radiative transfer code which employs Monte Carlo methods for the numerical simulation of radiation transport in expanding media. We discuss the application of this code to the calculation of synthetic spectra and light curves for a Type Ia supernova explosion model and describe the sensitivity of the results to certain approximations made in the simulations. ", "introduction": "Since almost all we know about astronomical objects is inferred from the radiation which they emit, the theory of radiation transport often has a key role in testing our understanding of astrophysics. Although there are a variety of competitive approaches used for radiative transfer simulations, Monte Carlo methods are particularly well-suited for many modern astrophysical applications. In the Monte Carlo approach, the radiation field is discretized into quanta which represent bundles of photons. By propagating these quanta through a model of an astrophysical object, and simulating their interactions, synthetic spectra and light curves can be obtained. This method has the particular advantage that matter-radiation interactions are always treated locally meaning that multi-dimensionality, time-dependence and large-scale velocity fields can all be incorporated readily. ", "conclusions": "" }, "0911/0911.0980_arXiv.txt": { "abstract": "We simulate the formation and evolution of galaxies with a self-consistent 3D hydrodynamical model including star formation, supernova feedback, and chemical enrichment. Hypernova feedback plays an essential role not only in solving the [Zn/Fe] problem, but also reproducing the cosmic star formation rate history and the mass-metallicity relations. In a Milky-Way type galaxy, kinematics and chemical abundances are different in bulge, disk, and thick disk because of different star formation histories and the contribution of Type Ia Supernovae. ", "introduction": "While the evolution of the dark matter is reasonably well understood, the evolution of the baryonic component is much less certain because of the complexity of the relevant physical processes, such as star formation and feedback. With the commonly employed, schematic star formation criteria alone, the predicted star formation rates (SFRs) are higher than what is compatible with the observed luminosity density. Thus feedback mechanisms are in general invoked to reheat gas and suppress star formation. Supernovae inject not only thermal energy but also heavy elements into the interstellar medium, which can enhance star formation. Chemical enrichment must be solved as well as energy feedback. ``Feedback'' is also important for solving the angular momentum problem and the missing satellite problem, and for explaining the existence of heavy elements in intracluster medium and intergalactic medium, and the mass-metallicity relation of galaxies (\\cite{kob07}, hereafter K07). In the next decade, high-resolution multi-object spectroscopy (HERMES) and space astrometry mission (GAIA) will provide kinematics and chemical abundances of a million stars in the Local Group. Since different heavy elements are produced from different supernovae with different timescales, elemental abundance ratios can provide independent information on ``age''. Therefore, stars in a galaxy are fossils to untangle the history of the galaxy. The galactic archeology technique can be used to study the galaxy formation and evolution in general. Metallicities are measured in various objects with different galaxy mass scale and as a function of redshift/time. The internal structure of galaxies have being observed with integral field spectrographs (e.g., the SAURON project, SINFONI on VLT). Chemodynamical simulations can provide useful predictions and physical interpretation of these observations. \\vspace*{-5mm} ", "conclusions": "" }, "0911/0911.2727_arXiv.txt": { "abstract": "{ In recent years it has emerged that the high energy behavior of gravity could be governed by an ultraviolet non-Gaussian fixed point of the (dimensionless) Newton's constant, whose behavior at high energy is thus {\\it antiscreened}. This phenomenon has several astrophysical implications. In particular in this article recent works on renormalization group improved cosmologies based upon a renormalization group trajectory of Quantum Einstein Gravity with realistic parameter values will be reviewed. It will be argued that quantum effects can account for the entire entropy of the present Universe in the massless sector and give rise to a phase of inflationary expansion. Moreover the prediction for the final state of the black hole evaporation is a Planck size remnant which is formed in an infinite time. } \\FullConference{Workshop on Continuum and Lattice Approaches to Quantum Gravity\\\\ September 17-19 2008\\\\ Brighton, University of Sussex, United Kingdom} \\begin{document} ", "introduction": "Cosmology is a natural setting to study quantum gravity, which may provide answers to fundamental questions as why is the expansion of the universe isotropic, can the initial singularity be avoided, why does the vacuum energy ``gravitate'' so little (Cosmological Constant problem)? In recent years it has emerged that the asymptotic safety scenario \\cite{wein2,2009AnPhy.324..414C,niedermaier_asymptotic_2006} could provide the right framework to address the above questions. According to this approach the ultraviolet (UV) behavior of quantum gravity is controlled by a fixed point at a non-zero value of the (dimensionless) coupling constant, so that the dimensionful Newton's constant reduces its strength at higher energies, it is thus {\\it antiscreened}. The non-perturbative renormalization group (RG) equation employed in this investigation predicts that the dimensionless cosmological constant reaches a non-gaussian fixed point (NGFP) in the infinite cutoff limit, so that the full Einstein-Hilbert Lagrangian is renormalizable at a non-perturbative level around this fixed point. The gravitational {antiscreening} behavior is very similar to the running of the non-Abelian gauge coupling in Yang-Mills Theory, but only after the introduction of the effective average action and its functional renormalization group equation for gravity \\cite{1998PhRvD..57..971R} detailed investigations of the scaling behavior of the Newtons's constant have become possible \\cite{1998PhRvD..57..971R,1998CQGra..15.3449D,2002PhRvD..65b5013L,2002PhRvD..66b5026L,2002CQGra..19..483L, 2002PhRvD..65f5016R,2002PhRvD..66l5001R,1999PThPh.102..181S,2003PhRvD..67h1503P,2004PhRvL..92t1301L, 2005JHEP...02..035B,2006PhRvL..97v1301C,2009GReGr..41..983R,2009PhRvD..79j5005R,2009arXiv0904.2510M,2009arXiv0905.4220M}. The non-perturbative renormalization group equation underlying this approach defines a Wilsonian RG flow on a theory space which consists of all diffeomorphism invariant functionals of the metric $g_{\\mu\\nu}$. This framework turned out to be an ideal setting for investigating the asymptotic safety scenario in gravity \\cite{wein2,2009AnPhy.324..414C,niedermaier_asymptotic_2006} and, in fact, substantial evidence was found for the non-perturbative renormalizability of Quantum Einstein Gravity. The theory emerging from this construction (``QEG\") is not a quantization of classical general relativity. Instead, its bare action corresponds to a nontrivial fixed point of the RG flow and therefore is a {\\it prediction}. The effective average action \\cite{1998PhRvD..57..971R,2002PhR...363..223B} has crucial advantages as compared to other continuum implementations of the Wilson RG, in particular it is closely related to the standard effective action and defines a family of effective field theories $\\{ \\Gamma_k[g_{\\mu\\nu}], 0 \\leq k < \\infty \\}$ labeled by the coarse graining scale $k$. The latter property opens the door to a rather direct extraction of physical information from the RG flow, at least in single-scale cases: If the physical process or phenomenon under consideration involves only a single typical momentum scale $p_0$ it can be described by a tree-level evaluation of $\\Gamma_k[g_{\\mu\\nu}]$, with $k=p_0$. The precision which can be achieved by this effective field theory description depends on the size of the fluctuations relative to the mean values. If they are large, or if more than one scale is involved, it might be necessary to go beyond the tree analysis. The qualitative scale dependence of Newton's constant can be grasped with the help of the following physical argument. Let us imagine that in the large distance limit the leading quantum effects of the geometry are described by quantizing the linear fluctuations of the metric, $g_{\\mu \\nu}$. The resulting theory is a minimallu coupled theory in a curved background spacetime whose elementary quanta, the gravitons, carry energy and momentum. The vacuum of this theory will be populated by virtual graviton pairs, and the problem is to understand how these virtual gravitons respond to the perturbation by an external test body which we immerse in the vacuum. Assuming that also in this situation gravity is universally attractive, the gravitons will be attracted towards the test body. It will thus become ``dressed\" by a cloud of virtual gravitons surrounding it so that its effective mass seen by a distant observer is larger than it would be in absence of any quantum effects. This means that while in QED the quantum fluctuations {\\it screen} external charges, in quantum gravity they have an {\\it antiscreening} effect on external test masses. The consequence of this simple {\\it Gedanken} experiment entails Newton's constant becoming a scale dependent quantity $ G\\!\\left(k \\right)$ which is small at small distances $ r\\sim 1/k$, and which becomes large at larger distances. In QED the {screening} behavior is well-known but it is interesting to recall how this result is obtained from the \"renormalization group improvement\", a standard device, in particle physics, in order to add the dominant quantum corrections to the Born approximation of a scattering cross section for instance. One starts from the classical potential energy $V_{\\rm cl}(r) = e^2/4\\pi r$ and replaces $e^2$ by the running gauge coupling in the one-loop approximation: \\begin{equation}\\label{uel} e^2(k)=e^2(k_0)[1-b\\ln(k/k_0)]^{-1},\\;\\;\\;\\;\\; b\\equiv e^2(k_0)/6\\pi^2. \\end{equation} The crucial step is to identify the renormalization point $k$ with the inverse of the distance $r$ so that result of this substitution reads \\be\\label{rea} V(r)=-e^2(r_0^{-1})[1+b\\ln(r_0/r) + O(e^4)]/4\\pi r \\ee where the IR reference scale $r_0\\equiv 1/k_0$ has to be kept finite in the massless theory. We emphasize that eq.(\\ref{rea}) is the correct (one-loop, massless) Uehling potential which is usually derived by more conventional perturbative methods \\cite{dittrich_effective_1985}. Obviously the position dependent renormalization group improvement $e^2\\rightarrow e^2(k)$, $k\\propto 1/r$ encapsulates the most important effects which the quantum fluctuations have on the electric field produced by a point charge. The effective field theory techniques proved useful for an understanding of the scale dependent geometry of the effective QEG spacetimes \\cite{2005JHEP...10..050L,2006JHEP...01..070R,2007JHEP...01..049R}. In particular it has been shown \\cite{2002PhRvD..65b5013L,2005JHEP...10..050L} that these spacetimes have fractal properties, with a fractal dimension of 2 at small, and 4 at large distances. The same dynamical dimensional reduction was also observed in numerical studies of Lorentzian dynamical triangulations \\cite{2005PhLB..607..205A,2004PhRvL..93m1301A,2005PhRvL..95q1301A} and in \\cite{2006JHEP...11..081C} A.Connes et al. speculated about its possible relevance to the non-commutative geometry of the standard model. In order to extract all the relevant information from the RG evolution, it is thus necessary to relate the cutoff scale $k$ which corresponds to the resolution of the RG flow, to the spacetime properties. This procedure is called ``cutoff identification\" for which the relevant energy scale $k$ is related to a characteristic length scale where the quanta with energy $k$ propagate. In the case of massless QED the choice $k\\propto 1/r$ was clearly the only possible one, as there are no other relevant scales in the problem. When several scales are present the prescription which emerges from the general theory of the Effective Average Action \\cite{2002PhR...363..223B} is that $\\Gamma_k$ is defined at a scale $k$ which is the {\\it largest} one of the various competing scales in the fluctuation determinant of the Average Action, $\\Gamma^{(2)}_k$, namely \\be\\label{flucd} \\Gamma^{(2)}_k=\\frac{\\delta^2 \\Gamma_k}{\\delta \\Phi^2 } \\ee where $\\Phi$ is the so-called ``blocked'' field \\cite{1995PhRvD..52..969B}. The difficulty arises when we decide to apply the same ``recipe'' in gravity by writing \\be\\label{klen} k \\sim 1/\\ell(x^\\mu)), \\;\\;\\;\\;\\;\\;\\;\\;\\;\\;\\; \\ell=\\ell(g_{\\mu\\nu}) \\ee being $\\ell$ a characteristic length where the fluctuations with energy $k$ propagate. The reason is that the flow equation is by construction diffeomorphism invariant at any $k$ so that the RG flow itself does not know anything about the background field metric $g_{\\mu\\nu}$ that has been used for projecting on a finite-dimensional subspace of the ``theory space''. There are two possible strategies to overcome this issue. The first one amounts to choose a fiducial metric which is a {\\it solution} of the Einstein equations and RG-improve it by substituting the Newton constant $G$ with the running $G(k)$ together with a cutoff identification of the type (\\ref{klen}). The limitation of this approach lies in the fact that in general the improved metric may not be a solution of Einstein equation, but one can imagine that this is a sort of ``Thomas-Fermi'' approximation where only the leading quantum corrections are taken into account \\cite{2000PhRvD..62d3008B,2006PhRvD..73h3005B}. The improved $g_{\\mu\\nu}(k)$ metric represents then a sort of ``emergent'' spacetime description of the effective geometry \\cite{2005LRR.....8...12B,2004CQGra..21.1725M,2007arXiv0712.0810V} according to the scale dependence of the Newton constant. A second possibility is to consider the energy scale $k$ associated to the field strength itself rather than to an observational scale $\\ell$. This is motivated by the analogy with the QED (and QCD) case, where higher loop contributions to the Uehling potential are obtained by renormalization group improvement of the QED {\\it action} by using the field strength $(F_{\\mu\\nu}F^{\\mu\\nu})^{1/4}$ as a cutoff instead than $1/r$ \\cite{1973NuPhB..52..483M,1978NuPhB.134..539M,1978NuPhB.143..485P}. In this case the short distance correction to the static potential is obtained from the non-linear differential equations \\begin{eqnarray}\\label{maty} &&\\nabla \\cdot \\mathbf D = J_0, \\;\\;\\;\\;\\; \\mathbf D = \\mathbf E \\; \\epsilon (E) , \\;\\;\\; \\mathbf E = -\\nabla A^0\\\\[2mm]\\nonumber && \\epsilon(E)=1-\\frac{e^2}{12\\pi^2} \\log (e E/k_0^2)+\\ldots \\end{eqnarray} whose solution reproduces the Uehling potential in the long distance limit, but in general the solutions of Eq.(\\ref{maty}) include higher loop effects due to the non-linearities of the effective action in the short distance limit. The two approaches discussed above are obviously related, at least in some limit. In the case of Robertson-Walker spaces it will be shown that due to the very high degree of symmetry of the spacetime, the time-scale defined by ``Hubble parameter'' behaves essentially like the characteristic time scale associated to the relevant curvature invariants\\cite{2004ForPh..52..650R,2005JCAP...09..012R,2007JCAP...08..024B}. In the case of spherically symmetric spacetimes \\cite{2000PhRvD..62d3008B,2004JCAP...12..001R,2004PhRvD..70l4028R,2006PhRvD..73h3005B}, near the singularity the proper distance of a radially free falling observer behaves essentially as $1/\\sqrt{\\Psi_2}$, being $\\Psi_2$ the ``Coulombian'' component of the Weyl tensor. From the above discussion it is then clear that in general there is not a preferred strategy to perform the RG improvement in gravity. In some case it might be more interesting to RG improve {\\it solutions} and to make contact with an emergent spacetime description of the effective geometry. In some other cases it could be more convenient to work with a RG improvement at the level of {\\it field equations} or {\\it actions} \\cite{2004PhRvD..69j4022R,2004PhRvD..70l4028R,2004CQGra..21.5005B,2005IJMPA..20.2358B}. It is important to remark that it is not surprising that different cutoff might provide quantitavely different evolutions, as usually the $\\beta$-functions are not ``universal'' quantities. For instance it is well known \\cite{Goldenfeld:1992qy} that different realizations of the block-spin RG transformation applied to the Ising model may provide different values for fixed points, as we are essentially using different type of ``microscopes''\\footnote{From this point of view the criticism expressed in \\cite{2008PhRvL.101h1301W} should not be seriously considered.}. On the other hand truly universal quantities, like the critical exponents, are essentially insensitive to the cutoff choice. In this review recent results obtained with the RG improvement of Einstein theory will be discussed in the framework QEG will be reviewed, with particular emphasis on recent results obtained in cosmology \\cite{2007JCAP...08..024B}. In particular in Sec.2 the RG evolution of the Newton constant and Cosmological constant describing our Universe are described. In Sec.3 a covariant formalism to improve the Einstein field equation is presented while in Sec.4 the RG improved Robertson-Walker Cosmology is discussed. In Sec.5 the basic mechanism to produce the entropy of the Universe is presented. In Sec.6 the properties of a class of solutions of the RG equations are discussed. In Sec.7 a mechanism to produce a power-law inflation is studied. In Sec.8 and Sec.9 the properties of RG improved Black Hole metric is studied. In Sec.10 the possibility that Quantum Gravity effects are present on Astrophysical distances is reviewed. Sec.11 is devoted to the Conclusions. ", "conclusions": "In these notes some important astrophysical consequences of the Asymptotic Safety Scenario have been reviewed. In particular it was advocated the point of view that the scale dependence of the gravitational parameters has an impact on the physics of the Universe we live in and, in particular, it has been possible to identify known features of the Universe which could possibly be due to this scale dependence. Three possible candidates for such features are proposed: the entropy carried by the radiation which fills the Universe today, a period of automatic, $\\Lambda$-driven inflation that requires no ad hoc inflaton, and the primordial density perturbations. Moreover, the impact of the leading quantum gravity effects on the dynamics of the Hawking evaporation process of a black hole have also been investigated. Its spacetime structure is described by a renormalization group improved Vaidya metric. Its event horizon, apparent horizon, and timelike limit surface have been obtained taking the scale dependence of Newton's constant into account. The emergence of a quantum ergosphere is discussed. The final state of the evaporation process is a cold, Planck size remnant. It would be interesting to investigate the possible astrophysical implications of a population of stable Planck size mini-black holes produced in the Early Universe or by the interaction of cosmic rays with the interstellar medium. I hope to address this issue in a subsequent publication." }, "0911/0911.2511_arXiv.txt": { "abstract": "We compare H$\\alpha$, ultraviolet (UV) and infrared (IR) indicators of star formation rate (SFR) for a well-defined sample of $z=0.84$ emission-line galaxies from the High-$z$ Emission Line Survey (HiZELS). Using emission-line, optical, IR, radio and X-ray diagnostics, we estimate that 5 -- 11~per~cent of H$\\alpha$ emitters at this redshift are active galactic nuclei. We detect 35~per~cent of the H$\\alpha$ emitters individually at 24~$\\mu$m, and stack the locations of star-forming emitters on deep 24-$\\mu$m {\\it Spitzer Space Telescope} images in order to calculate the typical SFRs of our H$\\alpha$-emitting galaxies. These are compared to the observed H$\\alpha$ line fluxes in order to estimate the extinction at $z=0.84$, and we find a significant increase in dust extinction for galaxies with higher SFRs. We demonstrate that the relationship between SFR and extinction found in the local Universe is also suitable for our high-redshift galaxies, and attribute the overall increase in the typical dust extinction for $z=0.84$ galaxies to an increase in the average SFR, rather than to a change in dust properties at higher redshift. We calculate the UV extinction, and find a similar dependence on SFR to the H$\\alpha$ results, but no evidence for a 2175~\\AA\\ UV bump in the dust attenuation law for high-redshift star-forming galaxies. By comparing H$\\alpha$ and UV indicators, we calculate the conversion between the dust attenuation of nebular and stellar radiation, $\\gamma$, and show that $\\gamma=0.50\\pm0.14$. The extinction / SFR relationship is shown to be applicable to galaxies with a range of morphologies and bulge-to-disk ratios, to both merging and non-merging galaxies, and to galaxies within high- and low-density environments, implying that it is a fundamental property of star-forming regions. In order to allow future studies to easily correct for a SFR-dependent amount of dust extinction, we present an equation to predict the extinction of a galaxy, based solely upon its observed H$\\alpha$ luminosity, and use this to recalculate the H$\\alpha$ luminosity function and star formation rate density at $z=0.84$. ", "introduction": "\\label{sec:introduction} There are many methods of estimating the global star formation rate (SFR) of a galaxy, from measurements of the ultra-violet (UV) emission coming directly from visible young massive stars, to the bolometric infrared (IR) luminosity resulting from dust-reprocessed stellar emission \\citep[see][for a review of various star formation indicators]{Kennicutt98}. All SFR indicators suffer from different biases and uncertainties, and while SFRs estimated in different ways in the local Universe are now in general agreement, there is much greater scatter in the results at $z\\sim1$ \\citep[see, e.g.][]{Hopkins06}. Locally one of the best measures of the current rate of star formation within a galaxy comes from observations of the H$\\alpha$ transition of atomic hydrogen. The luminosity of this recombination line scales directly with the rate of ionising flux from young, massive stars and gives an excellent indication of the presence of ongoing star formation. However, its usefulness as a quantitative SFR indicator is reduced by the effect of dust obscuration; observed H$\\alpha$ line fluxes need to be corrected for extinction before the intrinsic H$\\alpha$ luminosity can be recovered, and the SFR calculated. The initial method used to estimate H$\\alpha$ extinction came from a comparison of thermal radio emission, which is unaffected by dust attenuation, and the observed H$\\alpha$ + [N\\,{\\sc ii}] line emission within local galaxies \\citep{Kennicutt83}. Disentangling thermal and non-thermal radio emission is non-trivial \\citep[see, e.g.][for more details]{Condon92}, as thermal radio emission is only a small fraction of the total radio luminosity of a galaxy \\citep[$\\lesssim10$~per~cent at 1.4~GHz;][]{Condon90}, and so this method is not often used in practise. A much more common method to estimate extinction in individual galaxies comes from the `Balmer decrement', comparing the observed value of the H$\\alpha$/H$\\beta$ line flux ratio to the intrinsic value when no dust is present. This allows a calculation of the reddening between the wavelengths of the H$\\alpha$ and H$\\beta$ lines and, with a choice of dust attenuation model \\citep[e.g.][]{Calzetti00}, permits an estimate of the attenuation at a given wavelength \\citep*[see e.g.][]{Moustakas06}. However, this technique requires sensitive spectroscopy to obtain the line flux ratio with a good signal-to-noise ratio (SNR), which is time-consuming for large samples of galaxies at moderate redshifts. It is therefore more usual for studies with limited available photometry to simply assume that a single extinction can be applied to all galaxies. The canonical amount of H$\\alpha$ extinction for a typical galaxy is usually taken to be $A_{\\rm H\\alpha} \\sim1$~mag \\citep[e.g.][]{Kennicutt83,Bell01, Pascual01,Charlot02,Fujita03,Brinchmann04,Ly07,Geach08,Sobral09Halpha}, based upon measurements of extinction in the local Universe. At higher redshift, the appropriate H$\\alpha$ extinction for a typical galaxy is not well determined. Significant numbers of extremely dusty galaxies have been discovered at high redshift \\citep[e.g.][]{Chapman05}, and the large dust content implies a greater H$\\alpha$ attenuation. The global rate of star formation increases out to $z\\sim1$ \\citep[e.g.][]{Hopkins06}, and galaxies with higher SFR are thought to undergo greater extinction \\citep{Hopkins01, Sullivan01, Berta03}, also implying that the typical H$\\alpha$ attenuation will be greater at $z\\sim1$ than in the local Universe. \\citet{Villar08} have estimated that the typical H$\\alpha$ extinction of galaxies within their $z=0.84$ sample is $\\sim1.5$~mag. The total IR luminosity of a galaxy is a bolometric measurement of the thermal emission from dust grains, which have been heated by the UV and optical radiation produced by young massive stars \\citep[e.g.][]{Kennicutt98}. For star-forming galaxies with a reasonable dust content, this emission gives a direct measurement of the total stellar power output. However, since longer wavelength IR emission will more closely trace the cool dust component within a galaxy, which predominantly results from heating by an old stellar component rather than from recent star formation, it could be argued that mid-IR emission might be a better SFR indicator than the bolometric IR luminosity \\citep[e.g.][]{Dopita02}. \\citet{Rieke09} have demonstrated that the flux density measured in the {\\it Spitzer Space Telescope} 24-$\\mu$m band can be used as a good proxy for SFR for galaxies at $z=0$, and, after applying suitable corrections, has a comparable accuracy to estimating SFRs from the extinction-corrected Pa$\\alpha$ luminosity or total IR luminosity for normal galaxies in the redshift range $0L^*$ luminosities (eg. Skelton, Bell and Somerville 2009). Age also has a (secondary) role in explaining the colour trend - the most massive E/S0s formed early and rapidly (in $<1$ Gyr), while those of lower luminosity formed stars over longer timescales, giving slightly younger flux-weighted ages. There is also evidence that E/S0 galaxy formation occurred slightly earlier in dense environments (eg. Sheth et al. 2006, Bernardi 2009), while galaxies in cluster centres may have had an unusual and extreme merger history (e.g. Boylan-Kolchin, Ma and Quataert 2006). Thus merger history and cluster environment may both influence the CMR. Our studies of the CMR in Roche, Bernardi and Hyde 2009 (hereafter Paper I) suggested (see Section 3.1) that the radial colour gradients in E/S0 galaxies might be correlated with their other properties and it would be fruitful to investigate this further. These colour gradients are of interest in that they are sensitive tracers of evolution processes and vary strongly between the individual E/S0s. Most E/S0 galaxies have negative colour gradients, meaning that the centres are reddest and colours become bluer outwards. This could be the result of a negative gradient in age, metallicity or dust, or some combination of these. Tamura et al. (2000) and Tamura and Ohta (2003) estimate the mean colour gradient in E/S0s as ${{\\Delta(B-R)}\\over{\\Delta(\\rm log~r)}}=-0.09$ mag $\\rm dex^{-1}$ (with a scatter of 0.04), and attributed this to a radial gradient of metallicity (rather than age) ${{d(\\rm log~Z)}\\over{d(\\rm log~r)}}= -0.3\\pm 0.1$, approximately constant to $z\\simeq 1$. Wu et al. (2005) similarly estimated ${{d({\\rm log}~Z)}\\over{d(\\rm log~r)}}= -0.25$ with a large scatter $\\sigma=0.19$. Mehlert et al. (2003) found negative metallicity gradients in Coma-cluster ellipticals, but negligible gradients in age or $\\alpha$/Fe. Michard (2005) concluded colour gradients in 50 nearby ellipticals were primarily due to metallicity gradients and the effects of dust were usually small. La Barbera and Carvalho (2009) compare optical and near-IR colour gradients and again conclude they are produced by negative gradients in metallicity (they find the radial age gradients in E/S0s may be even be positive). Kobayashi (2004) performed detailed chemodynamical simulations of a set of over 100 model elliptical galaxies, with star-formation and merging, and predicted that spheroidals which formed monolithically, or at least with only minor mergers, would have steeper metallicity gradients (${{d({\\rm log}~Z)}\\over{d(\\rm log~r)}}\\simeq -0.3$ to -0.5) than those formed by major mergers (${{d({\\rm log}~Z)}\\over{d(\\rm log~r)}}\\simeq -0.2)$. Within the simulated galaxy set, merger history was the primary determinant (more so than mass or luminosity) of present-day metallicity/colour gradient. This might account for the observed wide distribution in colour gradients if similar numbers of spheroidals had monolithic, minor-merger or major-merger histories, and might also lead to an environmental dependence. In this paper we focus especially on Brightest Cluster Galaxies (BCGs), of which several thousand have been identified in the SDSS. Previously, the BCGs in the SDSS have been found to follow a steeper relation of radius to luminosity (${\\rm r}_{eff}\\propto L$) than the other E/S0s (${\\rm r}_{eff}\\propto L ^{0.6}$); (Bernardi et al. 2007; Bernardi 2009). In this paper, we look for systematic differences in their CMR and radial colour gradients with respect to other E/S0 galaxies, especially those in the same luminosity range ($M_r<-22.5$ to $M_r\\simeq -25$). Some models, for example, predict a much flatter CMR for the BCGs, as a result of formation from a relatively large number of mergers of (already) old and red galaxies (de Lucia and Blaizot 2007). Ko and Im (2005) found indications that the colour gradients of ellipticals in dense cluster environments tended to be less strong than in field environments, which again could be due to (more) mergers. In Section 2 of this paper we describe the data used and our selection of early-type galaxies and BCGs. In Sections 3 and 4 we investigate and compare their CMR and colour-$\\sigma$ relation. In Section 5 we use an estimator of colour gradient based on the ratio of $g$ and $r$-band effective radii to compare the colour gradients of BCGs and other E/S0s and examine the dependences on luminosity, radius, and density. In Section 6 we look at the influence of stellar age, as estimated from the spectra. We conclude in Section 7 with summary and further discussion. SDSS magnitudes are given in the AB system where $m_{AB}=-48.60-2.5$ log $F_{\\nu}$ (in ergs $\\rm cm^{-2}s^{-1}Hz^{-1}$); equivalently, $m_{AB}=0$ is 3631 Jy. We assume throughout a spatially flat cosmology with $H_0=70$ km $\\rm s^{-1}Mpc^{-1}$, $\\Omega_{M}=0.27$ and $\\Omega_{\\Lambda}=0.73$, giving the age of the Universe as 13.88 Gyr. ", "conclusions": "(i) The colour-magnitude relation (CMR) and colour-velocity dispersion relation (C$\\sigma$R) of BCGs are significantly flatter in slope(${{d(g-r)}\\over{d M_r}}$ and ${{d(g-r)}\\over{d{\\rm log}~\\sigma}}$) than the respective relations for non-BCG E/S0 galaxies of comparable (high) luminosity. The difference in slope is about a factor of two, whether we define the $g-r$ colour using k-corrected model magnitudes or rest-frame magnitudes derived from integrating the SDSS spectra. BCGs are also, on average, $\\sim 0.01$ mag redder in model magnitudes (k-corrected to rest-frame) than E/S0s of the same $M_r$ or $\\sigma$, but with less or no offset in spectra-derived colours. The hierarchical merging model of Lucia and Blaizot (2007) predicts a flat CMR for BCGs as a result of late assembly from a large number of red progenitors which formed their stars very much earlier ($z\\sim 5$). This model included a very early truncation of star-formation. Skelton, Bell and Somerville (2009) present a simplified model in which dry mergers of already merged E/S0s on a `creation red sequence' mildly flatten the CMR slope at higher luminosities. There may indeed be some flattening in the bright end of the CMR for all E/S0s (see Paper I), but for BCGs we find much more, although the BCG CMR is not entirely flat. This suggests the BCGs or their progenitors evolve by a less extreme scenario than the first of these models but perhaps with a higher dry merger rate compared to the E/S0s in the second model. In addition to being flattened, the BCG CMR is offset redwards. \\medskip (ii) As a simple quantifier of radial colour gradient we use ${{{\\rm r}_{eff}(g)}\\over{{\\rm r}_{eff}(r)}}-1$, with the effective radii ${\\rm r}_{eff}(g)$ and ${\\rm r}_{eff}(r)$, corrected to the rest-frame by linearly interpolating between the radii fitted in the bands $g$, $r$, $i$ and $z$. We show that at least for galaxies with angular $r_{eff}$ radii above 1.8 arcsec, this ratio is well and linearly correlated to ${{d(g-r)}\\over {d(\\rm log~r)}}$ across the observed range of colour gradients and galaxy luminosities, and estimate ${{d(g-r)}\\over {d(\\rm log~r)}}\\simeq -0.87 ({{{\\rm r}_{eff}(g)}\\over{{\\rm r}_{eff}(r)}}-1)$. The large scatter in this correlation, which we initially estimated as 0.122, is reduced to a much more acceptable 0.056 if only galaxies with $r_{eff}$ larger than 1.8 arcsec are included. After excluding the $r_{eff}<1.8$ arcsec galaxies and a few others which we believed to be spiral types, we found for the E/S0s a mean colour gradient, using our measure, of $0.11115\\pm 0.00061$ with a comparable galaxy-to galaxy scatter 0.1172. For the $M_r<-22.5$ non-BCG the mean is similar at $0.10550\\pm 0.00098$ (scatter 0.1141) and for max-BCGs the mean is $0.08142\\pm 0.00114$ (scatter 0.0764). Therefore we find at high significance that BCGs have a flatter (by an average of $23\\%$) radial colour gradient than other high luminosity spheroidals. This again could be due to intensified merging in the cluster-core environment. La Barbera et al. (2005), and similarly Ko and Im (2005), reported mean colour gradients in E/S0s of all luminosities were weaker by almost a factor two in the richest clusters ($N_{gal}>50$) compared to the field, and suggested this was because formation histories in denser cluster environments included more elliptical-elliptical mergers (e.g. Tran et al. 2005), which reduce colour/metallicity gradients (Kobayashi 2004). The simulations of Boylan-Kolchin, Ma and Quataert (2006) account for the steeper $R_e$--$L$ of BCGs, i.e. their `abnormally' large sizes (e.g. Bernardi 2009), as the result of repeated `dry' (dissipationless, non-star-forming) mergers, and especially of the anisotropic accretion of smaller spheroidals in near-radial, low angular momentum orbits, which might occur preferentially in a cluster core. A few BCGs have been observed undergoing such mergers (Liu et al. 2008; Tran et al. 2008). Di Matteo et al. (2009) predict from simulations that the metallicity gradient in a dissipationless (`dry') merger remnant will consistently be $\\simeq 0.6$ the mean of the progenitor galaxies' gradients (but not reduced to zero). \\medskip (iii) In non-BCG E/S0s, we examine the trends in mean colour gradient as a function of other galaxy properties. We find a dependence of colour gradient on absolute magnitude $M_r$, with a broad peak in colour gradient at $M_r\\simeq - 22$ and a decrease to lower and higher luminosities. The luminosity dependence is relatively mild, less than a factor 1.5, which may explain why it was not seen by La Barbera et al. (2005). However, Spolaor et al. (2009) do find a luminosity dependence similar to ours. Colour gradients tend to decrease with increasing velocity dispersion $\\sigma$, by almost 1/2 between 150 and 300 km $\\rm s^{-1}$, confirming the suggestion in our Paper I that the trends with $M_r$ and $\\sigma$ are different. \\medskip (iv) Colour gradients in E/S0s show strong correlations with some other galaxy properties. Colour gradients increase, by about a factor of two, with from small to larger effective radius, up to a maximum at 8-12 kpc (depending on luminosity). A positive correlation with radius was prevously reported by Tamura and Ohta (2003) for a small sample of E/S0s in Abell 2199. At even larger radii we find the mean colour gradients decrease, but this is actually because of a reduced gradient in the most luminous ($M_r<-23$) galaxies. If E/S0s are divided by luminosity, within low/moderate luminosity intervals there is simply a steep increase in colour gradient with $\\rm r_{eff}$. We find colour gradients to be negatively correlated with $\\rm 10~log~\\sigma+M_r$ (the $\\sigma$ residual relative to $\\langle \\sigma|M_r\\rangle$) and mass density ($\\sigma^2/\\rm r_{eff}^2$). Of course, these quantities are related -- a high density implies a high $\\sigma$ at a given dynamic mass (rather than luminosity). These negative correlations do not seem to be caused by a selection effect (from the flux limit of the sample), as they are seen within a wide range of narrow luminosity intervals, being strongest for E/S0s of moderate/high luminosity $-215\\rm \\AA$ in both $\\rm H\\beta$ and $\\rm H\\delta$. Suh et al. (2010) found that blue-cored early-types in general had at least mildly enhanced $\\rm H\\beta$ absorption, and showed that the Balmer enhancement and the positive colour gradients could both be accounted for by centrally concentrated, very young stellar populations (age $\\leq \\rm 0.5 Gyr$), comprising only $\\sim0.5$--$2\\%$ of the stellar mass. \\medskip (vi) Colour gradients in BCGs are consistently lower than in non-BCG spheroidals of the same luminosity, velocity dispersion or radius. Some classes of non-BCG ellipticals, those with (i) the highest luminosities and largest radii, (ii) the highest $\\sigma$ relative to $M_r$, or (iii) the highest densities, also have mean colour gradients of ${{{\\rm r}_{eff}(g)}\\over{{\\rm r}_{eff}(r)}}\\simeq 0.08$. But within the BCG class mean ${{{\\rm r}_{eff}(g)}\\over {{\\rm r}_{eff}(r)}}-1$ remains at $\\sim 0.08$, almost regardless of the other galaxy properties. The mean colour gradient in BCGs decreases a little for the largest radii and with increasing dynamic mass, but is uncorrelated with mass density, the $10~{\\rm log}~\\sigma+M_r$ residual, or stellar age (except for the drop at $<3$ Gyr). The process of colour gradient flattening through dry mergers appears to have operated most strongly on (all) BCGs, reducing the mean colour gradients to $\\sim 0.08$, but never much lower than this, whatever the initial value. In the Di Matteo et al. (2009) model of dry mergers, a galaxy with a strong colour gradient will probably merge with one of lower gradient and experience a significant flattening, but a galaxy with already a very flat colour gradient would probably merge with one with higher gradient and experience no further decrease or even a slight increase. With repeated mergers this would cause an exponential decrease in the mean gradient and also a narrowed distribution, which is what we seem to find for the merged classes of galaxies, with mean gradients of $\\simeq 0.08$ consistently. We found the BCGs had ${1\\over 3}$ less galaxy-to-galaxy variation in colour gradient than the E/S0s (this difference might be even greater after taking measurement errors into account). In addition, the radial mergers thought to occur in BCGs would strongly increase their radii (and reduce density) while simultanously flattening colour gradients, thus erasing the gradient-radius correlation (and gradient-density anticorrelation) seen in other spheroidals. \\medskip (vii) We examine colour as a function of stellar age, comparing the BCGs and other E/S0s. The BCGs are on average 0.5 Gyr older at observation than the $M_r<-22.5$ non-BCGs, and their mean lookback time to formation is estimated as 0.68 Gyr greater. This could simply reflect the correlation of earlier formation (by as much as $\\sim 1$ Gyr) with high-density cluster environment (e.g. Sheth et al. 2006; Rogers et al. 2010). The BCGs also have a flatter age-luminosity relation, as might be expected if BCGs (of all masses) form from dry major mergers of red early-type galaxies (with older ages associated with the cluster-core environment), which would increase luminosity without changing the mean stellar age. In the same way, their CMR is flattened. If, on the other hand, the BCGs had formed `monolithically', without major merging, they might show stronger trends in their stellar populations (age, composition) with luminosity/mass, extrapolating the trends seen in less massive E/S0s. In rest-frame model-magnitude $g-r$, BCGs are slightly redder than non-BCG E/S0s of the same age, whereas in spectra-derived $g-r$, they lie on the same (approximately linear) relation of colour to age. We conclude that the ($\\simeq 0.01$ mag) redder model-magnitude $g-r$ colours of the BCGs in the CMR and C$\\sigma$R can be explained simply by their greater mean ages combined with their flatter colour gradients, which would give them a slightly redder model-magnitude colour for a given central-aperture (e.g. spectra-derived) colour. \\medskip (viii) We re-examine the correlations of colour gradient with galaxy properties, dividing by age (star-formation redshift). The positive correlation of colour gradient with radius, and the anticorrelation with density and $10~{\\rm log}~\\sigma +M_r$ are strongest in younger galaxies but are seen within all age intervals. With increasing age, these correlations flatten somewhat while the mean colour gradients decrease, as might be expected for sequential dry mergers in the Di Matteo et al. (2009) model. The decrease in colour gradient with the $\\sigma$ residual $10~{\\rm log}~\\sigma+M_r$ is not solely due to the known correlation of this quantity with stellar age (Forbes and Ponman 1999; Bernardi et al. 2005; Gallazzi et al. 2006), as it is seen within each age interval. Rather, it seems the colour gradient must have independent anticorrelations with the $\\sigma$ to luminosity ratio and with stellar age. It seems the process of spheroid formation sets up a metallicity/colour gradient, with a steepness positively correlated with the mass/luminosity of stars and also with a large radius and low mass density for the galaxy. Subsequent mergers flatten the stronger gradients while adding low-gradient galaxies at the high-luminosity end of the E/S0 luminosity function. As dry mergers tend to increase size this also weakens the gradient-density relation (Figure 23). A comparison with colour gradients in much higher redshift E/S0s would help to confirm this. The mean colour gradient in E/S0s decreases somewhat at $M_{dyn}>10^{11.5}M_{\\odot}$ because most of these higher-mass galaxies are the products of elliptical-elliptical (dry) mergers (Tran et al. 2005; Naab, Khochfar and Burkert 2006). Yet some E/S0s remain today with colour gradients more than twice the observed average, like those in the `monolithic' models of Kobayashi et al. (2004). The position of BCGs on Figures 21 to 24 is that of galaxies which have experienced more age and/or merger-related colour gradient suppression than even the $z_{form}>2$ E/S0s. On Fig 18 the BCG colour gradient matches that of non-BCGs of mean stellar age $>10$ Gyr, although the BCG mean stellar age is only 7.5 Gyr. The colour gradients in BCGs could be effectively `super-aged' by several Gyr by their environment and their position in the path of infalling cluster spheroidals." }, "0911/0911.5727_arXiv.txt": { "abstract": "\\begin{center}{\\bf Abstract}\\end{center} A new accelerating cosmology driven only by baryons plus cold dark matter (CDM) is proposed in the framework of general relativity. In this model the present accelerating stage of the Universe is powered by the negative pressure describing the gravitationally-induced particle production of cold dark matter particles. This kind of scenario has only one free parameter and the differential equation governing the evolution of the scale factor is exactly the same of the $\\Lambda$CDM model. For a spatially flat Universe, as predicted by inflation ($\\Omega_{dm}+\\Omega_{baryon}=1$), it is found that the effectively observed matter density parameter is $\\Omega_{meff} = 1- \\alpha$, where $\\alpha$ is the constant parameter specifying the CDM particle creation rate. The supernovae test based on the Union data (2008) requires $\\alpha\\sim 0.71$ so that $\\Omega_{meff} \\sim 0.29$ as independently derived from weak gravitational lensing, the large scale structure and other complementary observations. ", "introduction": "It is well known that observations from Supernovae Type Ia (SNeIa) provide strong evidence for an expanding accelerating Universe \\cite{Riess07,Union08}. In relativistic cosmology, such a phenomenon is usually explained by the existence of a new dark component (in addition to cold dark matter), an exotic fluid endowed with negative pressure \\cite{review}. Many candidates for dark energy have been proposed in the literature \\cite{decaying,XM,SF,CGas}, among them: (i) A cosmological constant ($\\Lambda$), (ii) a decaying vacuum energy density or $\\Lambda(t)$-term, (iii) a relic scalar field slowly rolling down its potential, (iv) the ``X-matter'', an extra component characterized by equation of state $p_x=\\omega \\rho_x$, where $\\omega$ may be constant or a redshift dependent function, (iv) a Chaplygin-type gas whose equation of state is $p=-A/\\rho^{\\gamma}$, where A and $\\gamma$ are positive parameters. All these models explain the accelerating stage, and, as such, the space parameter of the basic observational quantities is rather degenerate. Nowadays, the most economical explanation is provided by the flat $\\Lambda$CDM model which has only dynamic free parameter, namely, the vacuum energy density. It seems to be consistent with all the available observations provided that the vacuum energy density is fine tuned to fit the data ($\\Omega_{\\Lambda} \\sim 0.7$). However, even considering that the addition of these fields explain the late time accelerating stage and other complementary observations \\cite{CMB,Clusters}, the need of (yet to be observed) dark energy component with unusual properties is certainly a severe hindrance. In general relativistic cosmology, the presence of a negative pressure is the key ingredient to accelerate the expansion. In particular, this means that cosmological models dominated by pressureless fluid like a CDM component expands in a decelerating way. However, as first discussed by Prigogine and coworkers \\cite{Prigogine} and somewhat clarified by Calv\\~{a}o and collaborators \\cite{LCW} through a manifestly covariant formulation, the matter creation process at the expense of the gravitational field is also macroscopically described by a negative pressure. Later on, it was also demonstrated that the matter creation is an irreversible process completely different from the bulk viscosity description \\cite{LG92} originally proposed by Zeldovich \\cite{Zeld70} to avoid the singularity, as well as to describe phenomenologically the emergence of particles in the begin of the Universe evolution early (see also \\cite{SLC02} for a more complete discussion comparing particle creation and bulk viscosity). Microscopically, the gravitationally induced particle creation mechanism has also been discussed by many authors \\cite{Parker,BirrellD}. A non-stationary gravitational background influences quantum fields in such a way that the frequency becomes time-dependent. In the case of a flat Friedmann-Robertson-Walker (FRW) spacetime described in conformal time coordinates, the key result is that the scalar field obeys the same equation of motion as a massive scalar field in Minkowski spacetime, except that the effective mass becomes time dependent (the dispersion relation in the FRW metric involves the scale factor and its second derivative). When the field is quantized, this leads to particle creation, with the energy for newly created particles being supplied by the classical, time-varying gravitational background. In this context, we are proposing here a new flat cosmological scenario where the cosmic acceleration is powered uniquely by the creation of cold dark matter particles. It will be assumed that the CDM particles are described by a real scalar field so that only particle creation takes place because in this case it is its own antiparticle. The model can be seen as a workable alternative to the cosmic concordance cosmology because it has only one free parameter and the equation of motion is exactly the same of the $\\Lambda$CDM model. As we shall see, in the case of a spatially flat Universe ($\\Omega_{dm}+\\Omega_{bar}=1$), the effectively observed matter density parameter is $\\Omega_{eff} = 1- \\alpha$, where $\\alpha$ is the constant parameter defining the creation rate. The supernovae test requires the central value $\\alpha\\sim 0.71$ so that $\\Omega_{eff} \\sim 0.29$ in accordance with the large scale structure and other complementary observations. ", "conclusions": "A new creation cold dark matter (CCDM) cosmology has been proposed. In this late time CDM dominated model, the vacuum energy density parameter is $\\Omega_{\\Lambda} = 0$, and, therefore, the so-called cosmological constant problem \\cite{weinb, decaying} is absent. The late time acceleration is powered here by an irreversible creation of CDM particles and the value of $H_0$ does not need to be small in order to solve the age problem. \\begin{table}[ht] \\caption{$\\Lambda$CDM vs. CCDM} \\centering \\begin{tabular}{c c} \\hline\\hline $\\Lambda$CDM & CCDM \\\\ [0.5ex] \\hline $\\Omega_\\Lambda$ & $\\alpha$ \\\\ $\\Omega_m$ & $\\Omega_{meff}\\equiv \\Omega_m - \\alpha$ \\\\ Vacuum DE & Creation of CDM \\\\ Acceleration ($z_t \\approx 0.71$, $k=0$)\\, & \\,acceleration ($z_t \\approx 0.71$, $k=0$) \\\\ [1ex] \\hline \\end{tabular} \\label{tab1} \\end{table} It is worth noticing the existence of a dynamic equivalence between CCDM and $\\Lambda$CDM cosmologies at the level of the background equations. Actually, the CCDM scenario can formally be interpreted as a two component fluid mixture: a pressureless matter with density parameter, $\\Omega_{eff} = \\Omega_m -\\alpha$, plus a ``vacuum fluid\" with $\\rho_v=-p_v = \\alpha \\rho_{co}$, where $\\rho_{co}$ is the critical density parameter. A simple qualitative comparison between both approaches is summarized on Table 1. Note that for nonflat CCDM, there are two dynamic free parameters, namely, $\\alpha$ and $\\Omega_m$ (or $\\Omega_{meff}$), similarly to nonflat $\\Lambda$CDM, whose dynamic free parameters are $\\Omega_{\\Lambda}$ and $\\Omega_m$. For the flat CCDM case, there is just one dynamic free parameter (like in flat $\\Lambda$CDM model), say, $\\alpha$. This formal equivalence explains why CCDM scenarios provide an excellent fit to the observed dimming of distant type Ia supernovae (see text and Figs. 1a and 1b). As it appears, one may say that $\\Lambda$CDM cosmology is one of the possible effective descriptions of cold dark matter creation scenarios. On the other hand, since the creation mechanism adopted here is classically described as an irreversible process \\cite{Prigogine,LCW}, the basic problem with this new cosmology is related to the absence of a consistent approach based in quantum field theory in curved spacetimes. However, the last 30 years thinking about the cosmological constant problem are suggesting that the possible difficulties in searching for a more rigorous quantum approach for matter creation in the expanding Universe are much smaller than in the so-called $\\Lambda$-problem. Indeed, the basic tools have already been discussed long ago \\cite{Parker,BirrellD}, and, as such, the problem now is reduced to take into account properly the entropy production rate present in the creation mechanism and the associated creation pressure. Finally, for those believing that $\\Lambda$CDM contains all the physics that will be needed to confront the next generation of cosmological tests we call attention for the model proposed here. It is simple like $\\Lambda$CDM, has the same dynamics, and, more important, it is based only in cold dark matter whose status is relatively higher than any kind of dark energy. Naturally, new constraints on the relevant parameters ($\\alpha$ and $\\Omega_m$) from complementary observations need to be investigated in order to see whether the CCDM model proposed here provides a realistic description of the observed Universe. In principle, additional tests measuring the matter power spectrum, the weak gravitational lensing distortion by foreground galaxies and the cluster mass function may decide between $\\Lambda$CDM and CCDM cosmologies. New bounds on the CCDM parameters coming from background and perturbed cosmological equations will be discussed in a forthcoming communication." }, "0911/0911.0872.txt": { "abstract": "An interesting new high-energy pulsar sub-population is emerging following early discoveries of gamma-ray millisecond pulsars (MSPs) by the \\textit{Fermi} Large Area Telescope (LAT). We present results from 3D emission modeling, including the Special Relativistic effects of aberration and time-of-flight delays and also rotational sweepback of B-field lines, in the geometric context of polar cap (PC), outer gap (OG), and two-pole caustic (TPC) pulsar models. In contrast to the general belief that these very old, rapidly-rotating neutron stars (NSs) should have largely pair-starved magnetospheres due to the absence of significant pair production, we find that most of the light curves are best fit by TPC and OG models, which indicates the presence of narrow accelerating gaps limited by robust pair production -- even in these pulsars with very low spin-down luminosities. The gamma-ray pulse shapes and relative phase lags with respect to the radio pulses point to high-altitude emission being dominant for all geometries. We also find exclusive differentiation of the current gamma-ray MSP population into two MSP sub-classes: light curve shapes and lags across wavebands impose either pair-starved PC (PSPC) or TPC / OG-type geometries. In the first case, the radio pulse has a small lag with respect to the single gamma-ray pulse, while the (first) gamma-ray peak usually trails the radio by a large phase offset in the latter case. Finally, we find that the flux correction factor as a function of magnetic inclination and observer angles is typically of order unity for all models. Our calculation of light curves and flux correction factor for the case of MSPs is therefore complementary to the ``ATLAS paper'' of Watters et al.\\ for younger pulsars. ", "introduction": "\\label{sec:intro} The field of gamma-ray pulsars has already benefited profoundly from discoveries made during the first year of operation of the \\textit{Fermi} / Large Area Telescope (LAT). These include detections of the radio-quiet gamma-ray pulsar inside the supernova remnant CTA~1 \\citep{Abdo08_CTA1}, the second gamma-ray millisecond pulsar (MSP) \\citep{Abdo09_J0030} following the \\textit{EGRET} $4.9\\sigma$-detection of PSR~J0218+4232 \\citep{Kuiper04}, the~6 high-confidence \\textit{EGRET} pulsars \\citep{Thompson99,Thompson04}, and discovery of~16 radio-quiet pulsars using blind searches \\citep{Abdo09_BS}. In addition, 8~MSPs have now been unveiled \\citep[][see Table~\\ref{tab1}]{Abdo09_MSP}, confirming expectations prior to \\textit{Fermi's} launch in June 2008 \\citep{HUM05,Venter05_Cherenkov}. A \\textit{Fermi} six-month pulsar catalog is expected to be released shortly \\citep{Abdo09_Cat}. \\textit{AGILE} has also reported the discovery of~4 new gamma-ray pulsars, and marginal detection of~4 more \\citep{Halpern08,Pellizzoni09b}, in addition to the detection of~4 of the \\textit{EGRET} pulsars \\citep{Pellizzoni09a}. Except for the detection of the Crab at energies above 25~GeV \\citep{Aliu08}, no other pulsed emission from pulsars has as yet been detected by ground-based Cherenkov telescopes \\citep{Schmidt05,Albert07,Aharonian07,Fuessling08,Kildea08,Konopelko08,Albert08,Celik08,DelosReyes09}. MSPs are characterized by relatively short periods $P\\lesssim30$~ms and low surface magnetic fields $B_0\\sim10^8-10^9$~G, and appear in the lower left corner of the $P\\dot{P}$-diagram (with $\\dot{P}$ the time-derivative of $P$; see Figure~\\ref{fig:PPdot}, where the newly-discovered \\textit{Fermi} MSPs are indicated by squares). MSPs are thought to have been spun-up to millisecond periods by transfer of mass and angular momentum from a binary companion during an accretion phase \\citep{Alpar82}. This follows an evolutionary phase of cessation of radio emission from their mature pulsar progenitors, after these have spun down to long periods and crossed the ``death line'' for radio emission. These ``radio-silent'' progenitors \\citep{Glendenning00} are thought to reside in the ``death valley'' of the $P\\dot{P}$-diagram, which lies below the inverse Compton scattering (ICS) pair death line \\citep{HM02}. The standard ``recycling scenario'' \\citep{Bhattacharya91} hypothesizing that MSP birth is connected to low-mass X-ray binaries (LMXRBs) might have been confirmed recently by the detection of radio pulsations from a nearby MSP in an LMXRB system, with an optical companion star \\citep{Archibald09}. Optical observations indicate the presence of an accretion disk within the past decade, but none today, raising the possibility that the radio MSP has ``turned on'' after termination of recent accretion activity, thus providing a link between LMXRBs and the birth of radio MSPs. High-energy (HE) radiation from pulsars has mainly been explained as originating from two emission regions. Polar cap (PC) models \\citep{Harding78,Daugherty82,Sturner95,DH96} assume extraction of primaries from the stellar surface and magnetic pair production of ensuing HE curvature radiation (CR) or ICS gamma rays, leading to low-altitude pair formation fronts (PFFs) which screen the accelerating electric field \\citep{HM98,HM01,HM02}. These space-charge-limited-flow (SCLF) models have since been extended to allow for the variation of the CR PFF altitude across the PC and therefore acceleration of primaries along the last open magnetic field lines in a slot gap (SG) scenario \\citep{AS79,Arons83,MH03_SG,MH04_SG,Harding08_Crab}. The SG results from the absence of pair creation along these field lines, forming a narrow acceleration gap that extends from the neutron star (NS) surface to near the light cylinder. The SG model is thus a possible physical realization of the two-pole caustic (TPC) geometry \\citep{Dyks03}, developed to explain pulsar HE light curves. On the other hand, outer gap (OG) models \\citep{CHR86a,CHR86b,CR92,CR94,Romani96,Cheng00,Zhang04} assume that HE radiation is produced by photon-photon pair production-induced cascades along the last open field lines above the null-charge surfaces ($\\mathbf{\\Omega}\\cdot\\mathbf{B}=0$, with $\\Omega=2\\pi/P$), where the Goldreich-Julian charge density \\citep{GJ69} changes sign. The pairs screen the accelerating E-field, and limit both the parallel and transverse gap size \\citep{Takata04}. Classical OG models may be categorized as ``one-pole caustic models'', as the assumed geometry prevents observation of radiation from gaps (caustics) associated with both magnetic poles \\citep{Harding05}. More recently, however, \\citet{Hirotani06,Hirotani07} found and applied a 2D, and subsequently a 3D \\citep{Hirotani08} OG solution which extends toward the NS surface, where a small acceleration field extracts ions from the stellar surface in an SCLF-regime (see also \\citet{Takata04,Takata06}, and in particular \\citet{Takata08} for application to Vela). Lastly, \\citet{Takata09} modeled Geminga using an OG residing between a ``critical'' B-field line (perpendicular to the rotational axis at the light cylinder) and the last open field line. Current models using dipole field structure to model MSPs predict largely unscreened magnetospheres due to the relatively low B-fields inhibiting copious magnetic pair production. Such pulsars may be described by a variation of the PC model (applicable for younger pulsars), which we will refer to as a ``pair-starved polar cap'' (PSPC) model \\citep{MH04_PS,HUM05,MH09_PS}. In a PSPC model, the pair multiplicity is not high enough to screen the accelerating electric field, and charges are continually accelerated up to high altitudes over the full open-field-line region. The formation of a PSPC ``gap'' is furthermore naturally understood in the context of an SG accelerator progressively increasing in size with pulsar age, which, in the limit of no electric field screening, relaxes to a PSPC structure. Several authors have modeled MSP gamma-ray fluxes, spectra and light curves in both the PSPC \\citep{Frackowiak05a,Frackowiak05b,HUM05,VdeJ05,Venter_phd,Zajczyk08} and OG \\citep{Zhang03,Zhang07} cases. Collective emission from a population of MSPs in globular clusters \\citep{HUM05,Zhang07,BS07,Venter08,Venter09} and in the Galactic Center \\citep{Wang06} have also been considered. \\citet{Watters09} recently calculated beaming patterns and light curves from a population of canonical pulsars with spin-down luminosities $\\dot{E}_{\\rm rot} > 10^{34}$~erg\\,s$^{-1}$ using geometric PC, TPC, and OG models. They obtained predictions of peak multiplicity, peak separation, and flux correction factor $f_\\Omega$ as functions of magnetic inclination and observer angles $\\alpha$ and $\\zeta$, and gap width $w$. The latter factor $f_\\Omega$ is used for converting observed phase-averaged energy flux $G_{\\rm obs}$ to the total radiated (gamma-ray) luminosity $L_\\gamma$, which is important for calculating the efficiency of converting $\\dot{E}_{\\rm rot}$ into $L_\\gamma$. A good example is the inference of the conversion efficiencies of globular-cluster MSPs which may be collectively responsible for the HE radiation observed from 47~Tucanae by \\textit{Fermi}-LAT \\citep{Abdo09_Tuc}. In this paper, we present results from 3D emission modeling, including Special Relativistic (SR) effects of aberration and time-of-flight delays, and rotational sweepback of B-field lines, in the geometric context of OG, TPC, and PSPC pulsar models. We study the newly-discovered gamma-ray MSP population \\citep{Abdo09_MSP}, and obtain fits for gamma-ray and radio light curves. Our calculation of light curves and flux correction factors $f_\\Omega(\\alpha,\\zeta,P)$ for the case of MSPs is therefore complementary to the work of \\citet{Watters09} which focuses on younger pulsars, although our TPC and OG models include non-zero emission width. Section~\\ref{sec:Model} deals with details of the various models we have applied. We discuss light curves from both observational and theoretical perspectives in Section~\\ref{sec:LCs}, and present our results and conclusions in Sections~\\ref{sec:Results} and~\\ref{sec:Con}. ", "conclusions": "\\label{sec:Con} We presented results from 3D emission modeling of gamma-ray and radio radiation in the framework of geometric PC, OG, and TPC pulsar models, and also for the full-radiation PSPC model. We have applied our results to recent measurements of newly-discovered MSPs by \\textit{Fermi}-LAT. In this sense, we present results complementary to those obtained by \\citet{Watters09} for young pulsars. Previously, it was believed that most MSPs should have unscreened magnetospheres \\citep{HUM05}, as they lie below the predicted CR pair death line on the $P\\dot{P}$-diagram. It was expected that such pair-starved MSPs should have single gamma-ray pulses roughly in phase with the radio \\citep{VdeJ05}. From Figure~\\ref{fig:LC_com1} and~\\ref{fig:LC_com3}, we see the surprising fact that there are indeed MSPs that have double-peaked light curves well fit by TPC / OG models, as are many of the young gamma-ray pulsars. This is interpreted as indicating the operation of a magnetic pair formation mechanism, and copious production of pairs to set up the required emitting gap structure. New ways of creating pairs in low-$\\dot{E}_{\\rm rot}$ pulsars will have to be found to explain this phenomenon. PSR~J0030+0451 illustrates this point very well in that it has the lowest $\\dot{E}_{\\rm rot}$ of the MSP sample ($3.5\\times10^{33}$~erg s$^{-1}$), therefore lying significantly below the calculated CR death line \\citep[e.g.,][]{HMZ02}, and yet exhibits the sharpest double peaks of the current population, implying emission originating in very thin TPC / OG gaps. The problem may be alleviated somewhat by increasing the stellar compactness $\\kappa^\\prime$ (larger mass or smaller radius), motivated by recent measurements of large MSP masses \\citep[up to $\\sim1.7M_\\odot$; see][and references therein]{Verbiest08,Freire09}. This will boost the GR E-fields, and enhance pair creation probability. Another way to do this would be to increase the magnetic field. B-fields that are larger than those usually inferred using the dipole spin-down model (and having smaller curvature radii) may be present when there are multipolar B-components near the surface (or an offset-dipole geometry). In fact, offset dipoles have been suggested in modeling the X-ray light curves of MSPs J0437$-$4715 and J0030$+$0451 \\citep{Bogdanov07,Bogdanov09}. However, detailed investigation of such a scenario and its implications for pair cascades is necessary to place this speculation on sure footing. Another possible origin for higher surface fields is the movement of magnetic poles toward the spin axis during the spin-up phase of an MSP \\citep{Lamb08}. It has been argued that during the spin-up to millisecond periods, the inward motion of the neutron star superfluid vortices produces a strain on the crust, causing the magnetic poles to drift toward the spin-axis \\citep{Ruderman91}. If the two poles are in the same hemisphere prior to spin-up, the poles drift toward each other, producing a nearly orthogonal rotator having the same dipole moment but a surface field that can be orders of magnitude higher \\citep{Chen93}. We find that there is exclusive differentiation between the TPC / OG models on the one hand, and the PSPC model on the other hand. Six MSPs have gamma-ray light curves which lag the radio and are explained using TPC or OG fits, but not PSPC fits. For the remaining two MSPs, the radio light curves slightly lag the gamma-ray light curves, and these are fit by the PSPC model (and not by the TPC / OG models). It therefore seems that there are two subclasses emerging within the current gamma-ray MSP sample, and it is not obvious which pulsar characteristics provide a means to predict subclass membership. From our model light curve fitting, we furthermore find $(\\alpha,\\zeta)$ values which are in reasonable agreement with values inferred from MSP polarization measurements. Although we find good PSPC fits for the last two MSPs, we caution that the E-field is only approximately known (e.g., it follows from a local electrodynamical model based on a GR dipolar B-field). Future models which take global current flow patterns into account, along with more sophisticated B-field structure, may produce more realistic solutions for the E-field. Our ability to discriminate between different classes of models derives from the fact that we produced both the gamma-ray and radio curves within the same model. We could then use the shape \\textit{and} relative radio-to-gamma phase lag provided by the data to obtain the best-fit model type for each MSP. The data also enabled us to conclude that the emission, in \\textit{all} models considered, must come from the outer magnetosphere. This has now been observed to be true for the bulk of the gamma-ray pulsar population \\citep{Abdo09_Cat}. In the case of PSR~J0437-4715 and PSR~J0613-0200, we find that the TPC model predicts a significant precursor to the main gamma-ray peak, while the OG model predicts no such low-level emission. With more statistics, this effect may possibly become a discriminator between the TPC and OG models. (We assumed that the TPC emission region starts at $r_{\\rm em}=R$ when creating our plots. However, the relative intensity of the precursor and low-level emission predicted by the TPC model may be reduced by limiting the emission region's extension, i.e.\\ only collecting photons above a certain radius $r_{\\rm em} \\geq R_{\\rm min} > R$.) We calculated the flux correction factor in the context of the different models, and found that $f_\\Omega \\sim1$. These values imply a wide beaming angle, and derives from the fact that we obtain best fits for large impact angles. \\citet{Venter_phd} previously found $\\Lambda_{\\rm avg} \\sim 10-30$ (i.e.\\ $f_\\Omega\\sim0.8-2.4$), and $\\Lambda_{\\max} \\sim300$ ($f^{\\max}_\\Omega\\sim24$) for the PSPF model. Now, we find $f_\\Omega$ $\\sim 0.5-2$, and $f^{\\max}_\\Omega\\sim 4$. These results are roughly consistent, with the differences stemming from the following: (i) \\citet{Venter_phd} used energy flux ratios to calculate $\\Lambda$, while we are using photon flux ratios to calculate $f_\\Omega$, assuming that the photon and energy fluxes have similar distributions across $(\\zeta,\\phi)$-space; (ii) \\citet{Venter_phd} only used $E_{||}^{(1)}$ and $E_{||}^{(2)}$ for the E-field, while we now also include the high-altitude solution ($E_{||}^{(3)}$) for the PSPF case. This leads to more intense high-altitude emission, and therefore smaller values of $f_\\Omega$ for off-beam emission. We lastly remark that the larger radio beam widths of MSPs compared to those of canonical pulsars should lead one to expect relatively few radio-quiet MSPs. The spectacular data from \\textit{Fermi}-LAT hold the promise of phase-resolved spectroscopy, at least for the brightest pulsars, and will challenge existing pulsar models to reproduce such unprecedented detail. Future work therefore includes using full acceleration and radiation models to study gamma-ray spectra, luminosities, and light curves, in order to constrain fundamental electrodynamical quantities, and possibly providing the opportunity of probing the emission geometry and B-field structure more deeply. Improved understanding of pulsar models will also feed back into more accurate population synthesis models \\citep[e.g.,][]{Story07}. In addition, we hope to obtain better understanding of important quantities such as MSP efficiencies, and whether this quantity is similar for Galactic-Field and globular-cluster MSPs \\citep{Abdo09_Tuc}." }, "0911/0911.0102_arXiv.txt": { "abstract": "{We investigate claims according to which the X-ray selection of AGN is not as efficient compared to that based on [OIII] selection because of the effects of X-ray absorption. We construct the predicted X-ray luminosity function both for all Seyferts as well as separately for Seyfert-1 and Seyfert-2 type galaxies, by combining the optical AGN [OIII] luminosity functions derived in SDSS with the corresponding $\\rm L_X-L_{[OIII]}$ relations. These relations are derived from {\\it XMM-Newton} observations of all Seyfert galaxies in the Palomar spectroscopic sample of nearby galaxies after correction for X-ray absorption and optical reddening. We compare the predicted X-ray luminosity functions with those actually observed in the local Universe by {\\it HEAO-1}, {\\it RXTE} as well as {\\it INTEGRAL}. The last luminosity function is derived in the 17-60 keV region and thus is not affected by absorption even in the case of Compton-thick sources. In the common luminosity regions, the optically and X-ray selected Seyfert galaxies show reasonable agreement. We thus find no evidence that the [OIII] selection provides a more robust tracer of powerful AGN compared to the X-ray. Still, the optical selection probes less luminous Seyferts compared to the current X-ray surveys. These low luminosity levels, are populated by a large number of X-ray unobscured Seyfert-2 galaxies. \\keywords {X-rays: general; X-rays: diffuse background; X-rays: galaxies}} ", "introduction": "Since the detection of the nearby AGN 3C273 in X-rays by {\\it UHURU} (Kellogg et al. 1971), X-ray observations have been considered to be the primary tool for selecting AGN. Recently, the deepest ever observations in the Chandra Deep Field North and South (Alexander et al. 2003, Giaconni et al. 2002, Luo et al. 2008) have resolved 80-90\\% of the extragalactic X-ray light, the X-ray background, in the 2-10 keV band. These observations reveal a sky density of about 5000 sources per square degree (Bauer et al. 2004), the vast majority of which are AGN (for a review see Brandt \\& Hasinger 2005). These surveys allowed the derivation of the luminosity function and probed with good accuracy the accretion history of the Universe (Ueda et al. 2003, La Franca et al. 2005, Barger et al. 2005). In the local Universe, the X-ray luminosity function has been derived from the wide-angle surveys of {\\it RXTE} and {\\it HEAO-1}. Sazonov \\& Revnitsev (2004) derived the luminosity function for bright AGN, detected at fluxes $>2.5\\times 10^{-11}$ \\funits in the 3-20 keV band, in the {\\it RXTE} slew survey. Shinozaki et al. (2006) derived the luminosity function of AGN using the all-sky HEAO-1 in the 2-10 keV band. Even these hard X-ray surveys may be missing a number of extremely obscured AGN. In the case of Compton-thick AGN, at column densities $>10^{24}$ \\cunits (equivalent to $A_V\\sim 450$ using the Galactic dust-to-gas ratio), a large fraction of the intrinsic flux will be absorbed. The {\\it SWIFT} (Gehrels et al. 2004) and the {\\it INTEGRAL} missions \\ (Winkler et al. 2003) which carry ultra-hard X-ray detectors ($>$15 keV), albeit with limited imaging capabilities, probed energies which are immune to X-ray obscuration up to column densities of $10^{25}$ \\cunits. These missions helped towards further constraining the number density of such heavily absorbed sources at very bright fluxes, $\\rm f_{17-60~keV} >10^{-11}$ \\funits in the local Universe, $z<0.1$ (Beckmann et al. 2006, Bassani et al. 2006, Winter et al. 2008, Winter et al. 2009). The luminosity function in these energies has been derived by Sazonov et al. (2007), Paltani et al. (2008) and Tueller et al. (2009). It appears that the fraction of Compton-thick sources is small and thus the {\\it RXTE} and {\\it HEAO-1} luminosity functions are little affected. In the optical, QSOs have been traditionally selected using colours (Schmidt \\& Green 1983, Marshall et al. 1987, Boyle et al. 2000). The advent of the 2dF (Croom et al. 2004) and the SDSS (York et al. 2000) surveys have generated vast samples of QSOs providing a leap forward in the study of their luminosity function. The derived sky density of QSOs (few hundred per square degree) is at least an order of magnitude lower than that derived from X-ray surveys. This is because the colour selection of AGN requires that the nuclear optical luminosity is much higher than that of the host galaxy for the AGN to be detected (typically $M_B<-23$). Therefore the optical selection based on colours is biased against low luminosity AGN in contrast to the X-ray selection. The limitations of colour optical selection techniques can be circumvented by selecting AGN via their emission lines. Such methods can extend the optical luminosity function to low luminosities. Ho et al. (1997) carried out a spectroscopic survey of about 500 nearby galaxies selected from the revised Shapley-Ames catalogue (Sandage \\& Tammann 1981). They identified Seyfert emission-line characteristics in 52 galaxies. Ulvestad \\& Ho (2001) derived the optical B-band luminosity function from their sample extending the AGN luminosity function to $\\rm M_B\\sim -17$ (see also Georgantopoulos et al. 1999). A limitation in the above works is that the B-band is strongly affected by the host galaxy light. Hao et al. (2005a) extended significantly these results using the SDSS survey to select a sample of about 3000 AGN in the redshift range $0$ & Dispersion \\\\ \\hline A. Seyfert-1 (Ho \\& Heckman)& 28 & $0.84\\pm0.09$ & $8.27\\pm 3.87$ & 1.66 & 0.60 \\\\ B. Seyfert-1 (Ho) & 8 & $0.66\\pm0.22$ & $15.31\\pm8.84$ & 1.84 & 0.82 \\\\ C. Seyfert-1 (Heckman) & 20 & $1.04\\pm0.17$ & $-0.17\\pm7.34$ & 1.59 & 0.48 \\\\ D. Seyfert-2 (Ho luminous+weak) & 23 & $0.83\\pm0.15$ & $7.3\\pm6.1$ & 0.87 & 0.80 \\\\ E. Seyfert-2 (Ho luminous) & 11 & $0.64\\pm0.17$ & $15.34\\pm7.15$ & 1.16 & 0.77 \\\\ F. Seyfert-2 (Ho weak) & 12 & $0.59\\pm0.29$ & $16.18\\pm 11.48$ & 0.61 & 0.78 \\\\ G. Seyfert 1+ 2 (luminous) & 17 & $0.65\\pm0.14$ & $15.35\\pm5.90$ & 1.44 & 0.84 \\\\ \\hline \\end{tabular} \\end{center} \\end{table*} ", "conclusions": "We derived the predicted X-ray luminosity function for Seyfert galaxies in the local Universe, by combining the optical SDSS [OIII] Seyfert luminosity functions with the corresponding {\\lxlo} relation. These relations have been derived using {\\it XMM-Newton} observations (Akylas \\& Georgantopoulos 2009) of the local, optically selected AGN sample of Ho et al. (1997). This sample covers a comparable luminosity range with the SDSS Seyfert sample. We have corrected the X-ray luminosity for the effects of absorption. Our analysis above shows that the predicted X-ray luminosity function is in reasonable agreement with the observed X-ray Seyfert luminosity functions derived in the 2-10 keV and 2-20 keV bands by {\\it HEAO-1} and {\\it RXTE} respectively. Most importantly, it is in agreement with the ultra-hard 17-60 keV {\\it INTEGRAL} luminosity function (Sazonov et al. 2007). As the {\\it INTEGRAL} luminosity function is practically immune to X-ray absorption, this suggests that absorption played little role in the optical/X-ray luminosity function discrepancy reported by Heckman et al. (2005). This is also independently supported by the {\\it XMM-Newton} observations of the Palomar optically selected Seyfert sample (Akylas \\& Georgantopoulos 2009). The fraction of Compton-thick AGN in this optically selected sample is small. In addition we examined separately, the Seyfert-1 and Seyfert-2 luminosity function. The Seyfert-1 luminosity function is in rough agreement or even somewhat above the observed X-ray luminosity function. The predicted Seyfert-2 luminosity function agrees quite well with the optical luminosity function again disfavouring the absorption hypothesis. Indeed, if absorption were the problem, then our predicted luminosity function, which takes X-ray absorption into account, would be well above that of {\\it RXTE}. Having addressed this matter, we need to understand why Heckman et al. (2005) found that the {\\it RXTE} X-ray luminosity function lies below the optical one by about a factor of 3. The comparison was based on the fact that the two luminosity functions have the same slope. However, this is true only for the bright part of the X-ray luminosity function (i.e. for L$_{\\star}(3-20~ \\rm keV)>3\\times 10^{43}$ \\lunits or equivalently L$_{\\star}(2-10 \\rm keV)>2.3 \\times 10^{43}$ \\lunits). Using $\\rm (L_X/L_{[OIII]}) \\approx 100 $ the corresponding optical luminosities must be greater than $5.7 \\times 10^7 L_{\\odot}$. Since the upper end of the optical luminosity function of Hao et al. (2005b) is $\\sim 3 \\times 10^8 L_{\\odot}$ the overlap is very limited (see Fig.5 in Heckman et al.) and thus the comparison is not quite robust. Moreover, the effect of the dispersion plays a critical role as shown here. Although, the densities of the [OIII] and the X-ray selected AGN are comparable, this does not mean that these methods favour the selection of the same objects. The optical selection favours X-ray weak Seyfert-2 (see Fig. 1). It has been proposed (see Ho 2008 for a review) that a large number of sources at low luminosities appear as Seyfert-2 possibly because no Broad-Line-Region is formed at low accretion rates (Nicastro 2000) or low luminosities (Elitzur \\& Shlosman 2006). If the above models hold true, these 'naked' Seyfert-2 galaxies should present no hidden Broad Lines in spectropolarimetric observations (e.g. Tran 2003). Interestingly, Spinoglio et al. (2009) find that the ratio of the X-ray to 12$\\mu m$ luminosity, $L_X/L_{12\\mu m}$, is much lower in the Seyfert-2 with no hidden Broad-Line-Region, suggesting that these are weak X-ray emitters. There is however, one caveat in these 'naked' X-ray weak Seyfert-2 interpretation. The [OIII] (or the mid-IR) luminosity is believed to be a good proxy of the ionizing nuclear luminosity and thus to the X-ray luminosity. But in these Seyfert-2 sources the $L_X/L_{[OIII]}$ ratio is low, implying that only the X-ray luminosity is weak. One explanation could be that star-formation is contributing a large part of the [OIII] emission. The future missions {\\it NUSTAR} and {\\it ASTRO-H} having superb imaging capabilities ($\\sim$ 1arcmin) at ultra-hard energies (10-70 keV), will provide the opportunity to obtain the least unbiased AGN samples, reaching flux levels at least two orders of magnitude fainter than {\\it SWIFT} and {\\it INTEGRAL}." }, "0911/0911.3101_arXiv.txt": { "abstract": "s{ \\ps\\ was successfully launched on May 14th, 2009, from the Kourou space port, in French Guyana. After recalling the objectives that we set out - back in 1996 - to fulfill with this project, I recall some of the technological breakthroughs which needed to be made and report on the exciting scientific outlook of the project in light of the knowledge we now have of the actual performances of the two on-board instruments. I also include one of our more recent results even though it was not yet available at the time of the conference. } The measurement goals of \\plancks may be stated rather simply: to build an experiment able to perform the ``ultimate'' measurement of the primary CMB temperature anisotropies, which requires:\\begin{itemize} \\item full sky coverage and a good enough angular resolution in order to completely mine all scales at which the Cosmic Microwave background (CMB) primary anisotropies contain information ($\\simgt 5$ minutes of arc) \\item a final sensitivity essentially limited by the ability to remove the astrophysical foregrounds, implying a large frequency coverage from 30 GHz to 1 THz (provided by the two instruments: HFI and LFI), with sensitivity at each of the 9 survey frequencies in line with the role of each map in determining the CMB properties . \\end {itemize} For the measurement of the polarisation of the CMB anisotropies, \\plancks goal was ``only'' to get the best polarisation performances with the technology available at the design time This is on these simple but ambitious goals and the proposed way of reaching them that, after 3 years of preparatory work, the project was selected by the European Space Agency (ESA), as the $3^{rd}$ Medium size mission of its Horizon 2000+ program. This selection occurred in march 1996, \\ie contemporaneously with that of WMAP by NASA, which rather proposed reaching earlier less ambitious goals with already existing technology. \\newcommand{\\hf}{\\hfill} \\begin{table}[htb] \\caption{Summary of \\plancks performance goals for the required 14 months of routine operations, which allows nearly all detectors to map the entire sky twice. Requirements on sensitivity are simply two times worse than the stated goals. Central band frequencies, $\\nu$, are in Gigahertz, the FWHM angular sizes are in arc minute, and the (sky-averaged) sensitivities $c^X_{noise}$, with $X = T, Q$ or $U$, are expressed in $\\mu \\mathrm{K.deg}$; this number indicates the rms detector noise, expressed as a equivalent temperature fluctuation in $\\mu \\mathrm{K}$, whcih is expected once it is averaged in a pixel of 1 degree of linear size).} \\vspace{0.4cm} \\begin{center} \\scalebox{1.}{ \\begin{tabular}{|c||c|c|c||c|c|c|c|c|c|} \\hline \\hline & \\multicolumn{3}{|c||}{\\lfi\\ goals}&\\multicolumn{6}{|c|}{\\hfis goals} \\\\ \\hline\\hline $\\nu$ \\hf [GHz] & \\hf 30 & \\hf 44 & \\hf 70 & \\hf 100 & \\hf 143 & \\hf 217 & \\hf 353 & \\hf 545 & \\hf 857 \\\\ FWHM \\hf [arcmin] & \\hf 33 & \\hf 24 & \\hf 14 & \\hf 9.5 & \\hf 7.1 & \\hf 5.0 & \\hf 5.0 & \\hf 5.0 & \\hf 5.0 \\\\ $c^T_{noise}$ \\hf [$\\mu \\mathrm{K.deg}$]& \\hf 3.0 & \\hf 3.0 & \\hf 3.0 & \\hf 1.1 & \\hf 1.4 & \\hf 2.2 & \\hf 6.8 & \\hf - & \\hf - \\\\ $c^{Q or U}_{noise}$ \\hf [$\\mu \\mathrm{K.deg}$]& \\hf 4.5 & \\hf 4.6 & \\hf 4.6 & \\hf 1.8 & \\hf 1.4 & \\hf 2.4 & \\hf 7.3 & \\hf & \\hf \\\\ \\hline \\end{tabular} } \\end{center} \\label{tab:perfs} \\end{table} Table~\\ref{tab:perfs} summarises the main performance goals of \\ps, expressed for instance as the average detector noise within a square patch of 1 degree of linear size, $c_{noise}$, for the 14 months baseline duration of the mission, which would allow covering twice all the sky by nearly all the detectors. It is interesting to note that if we take the noise performance figure for the average of the central CMB frequencies (the 100-143-217 GHz \\hfis\\ channels, assuming all the other channels are devoted to foregrounds removal), one finds $0.5\\ \\mathrm{\\mu K.deg}$ in temperature and $1\\ \\mathrm{\\mu K.deg}$ for the Q \\& U Stokes parameters. The magnitude of this step forward, if achieved, may be judged by comparing with the WMAP sensitivity which is given in Table~\\ref{tab:WMAPperfs}. The aggregate sensitivity of the WMAP 60 \\& 90 GHz channels is $\\sim 10.8\\ \\mathrm{\\mu K.deg}$ in a year, which would imply about $(10.8/.5)^2\\sim$ 460 years of operations to reach the baseline \\ps\\ sensitivity. In other words, the error bars from noise in the angular power spectrum should be at least hundred times smaller for \\plancks than for \\map (with an even larger difference at smaller scales which will be much better known from \\plancks\\ thanks to the twice higher angular resolution of HFI). \\begin{table}[htb] \\caption{Summary of WMAP in-flight performance per full year of operations. Same units than in Table~\\ref{tab:perfs}. } \\vspace{0.4cm} \\begin{center} { % \\begin{tabular}{|c||c|c|c|c|c|} \\hline \\hline \\multicolumn{6}{|c|}{\\maps (in flight) } \\\\ \\hline\\hline $\\nu$ \\hf [GHz] & \\hf 23 & \\hf 33 & \\hf 41 & \\hf 61 & \\hf 94 \\\\ FWHM \\hf [arc min] & \\hf 49.2 & \\hf 37.2 & \\hf 29.4 & \\hf 19.8 & \\hf 12.6 \\\\ $c^T_{noise}$ \\hf [$\\mu \\mathrm{K.deg}$] & \\hf 12.6 & \\hf 12.9 & \\hf 13.3 & \\hf 15.6 & \\hf 15.0 \\\\ \\hline \\end{tabular} } % \\end{center} \\label{tab:WMAPperfs} \\end{table} \\begin{figure}[htbp] \\begin{center} \\psfig{file=Planck-build-up.jpg, width=\\textwidth} \\end{center} \\caption[]{Planck build-up from the inside out. Going from left to right, one sees on the top row (t1) the HFI instrument with its 52 detector horns poking out of it's outer shell at 4\\,K. (t2) HFI is surrounded by the 11 larger horns from LFI. HFI and LFI together form (t3) the focal plane assembly from which (t4) the electrical signal departs (though a bunch of wave guides for the LFI and a harness of wires for HFI) to connect to the warm electronics parts of the detection chain which are located within the service module at $\\sim$ 300\\,K. (t5) The (top) cold and warm (bottom) parts are separated by three thermally isolating V-grooves which allow radiating to space heat from the spacecraft sideways and quite efficiently. The third (top) V-grooves operating temperature is about 40\\,K. On the middle row, one sees (m1) the beds of the sorption cooler and it's piping around the V-groove, bringing the overall focal-plane structure to LFI's operational temperature of $\\sim 18$\\, K. (m2) The back-to-back (to damp the first harmonics of the vibrations) compressors of the 4\\,K cooler allow bringing HFI outer shell to 4\\,K, while (m3) isotopes from the He3 tank and the three He4 tanks are brought to the mixing pipes within HFI to cool filters (within the horns) to 1.6\\,K and the bolometer plate to 0.1\\,K, before (m4) being released to space. (m5) The passive cooling and the three active stage constitute this complex but powerful cooling chain in space. On the last bottom row, one can also see (b1) some of the electronic boxes in the service module (SVM) which in addition to the warm part of the electronic and cooling chains also contain all ``services'' needed for transmitting data, reconstructing the spacecraft attitude, powering the whole satellite... The bottom of the SVM is covered with solar panels, while supporting struts begin on its top which allows (b3) positioning the secondary and primary reflectors. The top part is surrounded by a large baffle to shield at best the focal plane from stray-light. The back view (b5) allows distinguishing in the back the supporting structure of the primary mirror, and the wave guides from LFI. The spin axis of Planck (vertical on these plots) is meant to be close to the sun-earth line, with the solar panel near perpendicular to that line and the rotation of the line-of-sight (at 1 rpm) causing the detectors to survey circles on the sky with an opening angle around 85 degrees. Copyright ESA.} \\label{fig:build} \\end{figure} We proposed to achieve the ambitious sensitivity goals of \\plancks with a small number of detectors, limited principally by the photon noise of the background (for the CMB ones), in each frequency band. This implied to achieve several technological feats never achieved in space before (see in particular Lamarre et al. 2003, in New Astronomy Reviews, 47, pp. 1017): \\begin{itemize} \\item sensitive \\& fast bolometers with a Noise Equivalent Power $< 2 \\times 10^{-17}\\ \\mathrm{W/Hz}^{1/2}$ and time constants typically smaller than about 5 milliseconds (which thus requires cooling them down to $\\sim 100$ mK, and build them with a very low heat capacity \\& charged particles sensitivity) \\item total power read out electronics with very low noise, $< 6\\ \\mathrm{nV/Hz}^{1/2}$ from 10 mHz (1 rpm) to 100 Hz (\\ie from the largest to the smallest angular scales to measure at the \\ps\\ scanning speed) \\item excellent temperature stability, from 10 mHz to 100 Hz (cf. Lamarre et al. 04), such that the induced variation be a small fraction of the detector temporal noise: \\begin{itemize} \\item better than $ 10\\ \\mathrm{\\mu K/Hz}^{1/2}$ for the 4K box (assuming 30\\% emissivity) \\item better than $30\\ \\mathrm{\\mu K/Hz}^{1/2}$ on the 1.6K filter plate (assuming a 20\\% emissivity) \\item better than $20\\ \\mathrm{nK/Hz}^{1/2}$ for the detector plate (a damping factor $\\sim 5000$ needed) \\end{itemize} \\item very low noise HEMT amplifiers (therefore cooled to 20\\,K) \\& very stable cold reference loads (at 4\\,K) \\end{itemize} In addition, \\plancks requires:\\begin{itemize} \\item a low emissivity telescope with very low side lobes (\\ie strongly under-illuminated) \\item no windows, and minimum warm surfaces between the detectors and the telescope \\item a quite complex cryogenic cooling chain (cf. figure~\\ref{fig:build}) which begins by reaching $\\sim 40K$ via passive cooling, by radiating about 2 Watts to space, followed by three active stages, at 20\\,K, 4\\,K, and 0.1\\,K: \\begin{itemize} \\item 20\\,K for the LFI, with a large cooling power, $\\sim 0.7$ Watts (provided by $H_2$ Joule-Thomson sorption pumps developed by JPL, USA) \\item 4\\,K, 1.6\\,K and 100\\,mK for the HFI (the 15 milli-Watts cooling power at 4K is provided by mechanical pumps provided by the RAL, UK, in order to perform a Joule-Thomson expansion of He; the 1.6K stage has a pre-cooling power of about 0.5 milli-Watts, thanks to another Joule-Thomson expansion, while the final dilution fridge of He$^3$ \\& He$^4$, from a French collaboration between Air Liquide, the CRTBT, can deal with 0.2 micro-Watts at 0.1\\,K). \\item a thermal architecture optimised to damp thermal fluctuations (active+passive) \\end{itemize} \\end{itemize} Furthermore, a tight control of vibrations is needed, in particular since the dilution cooler does not tolerate micro-vibrations at sub-mg level. And as little as $7\\times 10^{10}$ He atoms accumulated on the dilution heat exchanger (an He pressure typically at the $1\\times 10^{-10}$ mb level) would be too much. These top-level design goals have now been turned into real instruments, which went through several qualification models. Before delivering the actual flight model of both instruments to industry for integration with the satellite, both instrumental consortia organised extensive calibrations campaigns starting at the individual components levels, then at the sub-systems levels (e.g. individual photometric pixels), then at instrument level. For HFI, the detector-level tests were done mainly at JPL in the USA, and the pixel level tests were performed in Cardiff in the UK, while the flight instrument calibration was performed at the Institut d'Astrophysique Spatiale in Orsay, France from April till the end of July 2006. During that period, we obtained in particular 19 days of scientific data at normal operating conditions. We could then confirm that HFI satisfies all our {\\em requirements}, and for the most part actually reaches or exceeds the more ambitious design {\\em goals}, in particular concerning the sensitivity, and speed of the bolometers, the very low noise of the read out electronics and the overall thermal stability. In addition, the total optical efficiency has been verified to be satisfactory, optical cross-talk appears negligible, as well as the Current cross-talk and the cross-talk in intensity is weak. Main beams are well-defined and are quite well described by the models, polarisation measurements confirm expectations, etc. The LFI instrument also went through detailed testing around the same time and it does reach most of its ambitious requirements. \\begin{figure}[htbp] \\begin{center} \\psfig{file=csl.jpg, width=\\textwidth} \\end{center} \\caption[]{Planck on May 2008, hanging outside the vacuum cryogenic chamber at CSL before completing it's first and last full thermal vacuum test. Copyright ESA.} \\label{fig:CSL} \\end{figure} The integration of the LFI and HFI instruments was performed at Thales premises in Cannes in November 2006 and within a year, by December 2007, the full satellite was ready for vibration testing. Planck was then flown from Cannes to ESA's ESTEC centre (in Noordwijk, Holland) where among other things it went through load balancing on April 7th, before travelling again to the ``Centre Spatial de Li\\'eges'' (CSL) in april 2008. Figure~\\ref{fig:CSL} is a picture of Planck hanging outside the vacuum cryogenic chamber at CSL, before the start of the first (and last) full thermal test with all elements of the cryogenic chain present and operating. This ultimate system-level (ground) test demonstrated in particular the following: \\begin{itemize} \\item the dilution system can work with the minimal Helium 3 and 4 flux, which should allow 30 months of survey duration (nominal duration being 14 months!) \\item the extremely demanding temperature stability required (at 1/5 of the detection noise) has been verified \\item bolometers sensitivities in flight conditions are indeed centered around the goal, as shown in figure~\\ref{fig:bolos}. \\end{itemize} \\begin{figure}\\begin{center} \\psfig{figure=Bolos.jpg,width=0.6\\textwidth} \\end{center} \\caption[]{Measured values of the Noise Equivalent Power of HFI detectors during the ground test in flight conditions at CSL on May 2008. One sees that the median value is at the goal level.} \\label{fig:bolos} \\end{figure} \\ps\\ was then shipped to Kourou, and after a few more nerve \u2013wracking delays, we finally lost sight of Planck for ever (when it was covered by the SYLDA support system on the top of which laid Herschel for a joint launch). Launch was on May 14th, and it was essentially perfect. After separating from Herschel, \\ps\\ was set in rotation and started its to the L2 Lagrange point of the sun-earth system, at 1.5 million kilometres away from earth, \\ie about 1\\% further away from the sun than the earth. The final injection in the L2 orbit was at the end of June, shortly after the end of the Blois meeting (see figure~\\ref{fig:trajcool}-a), at the same time than the cooling sequence ended successfully. Indeed, figure~\\ref{fig:trajcool}-b shows how the various thermal stages reached their operating temperature, cooling of coarse from the outside-in, and closely following the predicted pattern. \\begin{figure}\\begin{center} \\psfig{figure=trajectory.jpg,width=0.45\\textwidth} \\psfig{figure=cooling.jpg,width=0.54\\textwidth} \\end{center} \\caption[]{a) Spacecraft trajectory to and on the L2 orbit. b) Cooling sequence of Planck, showing the various stages reaching in turn their operational temperature, till the dilution plate actually reached 93 mK on July 3rd. Credit ESA and HFI consortium} \\label{fig:trajcool} \\end{figure} Once at L2, a calibration and performance verification phase was conducted till mid-august, to insure that all system are working properly and that instrumental parameters are all set at best. From August 13th to 27th, we conducted a ``First Light Survey'' (FLS) in normal operational mode for an ultimate verification of parameters and of the long-term stability of the experiment. We found the data quality to be excellent, and the Data processing Centre pipelines could be operated as hoped to produce the first images. Figure~\\ref{fig:strip} (extracted from the Press release we made on September 17th) illustrates the FLS coverage by showing an image generated from the data acquired from a single 100GHz detector of HFI superimposed to an image of the optical sky by Axel Mellinger. We also released (see the press release in English at ESA's site http://sci.esa.int/science-e/www/object/index.cfm?fobjectid=45543 and in French at http://public.planck.fr/actualiteFLS.php) a comparison of a high latitude field whose emission ought to be dominated by the CMB (shown by a small white square in the figure) as observed by an HFI and LFI detector. They demonstrate an excellent similarity while the two instruments are using quite different technologies. Nine images at all the frequencies covered by Planck of a Galaxy crossing area (indicated by the large white square of the figure) provide visual evidence to the richness of the dataset that Planck shall deliver, allowing very broad scientific studies outreaching its primary cosmological goals. Indeed an important part of Planck long term legacy will be the unique set of maps of the millimetric and sub-millimetric polarised full sky. \\begin{figure}\\begin{center} \\psfig{figure=PR1.jpg,width=\\textwidth} \\end{center} \\caption[]{Sky coverage during Planck First Light Survey (FLS). Planck measures a ring of about 1 degree per day (for a full rotation around the sun in a year), yielding this $\\sim 15$ degree strip during the two weeks of the FLS. The detector image shown in Galactic coordinates is superimposed to a background image where the Galactic plane is clearly visible. Credit ESA and HFI \\& LFI consortia. } \\label{fig:strip} \\end{figure} With the success of the FLS, the normal survey operations have now started. We should therefore be in a position to deliver in December 2010, as planned, an \u201cEarly Release Point Source Catalogue\u201d based on a rapid analysis of the first coverage of the sky which will be issued in time to allow the astrophysical community at large to propose follow-ups by Herschel during its expected cryogenic life. A first public release based on the data from the nominal 14 months mission is slated for december 2012. The release should contain the clean calibrated time-ordered data of each detector, the nine full sky maps at the six frequencies covered by HFI and the three ones by LFI, possibly supplemented by polarisation maps, as well as maps of identified astrophysical components (CMB, Galactic Emissions, Extragalactic sources catalogue), some ancillary information (\\eg on beams, spectral transmission, etc), accompanied by about 50 scientific papers describing the mission, how the ``products'' were obtained, validated, and the results of a first pass of scientific exploitation by the Planck collaboration itself, encompassing in particular the implications of the measured statistical properties of the CMB. Our anticipation from the measured Helium consumption in flight is that the mission duration will exceed the nominal duration of 14 months, and we plan a further release, about a year later on the basis of the extra data which might allow covering as much as five times in total the entire sky. In addition to an improved sensitivity, this extra duration will foremost allow greater data redundancy and therefore a tighter control of all systematic effect which can be searched for with a longer baseline. This should allow us detecting the gravitational wave stochastic background predicted in one interesting class of inflationary models, providing the long thought after ``smoking gun'' of inflation, or otherwise put meaningful constraints on the viable inflation models remaining. {\\em In conclusion, \\ps\\ is now in normal operation \\& performances are as expected or better.}\\\\ This gives us confidence that the scientific program of \\ps\\ can be carried through as anticipated. The dataset should in particular allow addressing many key cosmological questions, including the existence of a primordial gravitational wave background, or that of highly revealing deviations from the current minimal model, where the primordial fluctuation can be purely Gaussian, adiabatic, scale-free, in a strictly flat spatial geometry with a dark energy component describable as a pure cosmological constant, and (cosmologically) negligible neutrinos masses. A rather complete overview of the scientific Program of \\ps\\ can be found in the so-called ``Blue Book'' which was issued in 2004. It can be downloaded from \\\\ {\\em http://www.planck.fr/IMG/pdf/Planck\\_book.pdf}. \\\\ In addition, we submitted a series of pre-launch papers (all with a title starting with ``Planck pre-launch status:'') giving many details of the design and tests of the mission, the instruments, and some of their components. ", "introduction": " ", "conclusions": "" }, "0911/0911.3271_arXiv.txt": { "abstract": "% The habitability of planets is strongly affected by impacts from comets and asteroids. Indications from the ages of Moon rocks suggest that the inner Solar System experienced an increased rate of impacts roughly 3.8 Gya known as the Late Heavy Bombardment (LHB). Here we develop a model of how the Solar System would have appeared to a distant observer during its history based on the Nice model of Gomes et al. (2005). We compare our results with observed debris discs. We show that the Solar System would have been amongst the brightest of these systems before the LHB. Comparison with the statistics of debris disc evolution shows that such heavy bombardment events must be rare occurring around less than 12\\% of Sun-like stars. ", "introduction": " ", "conclusions": "" }, "0911/0911.1274_arXiv.txt": { "abstract": "The prediction of the spin of the black hole resulting from the merger of a generic black-hole binary system is of great importance to study the cosmological evolution of supermassive black holes. Several attempts have been recently made to model the spin via simple expressions exploiting the results of numerical-relativity simulations. Here, I first review the derivation of a formula, proposed in Ref.~\\cite{new}, which accurately predicts the final spin magnitude and direction when applied to binaries with separations of hundred or thousands of gravitational radii. This makes my formula particularly suitable for cosmological merger-trees and N-body simulations, which provide the spins and angular momentum of the two black holes when their separation is of thousands of gravitational radii. More importantly, I investigate the physical reason behind the good agreement between my formula and numerical relativity simulations, and nail it down to the fact that my formula takes into account the post-Newtonian precession of the spins and angular momentum in a consistent manner. ", "introduction": " ", "conclusions": "" }, "0911/0911.1056_arXiv.txt": { "abstract": "The recent detection of blazar 3C279 by MAGIC has confirmed previous indications by H.E.S.S. that the Universe is more transparent to very-high-energy gamma rays than previously thought. We show that this fact can be reconciled with standard blazar emission models provided photon oscillations into a very light Axion-Like Particle occur in extragalactic magnetic fields. A quantitative estimate of this effect explains the observed spectrum of 3C279. Our prediction can be tested in the near future by the satellite-borne GLAST detector as well as by the ground-based Imaging Atmospheric Cherenkov Telescopes H.E.S.S., MAGIC, CANGAROO III, VERITAS and by the Extensive Air Shower arrays ARGO-YBJ and MILAGRO. ", "introduction": "As is well known, in the very-high-energy (VHE) band above $100 \\, {\\rm GeV}$ the horizon of the observable Universe rapidly shrinks as the energy further increases. This comes about because photons from distant sources scatter off background photons permeating the Universe, thereby disappearing into electron-positron pairs~\\cite{stecker1971}. The corresponding cross section $\\sigma (\\gamma \\gamma \\to e^+ e^-)$ peaks where the VHE photon energy $E$ and the background photon energy $\\epsilon$ are related by $\\epsilon \\simeq (500 \\, {\\rm GeV}/E) \\, {\\rm eV}$. Therefore, for observations performed by Imaging Atmospheric Cherenkov Telescopes (IACTs) -- which probe the energy interval $100 \\, {\\rm GeV} - 100 \\,{\\rm TeV}$ -- the resulting cosmic opacity is dominated by the interaction with ultraviolet/optical/infrared diffuse background photons (frequency band $1.2 \\cdot 10^{3} \\, {\\rm GHz} - 1.2 \\cdot 10^{6} \\, {\\rm GHz}$, corresponding to the wavelength range $0.25 \\, \\mu {\\rm m} - 250 \\, \\mu {\\rm m}$), usually called Extragalactic Background Light (EBL), which is produced by galaxies during the whole history of the Universe. Neglecting evolutionary effects for simplicity, photon propagation is controlled by the photon mean free path ${\\lambda}_{\\gamma}(E)$ for $\\gamma \\gamma \\to e^+ e^-$, and so the observed photon spectrum $\\Phi_{\\rm obs}(E,D)$ is related to the emitted one $\\Phi_{\\rm em}(E)$ by \\begin{equation} \\label{a1} \\Phi_{\\rm obs}(E,D) = e^{- D/{\\lambda}_{\\gamma}(E)} \\ \\Phi_{\\rm em}(E)~. \\end{equation} Within the energy range in question, ${\\lambda}_{\\gamma}(E)$ decreases like a power law from the Hubble radius $4.2 \\, {\\rm Gpc}$ around $100 \\, {\\rm GeV}$ to $1 \\, {\\rm Mpc}$ around $100 \\, {\\rm TeV}$~\\cite{CoppiAharonian}. Thus, Eq.~(\\ref{a1}) entails that the observed flux is {\\it exponentially} suppressed both at high energy and at large distances, so that sufficiently far-away sources become hardly visible in the VHE range and their observed spectrum should anyway be {\\it much steeper} than the emitted one. Yet, observations have {\\it not} detected the behaviour predicted by Eq.~(\\ref{a1}). A first indication in this direction was reported by the H.E.S.S. collaboration in connection with the discovery of the two blazars H2356-309 ($z = 0.165$) and 1ES1101-232 ($z = 0.186$) at $E \\sim 1 \\, {\\rm TeV}$~\\cite{aharonian:nature06}. Stronger evidence comes from the observation of blazar 3C279 ($z = 0.536$) at $E \\sim 0.5 \\, {\\rm TeV}$ by the MAGIC collaboration~\\cite{3c}. In particular, the signal from 3C279 collected by MAGIC in the region $E<220$ GeV has more or less the same statistical significance as the one in the range 220 GeV $< E <$ 600 GeV ($6.1 \\sigma$ in the former case, $5.1 \\sigma$ in the latter). A suggested way out of this difficulty relies upon the modification of the standard Synchro-Self-Compton (SSC) emission mechanism. One option invokes strong relativistic shocks~\\cite{Stecker2007}. Another rests upon photon absorption inside the blazar~\\cite{Aharonian2008}. While successful at substantially hardening the emission spectrum, these attempts fail to explain why {\\it only} for the most distant blazars does such a drastic departure from the SSC emission spectrum show up. Our proposal -- usually referred to as the DARMA scenario -- is quite different~\\cite{drm}. Implicit in previous considerations is the hypothesis that photons propagate in the standard way throughout cosmological distances. We suppose instead that photons can oscillate into a new very light spin-zero particle -- named Axion-Like Parlicle (ALP) -- and vice-versa in the presence of cosmic magnetic fields, whose existence has definitely been proved by AUGER observations~\\cite{auger}. Once ALPs are produced close enough to the source, they travel {\\it unimpeded} throughout the Universe and can convert back to photons before reaching the Earth. Since ALPs do not undergo EBL absorption, the {\\it effective} photon mean free path ${\\lambda}_{\\gamma , {\\rm eff}} (E)$ gets {\\it increased} so that the observed photons cross a distance in excess of ${\\lambda}_{\\gamma}(E)$. Correspondingly, Eq. (\\ref{a1}) becomes \\begin{equation} \\label{a1bis} \\Phi_{\\rm obs}(E,D) = e^{- D/{\\lambda}_{\\gamma , {\\rm eff}}(E)} \\ \\Phi_{\\rm em}(E)~, \\end{equation} from which we see that even a {\\it slight} increase of ${\\lambda}_{\\gamma , {\\rm eff}}(E)$ gives rise to a {\\it huge} enhancement of the observed flux. It turns out that the DARMA mechanism makes ${\\lambda}_{\\gamma , {\\rm eff}}(E)$ shallower than ${\\lambda}_{\\gamma}(E)$ although it remains a decreasing function of $E$. So, the resulting observed spectrum is {\\it much harder} than the one predicted by Eq. (\\ref{a1}), thereby ensuring agreement with observations even for a {\\it standard} SSC emission spectrum. As a bonus, we get a natural explanation for the fact that only the most distant blazars would demand $\\Phi_{\\rm em}(E)$ to substantially depart from the emission spectrum predicted by the SSC mechanism. Our aim is to review the main features of our proposal as well as its application to blazar 3C279. ", "conclusions": "" }, "0911/0911.0927_arXiv.txt": { "abstract": "We demonstrate that stars beyond the virial radii of galaxies may be generated by the gravitational impulse received by a satellite as it passes through the pericenter of its orbit around its parent. These stars may become energetically unbound (escaped stars), or may travel to further than a few virial radii for longer than a few Gyr, but still remain energetically bound to the system (wandering stars). Larger satellites (10-100\\% the mass of the parent), and satellites on more radial orbits are responsible for the majority of this ejected population. Wandering stars could be observable on Mpc scales via classical novae, and on 100 Mpc scales via SNIa. The existence of such stars would imply a corresponding population of barely-bound, old, high velocity stars orbiting the Milky Way, generated by the same physical mechanism during the Galaxy's formation epoch. Sizes and properties of these combined populations should place some constraints on the orbits and masses of the progenitor objects from which they came, providing insight into the merging histories of galaxies in general and the Milky Way in particular. ", "introduction": "One of the triumphs of the last decade is orders of magnitude increase in the numbers of stars catalogued by large sky surveys such as the Sloan Digital Sky Survey and the Two Micron All Sky Survey. These catalogs contain significant samples of stars out to $\\sim$100kpc from the center of the Galaxy, allowing these regions to be probed through number counts to far lower surface brightness than previously possible and resulting in the discovery of numerous new dwarf satellite galaxies \\citep[e.g.][]{2005ApJ...626L..85W,2006ApJ...647L.111B}, as well as tidal tails from dwarf galaxies possibly disrupted long ago \\citep[e.g.][]{2003ApJ...599.1082M, 2005MNRAS.357...17B}. We move from this decade of large sky surveys into a decade of continual all-sky surveying with the introduction of the Panoramic Survey Telescope \\& Rapid Response System \\citep[Pan-STARRS, see][]{2004SPIE.5489...11K} and the prospect of the Large Synoptic Survey Telescope \\citep[LSST, see ][]{ 2007AAS...210.6605I}. These surveys will be sensitive to ever fainter magnitude objects and hence probe ever lower surface brightness structures in ever increasing volumes of space. The repeated nature of these surveys also adds the extra dimension of time, enabling large-sample statistical studies of variable phenomena in a way that was not previously possible. For example, \\cite{2006AJ....131.2980S} points out that LSST will detect many new classical novae (hereafter CN). While interesting in their own right, the special relationship between the peak brightness and decline time relationship of CN, coupled with their high luminosity, will allow us to map a volume out to 40 Mpc from Earth. This will be the first map of stars between galaxies and outside of clusters or groups of galaxies (i.e. intercluster stars) ever produced. But should we expect to find anything there? Red giant stars, planetary nebulae and classical novae have been detected outside of galaxies but {\\it within clusters} \\citep[i.e. intracluster light - hereafter ICL, for example see][]{2004MNRAS.355..159W,2005ApJ...618..692N}. These populations are thought to begin their lives in galaxies: stars are liberated from their original hosts during galactic collisions, galactic harassment and tidal shredding, which generate tails of debris that can be ripped from the galaxy system by the tidal field of the cluster \\citep{1996Natur.379..613M}. This evolutionary picture is confirmed by the general agreement of observations and simulations on the level of ICL \\citep{2004ApJ...607L..83M},which photometric studies define as $0 - 50\\%$ of the total cluster light \\citep{2004bdmh.confE..26A,2004IAUS..217...86F,2004MNRAS.355..159W}. The same dynamical processes can be appealed to on smaller scales to explain the existence of ICL around less massive, looser clusters and galaxy groups \\citep{2004bdmh.confE..26A}, as well as diffuse stellar halos around individual galaxies \\citep{2005ApJ...635..931B}. Thus stellar halos and the ICL can be thought of as testaments to galactic interactions during the epochs of galaxy and cluster formation. We explore whether dynamical interactions could plausibly lead to stars {\\it beyond} the virial radii of the parent galaxy, group or cluster, and, if so, what the existence of such a population might tell us about the nature of hierarchical structure formation. \\citet{1992ApJ...397L..75S} have already shown that N-body simulations of mergers of massive galaxies can give rise to unbound particles that form a gaussian tail to the energy distribution, in agreement with a statistical mechanical description of violent relaxation. Here, we develop a simple description of interactions that allows us to illustrate how the size of the escaped population should depend on the mass ratio and orbits of the progenitor systems. In \\S \\ref{Mechanism for Escape} we establish that a tidal impulse at perigalacticon is a viable path for stars from satellite dwarf galaxies to escape from the combined dark matter potentials of the parent and satellite galaxies. In \\S \\ref{Illustration} we first use restricted three-body simulations to test the mechanism described in \\S \\ref{Mechanism for Escape} and then analyze N-body results to establish trends, and quantitative expectations for the escaping population. In \\S \\ref{Discussion} we discuss the implications of our results for future surveys of giant stars, CN, supernovae and high velocity stars. In \\S \\ref{Conclusions} we summarize our conclusions. ", "conclusions": "We've outlined how tidal interactions can eject stars which are loosely bound to a satellite galaxy to beyond the virial radius of the satellite's parent. We determined that more massive satellites contribute the most to this escaped (entirely unbound from the satellite/parent system) or wandering (bound, but traveling beyond 2 virial radii for more than $\\sim$Gyr) population of stars, as do satellites on highly radial orbits. The majority of this population of stars originated during the epoch of galaxy formation, when large interacting satellites were more frequently infalling on radial orbits, approximately from a redshift of around 4 to 3. If this is indeed the case, we expect an old, uniform (to first order) population of escaped and wandering stars to exist beyond a few virial radii of the Milky Way. Subcategories (e.g. classical novae, supernovae) of this population should be detectable by LSST and Pan-STARRS. We estimate a lower limit of 0.05\\% of stars to be members of this population. An accurate number prediction warrants further investigation using full cosmological simulations which follow interacting subhalos as they form the parent halo. \"Not all those that wander are lost\"-Tolkien" }, "0911/0911.2780_arXiv.txt": { "abstract": " ", "introduction": "Primordial density perturbations are traditionally described by a Gaussian distribution, characterised by an almost scale-invariant power spectrum. However the detailed information about the primordial density perturbations over a range of cosmological scales offers the opportunity to test in detail the nature of the primordial perturbations, both their scale-dependence and Gaussianity \\cite{Komatsu:2008}. The local model for non-Gaussianity has proved a remarkably popular description of possible deviations from a purely Gaussian distribution of primordial perturbations. In the simplest case the primordial Newtonian potential includes a contribution from both the local value of a linear Gaussian field and a quadratic term proportional to the square of the local value of the Gaussian field: \\be \\label{originalPhi} \\Phi(\\vec{x}) = \\varphi_G(\\vec{x}) + \\fNLlocal \\left( \\varphi_G^2(\\vec{x}) - \\langle\\varphi_G^2\\rangle \\right) \\,, \\ee where ${f}^{\\rm local}_{\\rm NL}$ is a dimensionless parameter characterising the deviations from Gaussianity~\\cite{Komatsu:2001rj}. $\\langle\\varphi_G^2\\rangle$ denotes the ensemble average, or equivalently the spatial average in a statistically homogeneous distribution. Note that following Komatsu and Spergel~\\cite{Komatsu:2001rj} we adopt a sign convention for the Newtonian potential $\\Phi$ which is the opposite of that used, e.g., by Mukhanov et al~\\cite{Mukhanov:1990me}. In Fourier space we define the power spectrum and bispectrum as \\bea \\langle\\Phi_{\\vec{k_1}}\\Phi_{\\vec{k_2}} \\rangle = (2\\pi)^3 P_{\\Phi}(k_1) \\delta^{3}(\\vec{k_1}+\\vec{k_2}) \\,,\\\\ \\langle \\Phi_{\\vec{k_1}} \\Phi_{\\vec{k_2}} \\Phi_{\\vec{k_3}} \\rangle = (2\\pi)^3 B_{\\Phi}(k_1,k_2,k_3) \\delta^{3}(\\vec{k_1}+\\vec{k_2}+\\vec{k_3}) \\,, \\eea and the amplitude of the bispectrum relative to the power spectrum is conventionally given by the non-linearity parameter \\be \\label{deffnl} \\fnl(k_1,k_2,k_3) \\equiv \\frac{B_{\\Phi}(k_1,k_2,k_3)}{2 \\left[ P_{\\Phi}(k_1) P_{\\Phi}(k_2) + P_{\\Phi}(k_2) P_{\\Phi}(k_3) + P_{\\Phi}(k_3) P_{\\Phi}(k_1) \\right]} \\,. \\ee In the special case of the local model (\\ref{originalPhi}) we have $\\fnl=\\fNLlocal$ and it is clear that $\\fnl$ is, by construction, a constant parameter independent of spatial position or scale. This local model turns out to be a very good description of non-Gaussianity in some simple physical models for the origin of structure. In particular it can describe the primordial density perturbation on super-Hubble scales predicted, up to second-order, using the $\\delta N$-formalism \\cite{starob85,SS,LR} if the local integrated expansion is a function of a Gaussian random field, $N(\\sigma_g)$, at some initial time, $\\sigma_g(\\vec{x})=\\sigma(t_i,\\vec{x})$. We then have \\be \\label{Phizeta} \\Phi = \\frac35 \\zeta = \\frac35 \\left[ N - \\langle N \\rangle \\right] \\,. \\ee A good example is provided by the simplest curvaton scenario~~\\cite{Enqvist:2001zp,Lyth:2001nq,Moroi:2001ct}, where quantum fluctuations of a weakly-coupled field during inflation are well described by a Gaussian random field and the primordial density perturbation is determined by the density of the curvaton field when it decays, which is proportional the the square of the initial local field~\\cite{Linde:1996gt,Lyth:2002my}. In this paper we consider extensions of the simple local model (\\ref{originalPhi}). In particular we will characterise and quantify the scale-dependence of the parameter $\\fnl$ which arise in realistic inflationary models for the origin of structure. Scale-dependence arises due to two key features. Firstly we examine a multi-variate local model where the local expansion is a quadratic function of more than one Gaussian random field. In this case scale-dependent $\\fnl$ can arise if the Gaussian fields have differing scale-dependence, leading to a change with scale in the correlation of quadratic terms with the linear perturbation. Secondly, we show that scale-dependent $\\fnl$ can arise even when the expansion is a function of a single canonical scalar field. In this case, it is due to the development of intrinsic non-Gaussianity associated with the non-linear evolution of the initially Gaussian fluctuations after Hubble-exit. We will refer to this case as a quasi-local model. In the case of the so-called equilateral $\\fnl$, which arises for example from DBI inflation (see e.g. \\cite{DBI}), the scale dependence has already been quite well studied both in theoretical models \\cite{Chen:2005fe,Khoury:2008wj,Byrnes:2009qy,Leblond:2008gg} and forecasts have been made for future observational prospects \\cite{LoVerde:2007ri,Sefusatti:2009xu}. For a discussion of the different possible shape dependences of the bispectrum see \\cite{Fergusson:2008ra}. However for local-type models there has been little previous consideration of a scale dependence. The scale dependence of local-type $\\fnl$ was first calculated in Byrnes et al \\cite{Byrnes:2008zy} for a specific model of hybrid inflation with two-fields. It was found that although the scale dependence is slow-roll suppressed, it depends on a particular combination of slow-roll parameters which is not negligible and can easily be larger than the spectral index of the power spectrum. The scale dependence of $\\fnl$ was considered in the case of an ekpyrotic universe in \\cite{Khoury:2008wj}. In the case of an exact solution of two-field inflation, which can give rise to a large non-Gaussianity, it was shown that $\\fnl$ is scale independent \\cite{Byrnes:2009qy}. The observational prospects for local-type models were considered in \\cite{Sefusatti:2009xu} which showed that the CMB data is sensitive to a scale dependence of $\\fnl$. However they used a very simple Ansatz for the scale-dependent $\\fnl$: here we calculate for the first time the full scale dependence of a scale-dependent quasi-local $\\fnl$. We note that higher-order contributions to the primordial perturbations are expected beyond quadratic order. These only affect the bispectrum at subleading order in the scalar field perturbations, although in some cases they might provide the dominant contribution to observables \\cite{Boubekeur:2005fj}. In the language of \\cite{Byrnes:2007tm} they are loop corrections. The effective scale-dependence of $\\fnl$ due to higher-order terms is examined explicitly in \\cite{Kumar:2009ge} and was considered previously, for example, in \\cite{Suyama:2008nt}. Notice that the results depend at leading order on the infra-red cut-off, and that no infra-red complete theory is yet known \\cite{Seery:2009hs}. It would be interesting to further develop this interesting issue in a case in which the infra-red effects are well-understood. Note that the loop term is not well described as having constant scale dependence, unlike the cases we consider in this paper. We also note that \\cite{Kumar:2009ge} only consider the scale dependence coming from the logarithmic term which depends on the cut-off, while the other terms which multiply this will also have a scale dependence in general. This scale dependence can be calculated using the methods presented in this paper. This paper is organized as follow. In Section \\ref{mvarsec}, we focus on multi-variate local models, which are models that contain contributions to the primordial curvature perturbation from more than one Gaussian field, in which the scale dependence of $\\fnl$ is due to the different scale dependences of the fields that drive local expansion. As an example of this we study a mixed inflaton and curvaton scenario in sec.~\\ref{sec:inflatonandcurvaton}. In Section \\ref{sfsec}, we show that scale dependence of $\\fnl$ can arise also for single field models, due to the non-linear evolution of initially Gaussian fluctuations after Hubble exit. In Section \\ref{multisec} we extend the discussion to systems involving multiple fields, taking into account the non-linear evolution of fluctuations after horizon crossing in this context. We conclude in sec.~\\ref{concsec}. ", "conclusions": "\\label{sec:conclusion}\\label{concsec} We have studied generalisations of the local model of non-Gaussianity, thereby seeing in which cases the standard assumption that it can be described by a single, scale-independent parameter is valid. We have shown that this is only strictly valid in specific models, where the primordial curvature perturbation is generated by a single scalar field which acts as an isocurvature (``test'') field during inflation and has a quadratic potential. An example of this is the simplest curvaton scenario. However as soon as the scalar field has interactions these generate non-Gaussianity of the field perturbations after Hubble exit, and these give rise to a scale dependent non-Gaussianity which we call quasi-local, meaning that in the limit that the scale dependence of $\\fnl$ goes to zero one recovers the local model. We have also shown that even in a multi-variate local model where more than one Gaussian field contributes to the primordial curvature perturbation the effective $\\fnl$ has a scale dependence unless all fields have the same scale dependence. An example of this is a mixed inflaton and curvaton scenario or a multi-curvaton model, where the curvatons have quadratic potentials with different masses. We have developed a formalism, based on the $\\delta N$ approach, which allows us to obtain compact expressions for $\\nnl=\\,d \\ln{|\\fnl|}/d \\ln{k}$, in both the single and multi-field case. We have also discussed more generally the validity of defining a scale dependence of $\\fnl$ with respect to a single scale when the bispectrum is generally a function of three variables, except in the case of an equilateral triangle. We have shown that the scale dependence of $\\fnl$ is independent of the shape of the triangle it describes, provided that we consider scale variations which preserve its shape. It is then appropriate to perform the simpler calculation of $\\fnl$ for an equilateral triangle, and use this to calculate $\\nnl$ by taking its logarithmic scale dependence. We applied our formalism to discuss several specific models. Our results suggest that the scale dependence is typically first order in slow roll. The precise value however depends quite sensitively on details of the model which, makes it in principle possible to use $\\nnl$ to discriminate observationally between different models. We have shown that while the simplest realisation of the curvaton model has an exactly scale independent $\\fnl$, almost any extension to a more realistic scenario does give rise to a scale dependence. In the case of an interacting curvaton or a multiple-curvaton scenario the scale dependence is likely to be suppressed compared to the spectral index. This is because $\\nnl$ in these scenarios depends schematically on $\\eta_{{\\rm curvaton}}$ which is normally positive and, in light of the observational preference for a red spectrum, this is likely to be small. A detection of both $\\fnl$ and $\\nnl$ would therefore be a signal of non-trivial dynamics in the curvaton scenario, or that the inflaton perturbations have also made a significant contribution to the observed power spectrum. In the case of a mixed scenario, in which the primordial curvature perturbation has contributions from both the inflaton and curvaton perturbations it is quite natural for there to be a consistency relation between the scale dependence of the power spectrum and the bispectrum, $\\nnl\\simeq-2(n-1)$, and this could be observable with Planck assuming a large enough fiducial value for $\\fnl$, close to the current observational bounds. In this case the bispectrum would be larger on small scales and this could make a detection of $\\fnl$ using large scale structure data more likely. In any scenario where one includes the Gaussian perturbations of the inflaton field as well as the perturbations of a second field which generates non-Gaussianities one will have a scale dependent $\\fnl$, unless both fields have exactly the same spectral index. For the case where non-Gaussianity is generated during slow-roll hybrid inflation one generally has a non-negligble $\\nnl$ \\cite{Byrnes:2008zy}. In this case, since the magnitude of $\\fnl$ can only grow during inflation, larger scales, which exit the horizon earlier, will be more non-Gaussian and one necessarily has $\\nnl<0$. It should be straightforward to extend the formalism we have presented here to the primordial trispectrum (the four-point function) \\cite{Seery:2006js, Byrnes:2006vq} and we expect that the two non-linearity parameters which model it, $\\tau_{NL}$ and $g_{NL}$, would also have a first order in slow roll scale dependence. We have chosen the notation $\\nnl$ in a way that is extendible in an obvious manner to study this. For example in the case that a single field generates the primordial curvature perturbation one has a non-Gaussianity consistency relation, $\\tau_{NL}=(6\\fnl/5)^2$. Then assuming our formalism can be extended in the obvious way it follows that $n_{\\tau_{NL}}=2\\nnl$." }, "0911/0911.5396_arXiv.txt": { "abstract": "{ We consider cosmologies in which a dark-energy scalar field interacts with cold dark matter. The growth of perturbations is followed beyond the linear level by means of the time-renormalization-group method, which is extended to describe a multi-component matter sector. Even in the absence of the extra interaction, a scale-dependent bias is generated as a consequence of the different initial conditions for baryons and dark matter after decoupling. The effect is enhanced significantly by the extra coupling and can be at the 2-3 percent level in the range of scales of baryonic acoustic oscillations. We compare our results with N-body simulations, finding very good agreement. \\\\ PACS numbers: 95.35.+d, 95.36.+x, 98.80.Cq } \\newpage ", "introduction": "\\setcounter{equation}{0} The power spectrum of matter perturbations reflects the evolution of the Universe since the time of matter-radiation equality. For given initial conditions, determined by the primordial spectrum (usually assumed to be scale invariant), the growth of perturbations depends on the cosmological scenario. The calculation of the present matter power spectrum can constrain this scenario through the comparison of the deduced spectrum with the observed large-scale structure. A major technical difficulty in the realization of such a program is the failure of linear perturbation theory to describe present-day fluctuations with characteristic length scales below roughly 100 Mpc. At length scales below about 10 Mpc, the evolution is highly non-linear, so that only numerical N-body simulations can capture the dynamics of the formation of galaxies and clusters of galaxies. Fluctuations with length scales of around 100 Mpc fall within the mildly non-linear regime, for which analytical methods have been developed. These scales (corresponding to wavenumbers in the $0.03-0.25\\,h$/Mpc range) are of particular interest, because they correspond to the sound horizon at decoupling, which can be determined by reconstructing the oscillatory behavior of the matter power spectrum due to baryonic acoustic oscillations (BAO). The various analytical methods \\cite{CrSc1}--\\cite{matsubara} essentially amount to resummations of subsets of perturbative diagrams of arbitrarily high order, in a way analogous to the renormalization group (RG). In this work we shall follow the approach of \\cite{Max2}, named time-RG or TRG. In the context of the RG the various observables depend on a characteristic energy scale, and evolve as this scale is varied. The TRG uses time as the flow parameter that describes the evolution of physical quantities, such as the spectrum of perturbations. The method is characterized by conceptual simplicity. It has been applied to ${\\rm \\Lambda CDM}$ and quintessence cosmologies \\cite{Max2}, as well as models with massive neutrinos \\cite{lesgourgues}. The fundamental equations in the TRG approach are the ``equations of motion'', i.e. the continuity, Euler and Poisson equations. From these, equations can be derived for the time evolution of correlation functions for the density and velocity fields. The various spectra are obtained through appropriate Fourier transforms of the correlation functions. The method results in a coupled infinite system of integro-differential equations for the time evolution of the spectrum, bispectrum etc. The crucial approximation, that makes a solution possible, is to neglect the effect of the higher-order correlation functions in the evolution equations of the lower-order ones. The calculations performed so far take into account the spectrum and bispectrum and set the higher-level spectra to zero. The procedure of truncating the system of equations is commonly employed in the applications of the Wilsonian RG to field theory or statistical physics. (For a review, see \\cite{erg}.) The accuracy of the calculation can be determined either by enlarging the truncated system (by including the trispectrum, for example) and examining the stability of the results, or by comparing with alternative methods. The second approach is often followed, because enlarging the truncation can increase the complexity of the calculation considerably. In the case of the TRG, the agreement with results from N-body simulations for ${\\rm \\Lambda CDM}$ has been confirmed \\cite{Max2}. Also, a comparative analysis of several analytical methods, using N-body simulations as a reference, has been carried out in ref. \\cite{carlson}. The study demonstrates that the TRG remains accurate at the 1-2\\% level over the whole BAO range at all redshifts. In fig. 5 of ref. \\cite{carlson} the deviation of the TRG prediction from a reference spectrum derived through simulations for ${\\rm \\Lambda CDM}$ is depicted. At a redshift $z=1$ the deviation is at the 1\\% level, or smaller, over the whole depicted range $0\\leq k \\leq 0.15\\,h$/Mpc. For $z=0$ the deviation may exceed 1\\% for some values of $k$, but stays below 2\\% over the whole depicted range. Based on these findings and the analysis of \\cite{Max2} we estimate an accuracy of 1\\% for $z=1$ and 2\\% for $z=0$ over the range $0\\leq k \\leq 0.2\\,h$/Mpc. The accuracy of 1-loop standard perturbation theory (SPT) can be inferred from fig. 1 of ref. \\cite{carlson}. At $z=1$ the accuracy is at the 1\\% level for $0\\leq k \\leq 0.1\\,h$/Mpc, while at $z=0$ it is at the 2\\% level for $0\\leq k \\leq 0.05\\,h$/Mpc. We depict these ranges in figs. \\ref{ddcc}--\\ref{bias1} of the current paper. The additional approximations that we make in this paper for the study of models with non-zero coupling between dark matter and dark energy induce uncertainties at the sub-percent level. We discuss this issue in detail in subsection 3.1. For this reason the level of accuracy of our results for the power spectra in the coupled case is expected to be similar to that for ${\\rm \\Lambda CDM}$. Its magnitude is set by the truncations in the evolution equations (the omission of the effect of the trispectrum and higher spectra), which are similar in all models. The purpose of the present work is threefold: \\\\ 1) We extend the formalism to more complicated models. We introduce two modificiations to previous studies: a) Within the matter sector we allow for an arbitrary number of species with independent spectra. These include baryonic matter (BM) and cold (pressureless) dark matter (CDM). One may also consider contributions from massive neutrinos etc. b) We allow for an interaction between CDM and dark energy (DE). We consider a class of quintessence models, in which there is direct coupling between CDM and the quintessence field. The form of the interaction is a generalization of the universal coupling to all species present in scalar-tensor theories in the Einstein frame. It is modelled through the dependence of the mass of the CDM particles on the quintessence field \\cite{wetcosmon}-\\cite{baccigalupi}. \\\\ 2) We test the accuracy of the method in this enlarged framework by comparing with available N-body simulations. We perform our numerical analysis for a model for which results from simulations are given in ref. \\cite{baldi}. In the context of coupled quintessence, the cosmological evolution can be very diverse \\cite{amendola,baccigalupi,perturbations1,perturbations2}. It is very time-consuming to study exhaustively every model through N-body simulations. Our approach provides an alternative method, which can be much faster, while retaining the necessary accuracy. It is also important to note that, while the N-body simulations are highly accurate at the length scales of galaxies and clusters of galaxies, they are less accurate in the BAO range because of the required large volumes. On the other hand, analytical methods, like ours, are more accurate in the quasi-linear regime of large length scales. The two approaches can be viewed as complementary. \\\\ 3) We provide predictions for observables for which non-linear corrections can be important. As such, we study the bias between dark and baryonic matter in the BAO range for models of coupled quintessence. The couplings between the matter sector and DE are constrained by observations. For the BM-DE coupling, the bound from the Cassini spacecraft \\cite{Cassini} limits its order of magnitude to be below $10^{-3}$. As such small couplings produce negligible effects on the power spectrum, we assume that the BM-DE coupling is exactly zero. The coupling between CDM and DE is constrained by various considerations, such as the modification of the spectrum of the cosmic microwave background (CMB) or that of the matter distribution. A common feature of the class of models we are considering is that the presence of an additional long-range force between CDM particles, induced by the DE field, modifies their clustering properties. The various observable consequences have been discussed in the literature \\cite{largescale}--\\cite{baldi}. The strength of the CDM-DE interaction is constrained through the comparison with the observed CMB and matter spectra. It has been shown that, for particular models, the CDM-DE coupling must be considerably smaller than the gravitational one \\cite{bean,lavacca}. Such constraints cannot be considered generic, because the evolution of the cosmological background and the perturbations around it varies considerably from model to model. We work within the model of \\cite{baldi}, because our main objective is to compare with the results of N-body simulations presented there. The couplings that we consider are roughly consistent with the bounds deduced in \\cite{bean,lavacca} for a model similar to ours. In this work we follow a novel approach for the derivation of the fundamental equations. We derive the continuity, Euler and Poisson equations on an expanding background, starting from the conservation of the energy-momentum tensor. In order to cast these equations in a form that generalizes the standard expressions on a static background, we need to impose a certain hierarchy between the density perturbations, the velocity field and the potentials. Our derivation makes it straightforward to generalize these equations in future studies in order to take into account non-zero pressure and higher-order terms. Next, we derive the system of differential equations for the spectrum and bispectrum, within a truncation that neglects higher-level spectra. We integrate these equations numerically in order to produce the non-linear spectra at low redshift and compare with the results of N-body simulations. We study in detail the difference, usually characterized as bias, between the spectra of dark and baryonic matter. We find that the CDM-DE coupling enhances significantly the bias of the decoupled case. In the following section we derive the evolution equations for the spectra of dark and baryonic matter. The details of the derivation for the case of one massive component are given in appendix A. The generalization for an arbitrary number of massive components is presented in appendix B. The results of the numerical integration of the evolution equations are presented in section 3. We compare them with the results of N-body simulation. We also discuss in detail the form of the bias in the BAO range. ", "conclusions": "\\setcounter{equation}{0} In this paper we have extended the TRG formalism introduced in ref.~\\cite{Max2} in two respects. Firstly, we described the matter sector of the theory keeping track of the CDM and BM components separately, instead of treating them as a single fluid. Secondly, we introduced a new scalar force that couples differently to CDM and BM. As was discussed recently in \\cite{smith}, an accurate computation of the evolution of the different components is mandatory if one wants to achieve high precision modeling of structure formation. In order to test the accuracy of the TRG method for this more general class of cosmological scenaria, we analyzed the same cosmologies considered in \\cite{baldi}, and compared our results with those of the N-body simulations presented there. The agreement is very good up to $k \\simeq 0.5 \\,\\mathrm{h/Mpc}$ at $z=0$, where the non-linear power spectrum is roughly twice the linear one. These results confirm the reliability of the TRG as a computational tool that can fill the gap between linear and non-linear scales. Even in the absence of an extra force, or in the case that the extra force couples universally to all matter species, the evolution of BM and CDM differs as a consequence of different initial conditions after decoupling. In linear theory, the initial unbalance between BM and CDM fluctuations is almost completely washed out by the present epoch. However, when non-linearities are taken into account the bias persists. We have seen that the effect is at the percent level in the BAO range at low redshift in the uncoupled case, but it may grow up to the $2.5\\%$ level when a non-zero coupling is turned on with a value compatible with present bounds (obtained within the linear approximation \\cite{bean,lavacca}). As a result, these models will receive significant constraints from future galaxy surveys, which aim to measure the power spectrum within the BAO range with an accuracy at the percent level. A thorough investigation of this model-dependent issue goes beyond the purpose of this paper and is postponed for future work. The most interesting outcome of our analysis from the point of view of observations concerns the form of the bias between CDM and BM spectra. In linear theory, the bias is {\\it scale independent} in recent epochs \\cite{tocchini}. The non-linear corrections remove this feature, and make the bias {\\it scale dependent}, consistently with the results of \\cite{baldi}. We confirm and extend these results for the BAO region ($0.03\\,h$/Mpc $\\lta k \\lta 0.25\\, h$/Mpc), where we expect the TRG method to be reliable. The effect becomes more pronounced with increasing coupling between CDM and DE (assuming that the respective coupling for BM vanishes). The form of the bias provides an additional observational handle for the differentiation between various models. For example, in models with massive neutrinos the growth of the spectrum induced by the non-linearities at small length scales is compensated by the free-streaming of neutrinos \\cite{lesgourgues}. On the other hand, the bias is expected to remain scale-dependent at the non-linear level if there is a substantial CDM-DE coupling." }, "0911/0911.3390_arXiv.txt": { "abstract": "We have investigated the evolution of a pair of interacting planets embedded in a gaseous \\langeditorchanges{disc,} considering \\langeditorchanges{the} possibility of the resonant capture of a Super-Earth by a Jupiter mass gas giant. First, we have examined the situation where the Super-Earth is on the internal orbit and the gas giant on the external one. It has been found that the terrestrial planet is scattered from the disc or the gas giant captures the Super-Earth into an interior 3:2 or 4:3 \\langeditorchanges{mean-motion} resonance. The stability of \\langeditorchanges{the latter configurations} depends on the initial planet positions and \\langeditorchanges{on} eccentricity evolution. The behaviour of the system is different if the Super-Earth is the external planet. We have found that instead of being captured in the \\langeditorchanges{mean-motion} \\langeditorchanges{resonance,} the terrestrial planet is trapped at the outer edge of the gap opened by the gas giant. \\langeditorchanges{This effect} prevents \\langeditorchanges{the} occurrence of the first order \\langeditorchanges{mean-motion} commensurability. \\langeditorchanges{These} results are particularly interesting in light \\langeditorchanges{of} recent exoplanet discoveries and provide predictions of what will become observationally testable in the near future. ", "introduction": "So far we know only \\langeditorchanges{a} few Super-Earths, \\langeditorchanges{i.e.,} planets less massive than $10 M_\\oplus$. However, there is a good chance that in the near future, COROT and \\langeditorchanges{the} Kepler mission will find more terrestrial type planets. An interesting possibility is the existence \\langeditorchanges{of} Super-Earths close \\langeditorchanges{to} Jupiter-like planets. \\langeditorchanges{Different} mass objects embedded in a protoplanetary disc will migrate with different rates. The final configurations will depend on the intricate interplay among many physical processes including planet-planet, disc-planet and planet-star interactions. Here we focus on the possible resonant configurations of a gas giant and a Super-Earth orbiting a Solar-type star. The early divergent conclusions concerned with the occurrence \\langeditorchanges{of} terrestrial planets \\langeditorchanges{in} hot Jupiter systems (Armitage, \\cite{armitage}, Raymond et al., \\cite{raymond05}) have been clarified by the most recent studies (Fogg \\& Nelson, \\cite{fonel07a}, \\cite{fonel07b}, Raymond et al., \\cite{raymond}), which predict that terrestrial planets can grow and be retained \\langeditorchanges{in} hot-Jupiter \\langeditorchanges{systems,} both interior and exterior to the gas giant. Here we consider the evolution of an already formed Super-Earth and a gas giant, both embedded in a gaseous disc. In our studies we are interested in the possibility of the formation of resonant planetary configurations. First, we have examined the behaviour of the system when the terrestrial planet is on the internal orbit and the gas giant on the external one. Then we have reversed the situation such that the Jupiter is the \\langeditorchanges{internal} planet and the Super-Earth is \\langeditorchanges{the external planet}. % \\langeditorchanges{Two-dimensional} hydrodynamical simulations have been employed in order to determine the occurrence \\langeditorchanges{of first-order mean-motion} resonances as the outcome \\langeditorchanges{of} convergent orbital migration. Our predictions have \\langeditorchanges{interesting} astrobiological \\langeditorchanges{implications,} which we discuss in \\langeditorchanges{the Conclusions section}. ", "conclusions": "A large-scale orbital migration \\langeditorchanges{in} young planetary systems might play an important role in \\langeditorchanges{shaping} their architectures. \\langeditorchanges{Tidal} gravitational forces are able to rearrange the planet positions according to their masses and the disc parameters. The final configuration after \\langeditorchanges{disc} dispersal might be what we actually observe in the extrasolar systems. In particular, \\langeditorchanges{convergent} migration can bring planets into \\langeditorchanges{mean-motion} resonances \\langeditorchanges{as} has been found in Podlewska \\& Szuszkiewicz (\\cite{podszusz}) \\langeditorchanges{in the} case where the terrestrial planet is on the internal orbit. When the Super-Earth is on the external \\langeditorchanges{orbit,} it is captured at the outer edge of the gap opened by the gas giant (Podlewska \\& Szuszkiewicz, \\cite{paperII}). Despite \\langeditorchanges{the} absence of \\langeditorchanges{mean-motion} resonance, the trapping at the outer edge of the gap can slow down the migration of the \\langeditorchanges{Super-Earth.} Thus, depending on whether the Super-Earth is inside or outside \\langeditorchanges{the} gas giant \\langeditorchanges{orbit,} the most likely planet configurations will be different. We claim that the Super-Earth can survive in close proximity to the gas giant. \\langeditorchanges{The} terrestrial planet is \\langeditorchanges{either} locked in \\langeditorchanges{mean-motion} commensurability or it is captured in the trap which prevents \\langeditorchanges{fast} migration toward the star. Our results have an interesting implication \\langeditorchanges{for} astrobiological studies. Namely, if we have a gas giant in or near the habitable \\langeditorchanges{zone,} \\langeditorchanges{then a} terrestrial planet locked in the mean motion resonance or captured in the trap at the outer edge of the gap can be also located in the habitable zone. Some of the known extrasolar gas giants as well as the recently discovered Super-Earth in the Gliese 581 system (Udry et al., \\cite{udry}) are in the habitable \\langeditorchanges{zone.} We can expect that future observations will reveal terrestrial planets close to gas giants." }, "0911/0911.3445_arXiv.txt": { "abstract": "We present new high spatial resolution ($\\lesssim0$\\farcs1) 1--5\\mic\\ adaptive optics images, interferometric 1.3\\,mm continuum and $^{12}$CO 2--1 maps, and 350\\mic, 2.8 and 3.3\\,mm fluxes measurements of the {\\hv} system. Our adaptive optics images unambiguously demonstrate that {\\hvab}--C is a common proper motion pair. They further reveal an unusually slow orbital motion within the tight {\\hvab} pair that suggests a highly eccentric orbit and/or a large deprojected physical separation. Scattered light images of the {\\hvc} edge-on protoplanetary disk suggest that the anisotropy of the dust scattering phase function is almost independent of wavelength from 0.8 to 5\\mic, whereas the dust opacity decreases significantly over the same range. The images further reveal a marked lateral asymmetry in the disk that does not vary over a timescale of 2 years. We further detect a radial velocity gradient in the disk in our $^{12}$CO map that lies along the same position angle as the elongation of the continuum emission, which is consistent with Keplerian rotation around an 0.5--1$M_\\odot$ central star, suggesting that it could be the most massive component in the triple system. To obtain a global representation of the {\\hvc} disk, we search for a model that self-consistently reproduces observations of the disk from the visible regime up to millimeter wavelengths. We use a powerful radiative transfer model to compute synthetic disk observations and use a Bayesian inference method to extract constraints on the disk properties. Each individual image, as well as the spectral energy distribution, of {\\hvc} can be well reproduced by our models with fully mixed dust provided grain growth has already produced larger-than-interstellar dust grains. However, no single model can satisfactorily simultaneously account for all observations. We suggest that future attempts to model this source include more complex dust properties and possibly vertical stratification. While both grain growth and stratification have already been suggested in many disks, only a panchromatic analysis, such as presented here, can provide a complete picture of the structure of a disk, a necessary step towards quantitatively testing the predictions of numerical models of disk evolution. ", "introduction": "Circumstellar disks are an ubiquitous outcome of the stellar formation process and they are believed to be the birth place of planetary systems. The growth of dust particles towards planetesimal sizes along with their vertical settling due to gas drag are processes that are believed to be the first steps towards planet formation. Hydrodynamical models have shown that these processes can be efficient early in the disk evolution \\citep[e.g.,][]{weidenschilling97, dullemond04, barriere05}. To test these models, it is necessary to obtain an observation-based description of the structure and dust content of protoplanetary disks as a function of the age of the system and other relevant parameters (e.g., stellar mass). The dust component of protoplanetary disks has long been studied via its thermal emission from near-infrared to millimeter wavelengths which is frequently associated with low-mass pre-main sequence T\\,Tauri stars \\citep{kenyon87, bertout88, strom89, beckwith90}. Both grain growth and dust settling can alter the overall shape of the spectral energy distribution (SED) of a T\\,Tauri star \\citep{dalessio01, dullemond04, dalessio06}. Indeed, several studies that analyzed (elements of) the SED of young stars have concluded that both grain growth and settling is occuring in protoplanetary disks \\citep[e.g.,][and references therein]{beckwith91, mannings94, furlan06, kessler06, rodmann06, natta07}. Unfortunately, such studies suffer from the absence of spatial information inherent to photometric measurements and the high optical thickness of disks in the near- to mid-infrared regime. As a result, comparing an object's SED to radiative transfer models leaves many ambiguities \\citep{chiang01}. For instance, the inferred total dust mass and the maximum size of the dust grains are inversely correlated because of the dependency of dust opacity on grain size. In addition, most of these studies, which focus on a single type of observations (e.g., millimeter fluxes, silicate emission feature), only probe a limited region of the disk and a small fraction of the entire grain size distribution. To solve for the ambiguities inherent to SED studies, it is critical to obtain spatially-resolved observations. Such observations include thermal emission mapping with (sub)millimeter interferometers \\citep[e.g.,][]{keene90, simon92, lay94, dutrey96, andrews07} and scattered light imaging with optical and near-infrared high-resolution instruments \\citep[e.g.,][]{burrows96, roddier96, stapelfeldt98}. Disks are generally optically thin at long wavevengths, so the former type of observations can probe the entire disk structure. Furthermore, they are very sensitive to the presence of millimeter-sized particles. On the other hand, scattered light images, which only probe dust grains at the disk surface, are very sensitive to the size distribution of micronic grains especially when images at multiple wavelengths are analyzed simultaneously \\citep[][and references therein]{watson07ppv}. Both types of observations have already yielded important pieces of evidence supporting both grain growth and dust settling in disks \\citep[e.g.,][]{duchene03, duchene04, watson04}. While grain growth and settling appear to occur in protoplanetary disks, detailed {\\it quantitative} tests of hydrodynamical models can only be achieved with a detailed view of the entire structure of a disk. This can only be obtained via a multi-technique, panchromatic approach. Unfortunately, observational and computational limitations have so far limited the number of objects for which such an analysis could be conducted to a handful. The most notable examples are the studies of the ``Butterfly Star'' \\citep{wolf03}, IM\\,Lup \\citep{pinte08} and IRAS\\,04158+2805 \\citep{glauser08}. In the former two cases, these studies have unambiguously shown that the dust population is stratified, possibly indicating that dust settling is already occurring. Increasing the number of disks studied in such detail is necessary to disentangle individual peculiarities from genuine trends associated with disk evolution. {\\hv} is a triple system located in the Taurus star-forming region. It consists of a 550\\,AU wide pair whose optically brightest component is itself a tight (10\\,AU) visual binary \\citep{simon96}. Spectroscopic and photometric measurements revealed that this subsystem does not currently experience accretion nor does it show infrared excess. They further establish an age of about 2\\,Myr for the system \\citep{white01}. The third component of the system, {\\hvc}, is much fainter yet bluer than {\\hvab}. While \\citet{magazzu94} first thought that this source was an Herbig-Haro object, \\citet{woitas98} later proposed that {\\hvc} is a normal M0 T\\,Tauri star surrounded by an opaque edge-on disk similar to that found in HH\\,30 by \\citet{burrows96}. Subsequent high resolution imaging confirmed this hypothesis \\citep{monin00}. \\citet{stapelfeldt03} produced the first model of high resolution 0.8 and 2.2\\mic\\ scattered light images of {\\hvc}, finding that dust properties similar to those of interstellar dust grains can account for these images. Their 0.8\\mic\\ image also revealed the presence of a roughly spherical envelope, producing a symmetric halo that is more extended than the disk itself, that is likely the remnant of the core from which the system formed. As a consequence of their particular viewing geometry, edge-on protoplanetary disks offer a unique opportunity to determine their geometry and dust content. As such, they may be the best candidates to study vertical stratification in protoplanetary disks. They also are comparatively easy targets for high-angular resolution instruments since the contrast requirement is strongly relaxed. On the other hand, they are challenging from the modeling point of view because of the difficulty of accurately solving for radiative transfer including anisotropic scattering in high optical depth regions. Previous modeling efforts of edge-on protoplanetary disks have therefore focused on interpreting one type of observation at a time \\citep{burrows96, stapelfeldt98, stapelfeldt03, cotera01, wood02, watson04}. While these studies proved highly valuable to constrain some of the disk properties, no self-consistent model was used to model all data at once, leaving unexplained contradictions. Our objective in this work is to perform a global analysis of the {\\hvc} disk, combining scattered light images, millimeter interferometric data and the overall SED into a single fit. We present a series of new observations of the system in Section~\\ref{sec:obs} and discuss our empirical results in Section~\\ref{sec:results}. In Section~\\ref{sec:models}, we present radiative transfer models of the {\\hvc} disk and discuss their implications in Section~\\ref{sec:discus}. Section~\\ref{sec:concl} summarizes our results. ", "conclusions": "\\label{sec:concl} We have obtained new 1--5\\mic\\ adaptive optics images of the {\\hv} triple system at the VLT and Keck observatories, including the first 4.8\\mic\\ scattered light image of the edge-on {\\hvc} disk. All of our images reveal a steady lateral asymmetry in the disk that indicates a departure from pure axisymmetry in the disk. This could be related to the known variability of {\\hvc}. We extact precise relative astrometry from our new images and find that {\\hvab}--C constitutes a common proper motion pair. We also resolved the tight binary system {\\hvab} and found a surprisingly slow orbital motion compared to its projected separation, probably due to orbit eccentricity and/or a large deprojected physical separation. We have also obtained the first spatially resolved 1.3\\,mm continuum and $^{12}$CO 2--1 millimeter maps of the {\\hvc} disk with IRAM's PdBI interferometer. The continuum emission is resolved along the same position angle as the scattered light images, as expected from dust thermal emission. The CO map shows evidence of Keplerian rotation about an 0.5--1$M_\\odot$ central star. We also completed the SED of that component with new {\\it Spitzer}/MIPS 24, 70 and 160\\mic\\ observations as well as observations at 350\\mic, 2.8 and 3.3\\,mm at the CSO, PdBI and CARMA facilities. The 1--3\\,mm spectral index of {\\hvc} is relatively flat ($\\alpha_{\\rm mm}=2.5\\pm0.2$), indicative of the presence of millimeter-sized grains in the disk and/or high optical depth even at millimeter wavelengths. To interpret these data along with previously published HST images of {\\hvc}, we have computed a grid radiative transfer models that produced synthetic disk observations. While we can reproduce most observational properties of {\\hvc}, a Bayesian analysis of our model grid reveals that the SED and scattered light images of {\\hvc} provide mutually exclusive constraints. As found previously, the scattered light images are best fit with a relatively small total mass and interstellar-like dust properties. Fitting the SED, however, requires almost two orders of magnitude more mass as well as a grain size distribution that extends to millimeter sizes. The mismatch between the two sets of constraints reveals an intrinsic shortcoming of our parametrized model, which assumes perfectly mixed dust. Indeed, the small maximum dust grain size inferred from fitting the scattered light images is impossible to reconcile with the long-wavelength end of the SED. In turn, this suggests that both grain growth and vertical stratification are present in the {\\hvc} disk. While these phenomena have been suggested for other protoplanetary disks in the past, {\\hvc} is one of a handful of objects that have been studied in sufficient details to eventually provide a complete picture of these processes. Future observations of {\\hvc} can help refine the model proposed here. For instance, high-resolution (0\\farcs2 or better) millimeter mapping of the disk would better resolve it and provide more stringent constraints. In particular, obtaining observations at even longer wavelengths, 7\\,mm or 1.3\\,cm, would probe a regime in which the optical depth through the disk is much smaller and enable a more direct interpretation in terms of disk properties. On the longer run, mapping of the disk with ALMA in the submillimeter regime and in scattered light at 8--10\\mic\\ with the next generation of large ground-based telescopes will help bridge the gap between the scattering and thermal emission regimes at a roughly constant spatial resolution. In particular, very high resolution maps with ALMA may resolve this and other disks along the vertical axis, in order to probe their vertical stratification and, ultimately, to outline an evolutionary sequence." }, "0911/0911.1992_arXiv.txt": { "abstract": "We take advantage of good atmospheric transparency and the availability of high quality instrumentation in the $1\\, \\micron$ near-infrared atmospheric window to present a grid of F, G, K, and M spectral standards observed at high spectral resolution ($R\\approx$25,000). In addition to a spectral atlas, we present a catalog of atomic line absorption features in the $0.95-1.11\\,\\micron$ range. The catalog includes a wide range of line excitation potentials, from 0-13 eV, arising from neutral and singly ionized species, most frequently those of Fe {\\scshape i} and Ti {\\scshape i} at low excitation, Cr\\,{\\scshape i}, Fe {\\scshape i}, and Si {\\scshape i} at moderate excitation, and C {\\scshape i}, S {\\scshape i}, and Si {\\scshape i} having relatively high excitation. The spectra also include several prominent molecular bands from CN and FeH. For the atomic species, we analyze trends in the excitation potential, line depth, and equivalent width across the grid of spectroscopic standards to identify temperature and surface gravity diagnostics near $1\\, \\micron$. We identify the line ratios that appear especially useful for spectral typing as those involving Ti\\,{\\scshape i} and C {\\scshape i} or S {\\scshape i}, which are temperature sensitive in opposite directions, and Sr\\,{\\scshape ii}, which is gravity sensitive at all spectral types. ASCII versions of all spectra are available in the online version of the journal. ", "introduction": "\\label{sec:intro} Since the advent of infrared arrays in the early 1990s, there have been many low resolution spectroscopic atlases published covering the $J$, $H$, or $K$ bands. Previous efforts that included the full O--M spectral classes and I--V luminosity classes made use of moderate resolution Fourier transform spectroscopy techniques \\citep[e.\\/g.\\/,][]{Wall00,Mey98,Wall97}. An additional window of atmospheric transparency between these traditional near-infrared bands and the optical window, dubbed the $Y$ band, was highlighted by \\citet{Hill02}. With existing observational capabilities on instruments at the IRTF, Keck, Las Campanas, and UKIRT telescopes, and with upcoming wide-field survey facilities such as Pan-STARRS and VISTA, the $Y$ band is becoming an increasingly appreciated and scientifically important near-infrared bandpass. Several low resolution ($R\\equiv\\lambda/\\Delta\\lambda\\sim2000$) spectral atlases for M, L, and T dwarfs have been published including $Y$ band data, e.g. \\citet{Leg96}, \\citet{McLe03}, and \\citet{Cush05}. However, a high-resolution spectroscopic atlas covering earlier spectral types and the full range of luminosity classes has not yet appeared. To complement existing moderate resolution catalogs in the $J$, $H$, and $K$ bands, we present here a high-resolution infrared spectral atlas for the $Y$ band. Our data cover the $0.94-1.12\\,\\mu$m region sampled by NIRSPEC \\citep{McLe98} at the W. M. Keck Observatory. We obtained spectra of 20 MK-classified stars ranging in spectral type from F through M and in luminosity class from I--V for use as spectroscopic standards in future work. The data were processed as briefly described in Section \\ref{sec:data}. We present a representative grid of spectroscopic standards, a table of identified absorption features, and the resulting atlas of spectral lines in Section \\ref{sec:atlas}. In Section \\ref{sec:analysis}, we present relevant atomic data and illustrate as well as discuss line sensitivity to stellar effective temperature and surface gravity. ", "conclusions": "\\label{sec:summary} We have presented a spectral atlas for the $Y$ band region of the near infrared portion of the electromagnetic spectrum. Our data set consists of a grid of MK-classified stars for use as spectroscopic standards spanning the F through M spectral classes and I--V luminosity classes observed at R$\\approx$25,000. The associated catalog of identified lines contains numerous spectral absorption features, both atomic and molecular in origin, that span a wide range of excitation potentials. Many of the atomic features show strong line depth and equivalent width trends across the spectral atlas. This atlas and the identified atomic spectral lines are potentially useful for two-dimensional spectral classification, especially line depth ratios involving Ti {\\scshape i} and C {\\scshape i} or S {\\scshape i}, which are temperature sensitive in opposite directions, and Sr {\\scshape ii}, which is gravity sensitive. ASCII versions of all spectra are available to download with the online version of the journal (an example of the form and content of these spectra can be found in Table \\ref{tab:asciiex})." }, "0911/0911.4862_arXiv.txt": { "abstract": "{We observed 51~Peg, the first detected planet-bearing star, in a 55 ks {\\em XMM-Newton} pointing and in 5 ks pointings each with {\\em Chandra} HRC-I and ACIS-S. The star has a very low count rate in the {\\em XMM} observation, but is clearly visible in the {\\em Chandra} images due to the detectors' different sensitivity at low X-ray energies. This allows a temperature estimate for 51~Peg's corona of T$\\lesssim1\\mbox{MK}$; the detected ACIS-S photons can be plausibly explained by emission lines of a very cool plasma near 200~eV. The constantly low X-ray surface flux and the flat-activity profile seen in optical \\ion{Ca}{ii} data suggest that 51~Peg is a Maunder minimum star; an activity enhancement due to a Hot Jupiter, as proposed by recent studies, seems to be absent. The star's X-ray fluxes in different instruments are consistent with the exception of the HRC Imager, which might have a larger effective area below 200~eV than given in the calibration.} ", "introduction": "The star 51~Peg (GJ~882, HD~217014) shot to fame in 1995 when \\cite{mayorqueloz1995} detected an exoplanet in its orbit, the planetary parameters being quite unexpected at that time, because 51~Peg~b is a giant planet, located at only 0.05~AU distance. The star itself is a G5V star 15.4~pc away from the Sun. Its properties are quite similar to the Sun's, since 51~Peg is about 4~Gyr old and its mass, radius and effective temperature are comparable to solar values with $R=1.27R_{\\sun}$ \\citep{bainesmcalister2008}, $M=1.11M_{\\sun}$, $T_{eff}\\approx 5790K$ \\citep{fuhrmannpfeiffer1997}. However, 51~Peg is a metal-rich star, for which the metallicities given in the literature vary over a wide range of $+0.05\\leq [Fe/H]\\leq +0.24$, see for example \\cite{valentifischer2005}. Enhanced metallicities are a common feature of stars with giant planets \\citep{gonzales1997, santosisraelianmayor2001}. The activity profile of 51~Peg turned out to be unspectacular. In the Mount Wilson program \\citep{baliunasdonahuesoon1995}, which monitors the \\ion{Ca}{ii} H and K line fluxes of main sequence stars, the star shows a very low and nearly flat chromospheric activity level from 1977 until 1989 and a slight drop in 1990 and 1991. In the Lowell Observatory program \\citep{halllockwoodskiff2007}, it also shows low activity and little variability in \\ion{Ca}{} fluxes since the beginning of the program in 1994. The star was also observed in a 12.5~ks {\\em ROSAT} PSPC pointing in 1992 and detected as a weak X-ray source. The coronal activity of 51~Peg is of interest not only because the star is similar to the Sun, but also with regard to recent studies \\citep{kashyapdrakesaar2008}, which claim stars with close-in giant planets to be more X-ray active than stars with far-out ones. ", "conclusions": "We have detected X-ray emission from 51~Peg in a 55~ks observation with {\\em XMM-Newton} and 5~ks observations with {\\em Chandra} ACIS-S and HRC-I each. The detection of 51~Peg with a low count rate in the {\\em XMM} pointing and the clear source signal in the much shorter {\\em Chandra} observations can be explained by the different effective response of the detectors at low energies and 51~Peg having an extremely cool corona. Our main results are summarized as follows: \\begin{enumerate} \\item 51~Peg shows weak emission in the \\ion{O}{vii} triplet and emission around 200~eV which can be explained most likely by cool silicon emission lines. \\item A coronal temperature of $\\lesssim 1$~MK is consistent with the detected hardness ratios in different energy bands in both the {\\em XMM} and the {\\em Chandra} pointing as well as in the {\\em ROSAT} observation carried out 16~years earlier. \\item The {\\em Chandra} HRC-I count rate is higher than can be explained by differences in the effective areas of HRC and ACIS-S; HRC's effective area might be larger at low energies than given in the calibration so far. \\item The constant and very low surface X-ray flux level together with a flat-activity behavior in chromospheric \\ion{Ca}{ii}~H and K line fluxes suggests 51~Peg to be a Hot Jupiter-bearing Maunder minimum candidate. \\end{enumerate}" }, "0911/0911.1489_arXiv.txt": { "abstract": "We present X-ray analysis of two Wolf-Rayet (WR) binaries: V444 Cyg and CD Cru using the data from observations with XMM-Newton. The X-ray light curves show the phase locked variability in both binaries, where the flux increased by a factor of $\\sim 2$ in the case of V444 Cyg and $\\sim 1.5$ in the case of CD Cru from minimum to maximum. The maximum luminosities in the 0.3--7.5 keV energy band were found to be $5.8\\times10^{32}$ and $2.8\\times10^{32}$ erg s$^{-1}$ for V444 Cyg and CD Cru, respectively. X-ray spectra of these stars confirmed large extinction and revealed hot plasma with prominent emission line features of highly ionized Ne, Mg, Si, S, Ar, Ca and Fe, and are found to be consistent with a two-temperature plasma model. The cooler plasma at a temperature of $\\sim$ 0.6 keV was found to be constant at all phases of both binaries, and could be due to a distribution of small-scale shocks in radiation-driven outflows. The hot components in these binaries were found to be phase dependent. They varied from 1.85 to 9.61 keV for V444 Cyg and from 1.63 to 4.27 keV for CD Cru. The absorption of the hard component varied with orbital phase and found to be maximum during primary eclipse of V444 Cyg. The high plasma temperature and variability with orbital phase suggest that the hard-component emission is caused by a colliding wind shock between the binary components. ", "introduction": "\\label{sec:reduc} The log of the X-ray observations is given in Table~\\ref{tab:xray_observations}. We made use of the archival data obtained with the XMM-Newton observatory which consists of three co-aligned X-ray telescopes observed simultaneously and covered 30$^\\prime$ $\\times$ 30$^\\prime$ region of the sky. The X-ray photons were recorded with the European Photon Imaging Camera (EPIC), which forms images on three CCD-based detectors: the PN (Str$\\ddot{u}$der et al. 2001), and the twin MOS1 and MOS2 (Turner et al. 2001) with an angular resolution of 6$^{\\prime\\prime}$ (FWHM). During the observations, all the three EPIC detectors were active in full frame mode together with the Thick filter. For V444 Cyg, six separate observations were taken with exposure time ranging from 10 to 20 ks. These observations are spread over the half cycle of binary system. However, for the star CD Cru, three separate observations were taken covering almost full binary cycle. \\subsection{EPIC data reduction} \\label{sec:src_reduc} We reduced the X-ray data using standard XMM-Newton Science Analysis System software (SAS version 7.0.0) with updated calibration files (Ehle et al. 2004). Event files for the MOS and the PN detectors were generated using the tasks {\\sc emchain} and {\\sc epchain}, respectively. These tasks allow calibration of the energy and the astrometry of the events registered in each CCD chip and to combine them in a single data file. We restricted our analysis to the energy band 0.3--7.5 keV, as the data below 0.3 keV are mostly unrelated to bonafide X-rays, and above 7.5 keV is mostly dominated by the background counts. Event list files were extracted using the SAS task {\\sc evselect}. Data from the three cameras were individually screened for the time intervals with high background when the total count rates (for single events of energy above 10 keV) in the instruments exceed 0.35 and 1.0 $\\rm{counts~s^{-1}}$ for the MOS and PN detectors, respectively. The observations with observation ID 0206240401 for the source V444 Cyg were heavily affected by the high background events, and the data in MOS detectors were not useful. Only half of the observation time ($\\sim$5 ks) of PN was found useful. Light curves and spectra were extracted using a circular region with the source as the center in the energy range 0.3--7.5 keV of the EPIC detectors. The X-ray sources in the cluster were often found to be largely contaminated due to the emission from the neighboring sources. For this reason, the radii of extraction regions were varied between 20$^{\\prime\\prime}$ and 30$^{\\prime\\prime}$ depending upon the position of the sources in the detector and their angular separation between the neighboring X-ray sources. The background has been estimated from a number of empty regions close to the X-ray source in the same CCD of the detector. X-ray spectra of the sources were generated using SAS task {\\sc especget}, which also computes the photon redistribution as well as the ancillary matrix. For each source, the background spectrum was obtained from regions devoid of any sources chosen according to the source location. Finally, the spectra were re-binned to have at least 20 counts per spectral bin for both the sources in all the observations. \\subsection{RGS data reduction} \\label{sec:rgs} The Reflection Grating Spectrometers (RGS) are mounted on two of the three XMM-Newton X-ray telescopes and were operated in spectroscopy mode during the observations. We followed the standard procedure as outlined in the XMM-Newton handbook to generate the RGS spectra of the sources. The raw data were processed with the task {\\sc rgsproc} at the position of the sources. The other bright sources, found in the same field observed by RGS, have been excluded from background estimation. In addition, we have filtered the event list for high background level events. For star V444 Cyg, the data obtained in observations with observation ID 0206240401 were contaminated with high background events in RGS detector, therefore the data was not useful. For CD Cru, the two observations ID 0109480101 and 0109480201 were not useful as they are heavily contaminated by a brighter source HD 110432 and affected by high background episodes, respectively. \\section {Analysis and Results}\\label{sec:results} \\subsection{X-ray light curves}\\label{sec:WR_lt} \\begin{figure*} \\centering \\subfigure[]{\\includegraphics[width=3.0in]{V444Cyg_PN_JD_total.eps}} \\subfigure[]{\\includegraphics[width=3.0in]{CDCru_JD_PN_total.eps}} \\caption{X-ray light curves of the WR-binaries (a) V444 Cyg and (b) CD Cru in the 0.3--7.5 keV energy band.} \\label{fig:xlc} \\end{figure*} \\begin{figure*} \\centering \\subfigure[]{\\includegraphics[width=3.0in]{V444Cyg_PN_final.eps}} \\subfigure[]{\\includegraphics[width=3.0in]{V444Cyg_MOS_final.eps}} \\caption{(a) PN and (b) MOS light curves at three energy bands: the total (0.3--7.5 keV), the soft (0.3--2.0 keV) and the hard (2.0--7.5 keV), and hardness ratio HR curve as a function of orbital phase, where HR = hard/soft of V444 Cyg.} \\label{fig:V444Cygphaselc} \\end{figure*} \\begin{figure*} \\centering \\subfigure[]{\\includegraphics[width=3.0in]{CDCru_PN_final.eps}} \\subfigure[]{\\includegraphics[width=3.0in]{CDCru_MOS_final.eps}} \\caption{Similar to Fig. \\ref{fig:V444Cygphaselc} but for CD Cru.} \\label{fig:CDCruphaselc} \\end{figure*} The background subtracted X-ray light curves of the WR-binaries V444 Cyg and CD Cru, as observed with PN detector are shown in Fig. \\ref{fig:xlc}(a) and \\ref{fig:xlc}(b), respectively. The light curves are in the 0.3--7.5 keV (total) energy band show the variability which suggests colliding wind shocks. Further, we performed the $\\rm{\\chi^2}$ test to measure the significance of the deviations from the mean count rate in order to quantify the constancy of the data over the time-scale of observations. We found the variability in the light curves with a confidence level of greater than 99.999\\% for both binaries. In order to investigate the variability in the different energy bands, the light curve of these binaries obtained with the MOS and PN data are generated into two energy bands namely the soft (0.3--2.0 keV) and the hard (2.0--7.5 keV). The hardness ratio (HR) is defined by the ratio of hard to soft band count rates. The total, hard and soft band intensity curves, and the HR curve as a function of the orbital phase are shown in the subpanels running from top to bottom in Fig. \\ref{fig:V444Cygphaselc}a (PN) and \\ref{fig:V444Cygphaselc}b (MOS) for the WR binary V444 Cyg. Similar plots of intensity and HR for the star CD Cru are shown in Fig. \\ref{fig:CDCruphaselc}. The phases of the observations are reckoned using the ephemeris HJD = 2441164.337+4.213E for V444 Cyg (Underhill, Grieve \\& Louth 1990) and HJD = 2443918.4000+6.2399E for CD Cru (Moffat et al. 1990, Niemela, Massey \\& Conti 1980). Here, the phase 0.0 indicates the primary eclipse and the phase 0.5 indicates the secondary eclipse. The light curves in the individual bands show the phase locked variability for both binaries. For V444 Cyg, the count rates in total energy band were minimum at phase 0.0 and maximum at the phase 0.45. After the phase 0.45, the count rates were decreased upto the phase 0.5. The count rates are increased by a factor of $\\sim 2$ being minimum at phase 0.0\\ in the total energy band. In the soft band, the count rates were minimum at phase 0.0 and maximum during the phase 0.45-0.5. However, in the hard band light curve two minima at phase 0.0 and 0.5 were seen clearly. During the phases from 0.13 to 0.45 (i.e. outside the eclipse) the count rates in the hard band were constant. The HR curve can reveal the information about the spectral variations. The HR curve shows that the emitted X-ray is harder just after phase 0.0 being maximum at phase 0.13, and afterward decreased till the phase 0.5. In the case of CD Cru, the nature of the variability was found to be similar to that seen in V444 Cyg. In soft band, the count rates were found to be constant during the phases 0.47 and 0.78 (i.e. outside the eclipse). However, the count rates are decreased rapidly after the phase 0.47 \\ in the hard band. A small variation was also observed in the HR curve of the CD Cru. It was maximum at the phase 0.0 and minimum during the phase 0.47. \\subsection{X-ray spectra and spectral fits }\\label{sec:WR_spt} \\begin{figure*} \\centerline{\\hbox{\\hspace{0.5in} \\includegraphics[width=60mm, angle=-90]{V444Cyg_obs301_2Tvapec_new_formate.eps} \\hspace{0.15in} \\includegraphics[width=60mm, angle=-90]{V444Cyg_obs201_2Tvapec_new_formate.eps} }} \\vspace{0.15in} \\centerline{\\hbox{\\hspace{0.5in} \\includegraphics[width=60mm, angle=-90]{V444Cyg_obs701_2Tvapec_new_formate.eps} \\hspace{0.15in} \\includegraphics[width=60mm, angle=-90]{V444Cyg_obs501_2Tvapec_new_formate.eps} }} \\vspace{0.15in} \\centerline{\\hbox{\\hspace{0.5in} \\includegraphics[width=60mm, angle=-90]{V444Cyg_obs801_2Tvapec_new_formate.eps} \\hspace{0.15in} \\includegraphics[width=60mm, angle=-90]{V444Cyg_obs401_2Tvapec_new_formate.eps} }} \\vspace{3pt} \\caption{X-ray spectra of MOS and PN data with the best fit 2T {\\sc vapec} model in upper subpanels of each graph for V444 Cyg. The $\\chi^2$ distribution in terms of ratio are given in lower subpanels of each graph. The last three digits of observation ID and the corresponding phases of the observations are given in each graph. } \\label{fig:spt_V444Cyg} \\end{figure*} \\begin{figure*} \\centerline{\\hbox{\\hspace{0.5in} \\includegraphics[width=60mm, angle=-90]{CDCRu_obs401_2Tvapec_new_formate.eps} \\hspace{0.15in} \\includegraphics[width=60mm, angle=-90]{CDCru_obs201_2Tvapec_new_formate.eps} }} \\vspace{0.15in} \\centerline{\\hbox{\\hspace{0.5in} \\includegraphics[width=60mm, angle=-90]{CD_Cru_obs101_2Tvapec_new_formate.eps} }} \\vspace{3pt} \\caption{Same as Figure~\\ref{fig:spt_V444Cyg} but for CD Cru. } \\label{fig:spt_CDCru} \\end{figure*} The EPIC spectra of WR binaries in different phases are shown in Figs.~\\ref{fig:spt_V444Cyg} and \\ref{fig:spt_CDCru} for V444 Cyg and CD Cru, respectively. Below 1 keV, the spectra were found to be affected by the high extinction as previously observed with ASCA for V444 Cyg (Maeda et al. 1999). Strong emission lines are seen in the MOS and PN spectra of both WR binaries. The most prominent lines found in the spectra along with their laboratory energies are the following: Fe XVII (0.8 keV), Ne X (1.02 keV), Mg XII (1.47 keV), Si XIII (1.853 keV), S XV (2.45 keV), Ar XVII (3.12 keV), Ca XIX+XX (3.9 keV) and Fe XXV (6.63 keV). In order to trace the spectral parameters at different binary phases, we performed spectral analysis of each data set using simultaneous fitting of EPIC data by two models, (a) plane-parallel shock model ({\\sc vpshock}; Borkowski, Lyerly \\& Reynolds 2001), and (b) models of Astrophysical Plasma Emission Code ({\\sc apec}; Smith et al. 2001), as implemented in the XSPEC version 12.3.0. A $\\chi^2$ -- minimization gave the best fitted model to the data. The presence of interstellar material along the line-of-sight and the local circumstellar material around the stars can modify the X-ray emission from massive stars. We corrected for the local absorption in the line-of-sight to the source using the photoelectric absorption cross sections according to Baluci$\\acute{n}$ska-Church \\& McCammon (1992) and modeled as {\\sc phabs} (photoelectric absorption screens) with two absorption components, $\\rm{N_{H}^{ISM}}$ and $\\rm{N_{H}^{local}}$. The $\\rm{N_{H}^{ISM}}$ was estimated using the relation, $\\rm {N_H}$ $\\rm{= 5.0\\times}10^{21}\\times{E(B-V)~cm^{-2}}$ (Vuong et al. 2003), where $\\rm{E(B-V)=A_V/3.1}$, assuming a normal interstellar reddening law towards the direction of the cluster. We used the values of $\\rm{A_V}$ nearly 2.48 mag and 3.56 mag derived for the cluster Berkeley 86 (Massey, Johnson \\& Degioia-Eastwood 1995) and Hogg 15 (Sagar, Munari \\& Boer 2001), respectively. The estimated values of $\\rm{N_{H}^{ISM}}$ towards V444 Cyg and CD Cru are found to be $\\rm{4.0\\times10^{21}~cm^{-2}}$ and $\\rm{6.0\\times10^{21}~cm^{-2}}$, respectively. The $\\rm{N_{H}^{local}}$ was estimated by making a fit to the observed spectra by varying the local environment for the soft ($\\rm{kT_1}$) and the hard ($\\rm{kT_2}$) energy components in terms of $\\rm{N_H^1}$ and $\\rm{N_H^2}$, respectively. Because the WN stars are at evolved stages, the abundances of He, C, N, O, Ne, Mg, Si, S, Ar, Ca, and Fe were allowed to vary during the fitting procedure to account for the observed line emission. The solar abundances were adopted from Lodders (2003). First, we fitted {\\sc vpshock} plasma model to derive their spectral features. The constant temperature {\\sc vpshock} plasma model was considered without incorporating mass-loss and orbital parameters of WR stars. However, the model does account for non-equilibrium ionization effects and assumes an equal electron and ion temperature. The best fit {\\sc vpshock} models to the data for the WR stars are given in Table~\\ref{tab:V444Cyg_fit} and Table~\\ref{tab:CDCru_fit} for V444 Cyg and CD Cru, respectively. X-ray emitting plasma may not be isothermal and the observed X-ray spectrum may be a superposition of a cool stellar component and a hot colliding wind shock plasma, as a number of emission lines in the spectra can be formed over a range of temperatures. The cooler component is believed to arise from the instabilities in radiation-driven outflows. However, using the X-ray imaging data of XMM-Newton it is not possible to resolve the X-ray emission from colliding wind region and individual stars separately. Therefore, secondly, we fitted two temperature (2T) plasma model \"{\\sc vapec}\" to characterize such components. The form of 2T plasma model was {\\sc phabs(phabs*vapec+phabs*vapec)}. In terms of $\\chi^2$ the 2T plasma model provides better goodness-of-fit for both the sources over the {\\sc vpshock} model. The results for the 2T plasma model along with the data are displayed in Fig.~\\ref{fig:spt_V444Cyg} and Fig.~\\ref{fig:spt_CDCru} for V444 Cyg and CD Cru, respectively. The best-fit parameters are given in Table~\\ref{tab:V444Cyg_fit} and Table~\\ref{tab:CDCru_fit} for V444 Cyg and CD Cru, respectively. \\input{bestfit_WR_V444Cyg.tab} \\input{bestfit_WR_CDCru.tab} \\clearpage \\begin{figure} \\centering \\subfigure[]{\\includegraphics[width=3.0in]{analysis_V444Cyg.eps}} \\subfigure[]{\\includegraphics[width=3.0in]{analysis_CDCru.eps}} \\caption{Variation of $\\rm{L_X^S}$, $\\rm{L_X^H}$, $\\rm{N_H^1}$, $\\rm{N_H^2}$, $\\rm{kT_1}$ and $\\rm{kT_2}$ as a function of orbital phase of the the stars (a) V444 Cyg and (b) CD Cru.} \\label{fig:ana_phase} \\end{figure} \\subsection {Evolution of spectral parameters} \\label{sec:spt_evo} The spectral analysis at different phases of the WR binaries provides the dependence of the best fit values of parameters on orbital phase. The variation of the soft ($\\rm{L_X^S}$) and hard ($\\rm{L_X^H}$) band X-ray luminosities, column densities corresponding to the cool ($\\rm{N_H^1}$) and hot ($\\rm{N_H^2}$) temperature components, and cool(kT$_1$) and hot (kT$_2$) temperatures as a function of orbital phases of V444 Cyg and CD Cru are shown in Fig. \\ref{fig:ana_phase}(a) and \\ref{fig:ana_phase}(b), respectively. In the case of V444 Cyg, the $\\rm{L_X^H}$ was found to be minimum during the primary and secondary eclipses and maximum outside the eclipse. It was constant during the phase 0.13 to 0.47. However, the $\\rm{L_X^S}$ was found to be minimum during the primary eclipse only. At the phase of 0.5 the $\\rm{L_X^S}$ was found to be maximum. Both $\\rm{L_X^H}$ and $\\rm{L_X^S}$ were found to be minimum during the phase 0.0 and maximum at phase 0.5 in the WR binary CD Cru. The cool component ($\\rm{kT_1}$) was found to be constant with a mean value of 0.61$\\pm$0.05 keV and 0.57$\\pm$0.10 keV for V444 Cyg and CD Cru, respectively. The hard energy component ($\\rm{kT_2}$) was varied from a minimum value of 1.88 keV to a maximum value of 9.61 keV for V444 Cyg. It was maximum at primary eclipse and at the phase of 0.29, and minimum during the secondary eclipse and at the phase of 0.13. For CD Cru, $\\rm{kT_2}$ varies from a minimum value 1.63 keV during outside the eclipse to a maximum value of 4.27 keV at primary eclipse. For the star V444 Cyg, $\\rm{N_H^1}$ was maximum at phase 0.29, otherwise it was constant at all phases. However, $\\rm{N_H^1}$ was found to be constant throughout the orbital phase with a mean value of 1.14$\\pm$0.11 $\\times \\rm{10^{22}~cm^{-2}}$ for CD Cru. For V444 Cyg, $\\rm{N_H^2}$ was increased from phase 0.0 to its maximum value at phase 0.13. Afterward it was decreased and became almost constant from phase 0.29 to phase 0.47. For CD Cru, $\\rm{N_H^2}$ was minimum at phase 0.0 and maximum at phase 0.78. \\begin{figure*} \\centering \\includegraphics[width=6.0in, height=5.0in]{RGS_all_500b.eps} \\caption{RGS spectra of the WR binaries at different binary phase for the stars V444 Cyg (upper panel) and CD Cru (lower panel).} \\label{fig:rgs} \\end{figure*} \\subsection{RGS spectra} \\label{sec:an_rgs} We combined the first and second order spectra of detector RGS1 and RGS2 to inspect the main spectral lines. We obtained the RGS fluxed spectra using task {\\sc rgsfluxer} and is shown in Fig.~\\ref{fig:rgs} for different phases for both the binaries. We identified the prominent emission lines for V444 Cyg at wavelengths Mg XII (6.50 \\AA; 6.74 \\AA), Mg XI (9.23 \\AA), Fe XVIII (11.40 \\AA) and Ca XVI (22.61 \\AA, 21.43 \\AA, 22.30 \\AA). Individual lines show intensity variations from spectrum to spectrum. Unfortunately, our observations are at the limit of detectability with the data available with poor count statistics. Therefore, the signal-to-noise ratio of individual lines are not sufficient to perform a detailed quantitative analysis of the RGS line spectrum for any of the massive stars. ", "conclusions": "\\label{sec:discuss} We have analyzed the X-ray emission properties of two WR binaries with strong stellar winds. We discuss below the implications of the X-ray results. \\subsection{Phase-locked X-ray variability}\\label{sec:discuss_lt} The X-ray temporal and spectral analyses of WR binaries V444 Cyg and CD Cru showed a phase-locked phenomenon. The phase resolved X-ray spectroscopic observations show that the soft energy component and the corresponding $\\rm{N_H^1}$ were also found to be nearly constant throughout the binary phases of both stars. The $\\rm{L_X^S}$ was found to be minimum during the primary eclipse only. These results lead to the radiative wind shock origin of soft energy component. In the case of CD Cru, the modulation of $\\rm{L_X^S}$ well matches with the optical light curve reported by Moffat et al. (1990). Lamontagne et al. (1996) have reported a similar shape of optical photometric light curves for 13 WR binaries including CD Cru and they said that it could be due to the \"atmospheric eclipse\". The X-ray flux in the hard energy band was found to be minimum at both primary and secondary eclipse for both stars. This implies that the hard energy component could be due to the presence of wind-wind collision zone which is located somewhere in between the O-type and WR star, and therefore, minimum when either of the star in the binary system is eclipsed. A similar behavior of X-ray light curves was seen in the EINSTEIN, ROSAT and ASCA observations of V444 Cyg (Moffat et al. 1982; Corcoran et al. 1996 ; Maeda et al. 1999). The hot temperature ($\\rm{kT_2}$) and the corresponding $\\rm{N_H^2}$ were also varied with orbital phases of both binaries. The eccentricity of both binary systems is almost zero, therefore, the variation in the temperature corresponding to the hard energy component could be the result of variation in $\\rm{N_H^2}$, i.e., varying optical depth (see \\S\\ref{sec:intro}). This further supports the wind-wind collision phenomenon in these binaries. \\subsection{X-ray Temperatures of Plasma}\\label{sec:discuss_spt} The X-ray spectra of WR binaries are well defined by two temperature plasmas. The values of $\\rm{log(L_X/L_{bol})}$ in the total energy band are found to be -6.6 and -7.2 in for V444 Cyg and CD Cru, respectively, which are similar to those for other WR (WN) stars observed from {\\sc XMM-Newton} and {\\sc CHANDRA} (G$\\ddot{u}$del \\& Naz$\\acute{e}$ 2009). The temperature of the cool component of WR binaries are comparable to other WR binaries e.g., 0.56--0.67 keV for WR1 (Ignace, Oskinova \\& Brown 2003) and 0.7--0.8 keV for WR 147 (Skinner et al. 2007). However, the temperatures corresponding to hot energy component o f these WR binaries are slightly more than some of the similar binaries. The possible mechanisms for generation of X-rays are discussed below. \\subsubsection{ Instabilities driven radiative wind shocks } The wind-shock model predicts the intrinsic instability of the line driving mechanism. The standard model estimates the shock velocities by the relation $\\rm{kT_{sh}=1.95 \\mu v^2_{shock}}$ (Lucy 1982; Luo, McCray \\& Maclow 1990). The temperatures of the cool energy components are found to be almost similar ($\\sim$ 0.6 keV) for both the binary systems (see Table~\\ref{tab:V444Cyg_fit} and Table~\\ref{tab:CDCru_fit}). Adopting the mean particle weight $\\rm{\\mu\\approx}$1.16 (Skinner et al. 2007) for WN stars and $\\rm{\\mu\\approx}$0.62 for O-type stars (Cassinelli et al. 2008), we derived the \"average\" value of shock velocities $\\approx$ 515 $\\rm{km~s^{-1}}$ for WN stars and $\\approx$ 704 $\\rm{km~s^{-1}}$ for O-type stars corresponding to the cool component. These values are about a factor of 2 larger than those predicted by radiative shock model of Lucy (1982). However, the advance version of the wind shock model by Owocki et al. (1988) predicts X-ray emission up to 1 keV. Moreover, the temperatures of the cool energy component and $\\rm{N_H^1}$ are not varying with orbital phase. Therefore, it appears that the cool energy component from V444 Cyg and CD Cru could be generated by either of the binary components and may be explained by instabilities in radiation-driven wind shocks. \\subsubsection{Magnetically Confined Wind Shock} Babel \\& Montmerle (1997) have suggested that the presence of magnetic fields confine the wind, which may be an important ingredient for the production of X-ray emission (Babel \\& Montmerle 1997). The degree of confinement of wind by the magnetic field is derived in terms of a confinement parameter $\\rm{ \\Gamma=B^{2}_{0} R^{2}_{\\ast}/\\dot{M}v_{\\infty} }$, where $\\rm{B_0}$ is the surface equatorial magnetic field strength (ud-Doula \\& Owocki 2002). For a confined wind model $\\rm{\\Gamma>>1}$. Using the values of stellar radius ($\\rm{R_{\\ast}}$) and stellar mass loss rate ($\\rm{\\dot{M}v}$) as given in Table~\\ref{tab:par_str} for V444 Cyg binary components, the minimum magnetic field required to confine the wind is derived to be 0.14 kG for the primary and 1.25 kG for the secondary. Similarly, for CD Cru, we derived minimum magnetic field of 0.16 kG and 1.96 kG for the primary and the secondary components, respectively. Based on the formula given by Babel \\& Montmerle (1997) and using above estimated values of B, the X-ray luminosity are estimated to be $10^{34.66}$~\\egs~and $10^{35.74}$ ~\\egs~for primary and the secondary components of V444 Cyg, and $10^{35.06}$~\\egs~and $10^{36.8}$ ~\\egs~for primary and the secondary components of CD Cru. Such a high X-ray luminosity has not been observed from any of the WR binaries. However, smaller fields can exist and the confinement may be limited to just above the magnetic equatorial plane. \\subsubsection{Colliding wind shock model} \\input{model.tab} Assuming that shock velocities have reached to the terminal velocities, the standard model predicts a maximum temperature for the shocked region from winds of massive stars by the relation $\\rm{kT^{max}_{sh}=1.95 \\mu v^2_{\\infty}}$ (Luo, McCray \\& Maclow 1990). The nominal wind parameters of massive stars (see Table~\\ref{tab:par_str}) and the mean particle weight $\\rm{\\mu}$ for the WN stars and for the O-type stars give the maximum temperature generated by the individual components of binaries. The derived values are 7.80 keV + 7.21 keV for V444 Cyg (O6+WN5) and 10.88 keV + 13.69 keV for CD Cru (O5+WN6). The maximum observed temperature at phase 0.29 corresponding to hot component is found to be similar to that of maximum possible values in the case of V444 Cyg. However, in CD Cru maximum observed temperature is lesser than that of maximum possible values. The phase locked variability of the hard energy component shows that the observed hot X-ray temperatures in massive binaries could originate from the collision of stellar winds. The distances from the stars where these winds meet are derived using the relations (De Becker 2007; Stevens, Blondin \\& Pollock 1992), \\begin{equation} \\rm{ r_{OB}=\\frac {1} {1+\\eta^{1/2}} D} \\end{equation} \\noindent where $\\rm{r_{OB}}$ and D are the distance to the collision zone from the primary and the separation between the binary components, respectively. The wind momentum ratio $\\rm{\\eta}$ is expressed as, \\begin{equation} \\rm{\\eta=\\frac{\\dot{M_2}v_{\\infty,2}}{\\dot{M_1}v_{\\infty,1}}} \\end{equation} \\noindent where $\\rm{\\dot{M_1}}$ and $\\rm{\\dot{M_2}}$ are mass loss rates of primary and secondary components, respectively, and $\\rm{v_{\\infty,1}}$ and $\\rm{v_{\\infty,2}}$ are terminal velocities of primary and secondary components, respectively. Using the above relations, we derived wind momentum ratios of 7.03 and 24.60 for the stars V444 Cyg and CD Cru , respectively. Distances to the collision zone from the primary are estimated to be 0.27 and 0.17 times of their corresponding binary separation (D) for V444 Cyg and CD Cru, respectively. Therefore, the collision zone exists to be very close to the primary O-type star in case of both the binaries. Further, the collision wind zone will produce a bow shock around the O-type star , the shock volume and the emission measure are dominated by the WR winds, therefore, X-ray emitting plasma also shows non-solar abundances. The gas in a colliding wind region could be either adiabatic or radiative and depends upon the value of cooling parameter ($\\chi$), which is defined as the ratio of the cooling time ($\\rm{t_{cool}}$) of the shocked gas to the escape time ($\\rm{t_{esc}}$) from the intershock region (Stevens, Blondin \\& Pollock 1992). \\begin{equation} \\rm {\\chi = \\frac{t_{cool}}{t_{esc}} = \\frac{v^4_3 d_7} {\\dot{M}_{-7}}} \\end{equation} \\noindent where $\\rm{v_{3}}$ is the pre-shock velocity in the unit of 10$^3$ $\\rm{km~s^{-1}}$, $\\rm{d_{7}}$ is the distance from stellar center to the shock in the unit of 10$^7$ km and $\\rm{\\dot{M}_{-7}}$ is the mass loss rate in the unit of $\\rm{10^{-7}M_{\\odot} yr^{-1}}$. Using the parameters given in Table~\\ref{tab:par_str} and the values from above estimation, we derived cooling parameters $\\rm{\\chi_1}$ and $\\rm{\\chi_2}$ for primary and secondary components, respectively. The estimated values of $\\rm{\\chi_1}$ and $\\rm{\\chi_2}$ are given Table~\\ref{tab:model}. The winds from WR stars are found to be clearly radiative ($\\rm{\\chi<1}$) for both the binaries. The intrinsic luminosity can be estimated for each component of the binary using the following relation (Pittard \\& Stevens 2002) : \\begin{equation} \\rm {L_X = 0.5~\\Xi~\\dot{M}~v^2 } \\end{equation} \\noindent where $\\rm{\\Xi}$ accounts for a geometrical and inefficiency factors and v is the wind speed at contact surface (i.e. pre-shock velocities derived from observed temperatures). The values of $\\rm{\\Xi_{1}}$ (0.403 for V444 Cyg ; 0.564 for CD Cru) and $\\rm{\\Xi_{2}}$ (0.033 for V444 Cyg ; 0.0042 for CD Cru) for primary and secondary stars, respectively, are taken from Pittard \\& Stevens (2002). The estimated values of intrinsic luminosities of each component and the total intrinsic luminosities of binary systems are given in Table~\\ref{tab:model}. These values are nearly 3 orders of magnitude higher than those of the observed values. A similar results have been found for other WR binaries, e.g., HD 159176 (De Becker et al. 2004); WR 147 (Skinner et al. 2007). De Becker et al. (2004) suggested that the disagreement between observed and theoretical predictions could possibly be explained by (a) the kinetic power of the collision should be considered as an upper limit on the X-ray luminosity, (b) higher value of the parameter $\\rm{\\eta}$ should be considered, (c) diffusive mixing between hot and cool material is likely to exist due to the instability of the shock front, and (d) orbital effects should also be included to study such systems. The observed and theoretically predicted values of X-ray luminosities are shown in Fig.~\\ref{fig:model}. This model also predicts the phase-locked variations which can be seen in Fig.~\\ref{fig:ana_phase}. Therefore, it appears that the hard energy component is most likely associated with wind collision zone. \\begin{figure} \\centering \\includegraphics[width=3.5in, height=3.5in]{model_obs_hard.eps} \\caption{Observed vs Model X-ray luminosity.} \\label{fig:model} \\end{figure} \\subsubsection{ Non-thermal emission} There is no strong justification for invoking non-thermal origin of X-ray emission since the observed X-ray emission at higher energies can be fitted satisfactorily with a thermal model (see Fig.~\\ref{fig:spt_V444Cyg} and Fig.~\\ref{fig:spt_CDCru}). However, the present data is limited up to 10 keV and it is difficult to eliminate the contribution of non-thermal components from the spectra which is dominated by the presence of the strong thermal emission, and the non-thermal emission is overwhelmed by the thermal emission (De Becker 2007). Therefore, we can state that the spectra of massive stars in the present sample are fairly well explained by two-temperature plasma models." }, "0911/0911.4924_arXiv.txt": { "abstract": "{}{We investigate the dependence of $\\gamma$-ray brightness of blazars on intrinsic properties of their parsec-scale radio jets and the implication for relativistic beaming.}{By combining apparent jet speeds derived from high-resolution VLBA images from the MOJAVE program with millimetre-wavelength flux density monitoring data from Mets\\\"ahovi Radio Observatory, we estimate the jet Doppler factors, Lorentz factors, and viewing angles for a sample of 62 blazars. We study the trends in these quantities between the sources which were detected in $\\gamma$-rays by the $Fermi$ Large Area Telescope (LAT) during its first three months of science operations and those which were not detected.}{The LAT-detected blazars have on average higher Doppler factors than non-LAT-detected blazars, as has been implied indirectly in several earlier studies. We find statistically significant differences in the viewing angle distributions between $\\gamma$-ray bright and weak sources. Most interestingly, $\\gamma$-ray bright blazars have a distribution of comoving frame viewing angles that is significantly narrower than that of $\\gamma$-ray weak blazars and centred roughly perpendicular to the jet axis. The lack of $\\gamma$-ray bright blazars at large comoving frame viewing angles can be explained by relativistic beaming of $\\gamma$-rays, while the apparent lack of $\\gamma$-ray bright blazars at small comoving frame viewing angles, if confirmed with larger samples, may suggest an intrinsic anisotropy or Lorentz factor dependence of the $\\gamma$-ray emission.}{} ", "introduction": "One of the most important discoveries of the Energetic Gamma-Ray Experiment Telescope (EGRET) on-board the {\\it Compton Gamma-Ray Observatory} in the 1990s was the detection of over 65 active galactic nuclei (AGN) at photon energies above 100\\,MeV \\citep{har99,mat01}. The detected sources were almost exclusively blazars, a class of highly variable AGN comprised of flat spectrum radio quasars and BL Lac objects. The distinctive characteristic of blazars is a relativistic jet oriented close to our line-of-sight. Synchrotron radiation of energetic electrons in the jet dominates the low energy end of the blazar spectral energy distribution. This emission is strongly beamed due to relativistic effects which increase the observed flux density of a stationary jet by a factor of $\\delta^{2-\\alpha}$ and that of distinct ``blobs'' in the jet by a factor of $\\delta^{3-\\alpha}$. Here $\\delta$ is the jet Doppler factor and $\\alpha$ is the spectral index defined as $S_\\nu \\propto \\nu^{+\\alpha}$ \\citep{bla79}. The Doppler factor is defined as $\\delta = [\\Gamma (1-\\beta \\cos \\theta)]^{-1}$, where $\\Gamma = (1-\\beta^2)^{-1/2}$ is the bulk Lorentz factor, $\\beta$ is the jet speed divided by the speed of light, and $\\theta$ is the angle between the jet and our line-of-sight. The requirement that the $\\gamma$-ray-bright sources are transparent to $\\gamma\\gamma$ pair production, together with their small sizes deduced from the fast $\\gamma$-ray variability, strongly suggest that the $\\gamma$-ray emission originates in the jet and is also relativistically beamed, in the same manner as the radio emission \\citep{mon95,mat93}. The many correlations found between the $\\gamma$-ray emission detected by EGRET and the radio/mm-wave properties of blazars further support this scenario \\citep{val95, jor01b, jor01a, lah03, kel04, kov05}. Since the $\\gamma$-ray spectrum is typically steeper than the radio spectrum \\citep{abd09b}, it is possible that the $\\gamma$-ray emission is even more enhanced by Doppler boosting than the radio emission. Although inverse Compton (IC) scattering of soft photons off relativistic electrons in the jet is currently the favoured model for the $\\gamma$-ray emission, there is a substantial controversy within this model about the origin of the target photon field and the location of the emission site. The seed photons could be, for example, synchrotron photons emitted by the same electrons which scatter them later, \\citep[synchrotron self-Compton model, SSC;][]{mar92, blo96} or synchrotron photons emitted by electrons in a different layer of the jet \\citep{ghi05}. The seed photons can also originate in sources external to the jet like the accretion disk, the broad line region clouds or the dust torus \\citep[external Compton model;][]{der92,ghi96,sik94,bla00}. Beside these leptonic models there are also a number of models where $\\gamma$-rays are produced by hadronic processes initiated by relativistic protons co-accelerated with electrons \\citep[e.g.,][]{man93}. The Large Area Telescope (LAT) on-board the {\\it Fermi Gamma-ray Space Telescope} is a successor to EGRET, with much better sensitivity, larger energy range (up to 300 GeV), better angular resolution, and larger field-of-view \\citep{atw09}. In only its first three months of science operations, the LAT detected 205 bright $\\gamma$-ray sources at $>10\\sigma$ level \\citep{abd09a}, 116 of which are associated with AGN at high galactic latitudes ($|b| \\ge 10^\\circ$) \\citep{abd09b}. Here we refer to these 116 sources as ``LAT-detected sources''. Based on the long-term monitoring of the radio jet motions with the Very Long Baseline Array (VLBA), it was recently shown that the LAT-detected quasars have significantly faster apparent jet speeds and higher core brightness temperatures than non-LAT-detected quasars \\citep{lis09b,kov09}. The LAT $\\gamma$-ray photon flux also correlates with the compact radio flux density and the flares in $\\gamma$-rays and radio seem to happen in the VLBI cores within a typical apparent time separation of up to several months \\citep{kov09}. The $\\gamma$-ray bright blazars also have larger-than-average apparent jet opening angles \\citep{pus09}. These findings indicate that $\\gamma$-ray bright blazars are likely more Doppler boosted than $\\gamma$-ray faint ones, but what remains unknown is a possible dependence of {\\it intrinsic} $\\gamma$-ray luminosity on parsec-scale jet properties such as the bulk Lorentz factor or the viewing angle in the comoving frame of the jet \\citep{lis09b}. In this paper we confirm the connection between the $\\gamma$-ray brightness and the Doppler factor of the parsec-scale jet for a sample of 62 blazars. We have also combined measurements of the apparent jet speeds and temporal variability at mm-wavelengths to derive jet Lorentz factor and the viewing angle in the observer's frame and in the frame comoving with the jet for 57 blazars in order to study how the $\\gamma$-ray detection probability depends on these intrinsic jet parameters. \\begin{figure*} \\centering \\includegraphics[width=0.75\\textwidth]{savolainen_v4fg1.eps} \\caption{{\\it Upper left:} Variability Doppler factor ($\\delta_\\mathrm{var}$) distributions for non-LAT-detected (upper sub-panel) and LAT-detected (lower sub-panel) blazars in the Mets\\\"ahovi-MOJAVE sample. Quasars are denoted by shaded bins. {\\it Lower left:} Bulk Lorentz factor ($\\Gamma$) distributions. {\\it Upper right:} Distributions of the angles between the jet and our line-of-sight in the observer's frame ($\\theta$). {\\it Lower right:} Distributions of the viewing angles in the comoving frame of the jet ($\\theta_\\mathrm{src}$). The values were calculated assuming $\\Lambda$CDM cosmology with $H_0 = 71$\\,km\\,s$^{-1}$\\,Mpc$^{-1}$, $\\Omega_\\mathrm{M} = 0.27$, and $\\Omega_\\mathrm{\\Lambda} = 0.73$.} \\label{fig1} \\end{figure*} ", "conclusions": "The variability Doppler factors from \\citet{hov09} were determined over a multi-year monitoring program while the $\\gamma$-ray detections used in our analysis are based on just three-months of LAT monitoring. The sharp distinction seen in the $\\delta_\\mathrm{var}$ values between the LAT-detected and non-LAT-detected sources strongly supports the idea that $\\delta_\\mathrm{var}$ must remain fairly constant with time. Since the LAT-detected and non-LAT-detected sources are treated in the exact same way with respect to the derivation of their $\\delta_\\mathrm{var}$, any temporal variation in the Doppler factor (or in $T_\\mathrm{b,int}$) would thus only serve to destroy the possible correlation and could not create one. Therefore, we consider the result regarding higher $\\delta_\\mathrm{var}$ for the LAT-detected blazars to be very robust. A similar result regarding the high $\\delta_\\mathrm{var}$ of $\\gamma$-ray bright sources was found earlier using the less uniform EGRET data \\citep{lah03}. The simplest and arguably most likely interpretation is that the $\\gamma$-ray bright blazars are indeed systematically more Doppler boosted than the $\\gamma$-ray weak ones. This interpretation is compatible with the LAT-detected quasars having faster apparent jet speeds \\citep{lis09b}, wider apparent jet opening angles \\citep{pus09}, and higher VLBI brightness temperatures \\citep{kov09}. An alternative interpretation would be that, for some reason, the intrinsic brightness temperature is systematically about a factor of three higher in the LAT-detected blazars than in the non-LAT-detected ones. In the latter case our assumption of a constant limiting brightness temperature would lead to an overestimation of $\\delta_\\mathrm{var}$ in the $\\gamma$-ray bright blazars. This alternative explanation would not, however, explain the faster apparent jet speeds or wider apparent jet opening angles of the LAT-detected sources. The observed difference in the comoving-frame viewing angle distributions between the $\\gamma$-ray bright and weak blazars is an unanticipated result. The left panel of Fig.~\\ref{fig2} shows that the lack of LAT-detected blazars at large comoving-frame viewing angles can be explained by low Doppler factors of the sources at large $\\theta_\\mathrm{src}$. The beaming model does not, however, explain the lack of LAT-detected sources at {\\it small} values of $\\theta_\\mathrm{src}$. If this lack is real, it may reflect an intrinsic anisotropy of the $\\gamma$-ray emission in the comoving frame of the jet. This would have wide implications for the theoretical models of the high energy emission from blazars, since almost all of these models rely on the assumption that the $\\gamma$-ray emission is (nearly) isotropic in the rest frame of the relativistically moving sub-volume \\citep[e.g.][]{der95}. Possible sources of anisotropy in the $\\gamma$-ray emission include, for example, anisotropic $\\gamma\\gamma$ absorption or anisotropic seed photon field for inverse Compton scattering. The right hand panel of Fig.~\\ref{fig2} shows that the non-LAT-detected sources at small comoving frame viewing angles have small Lorentz factors, except for \\object{B0804+499}. This may provide an alternative explanation for the apparent lack of $\\gamma$-ray bright sources at small comoving frame viewing angles if the \\textit{intrinsic} $\\gamma$-ray luminosity depends on the bulk Lorentz factor in addition to being relativistically beamed. The fact that \\object{B0804+499} is not a bright $\\gamma$-ray source, despite having a very high Doppler factor and a moderately high Lorentz factor, poses a problem for this explanation, but does not rule it out since the $\\gamma$-ray emission may be intermittent (or there may be something unusual in this particular source)." }, "0911/0911.5113_arXiv.txt": { "abstract": "% The SuperAGILE (SA) instrument is a X-ray detector for Astrophysics measurements, part of the Italian AGILE satellite for X-Ray and Gamma-Ray Astronomy launched at 23/04/2007 from India. SuperAGILE is now studying the sky in the 18 - 60 KeV energy band. It is detecting sources with advanced imaging and timing detection capabilities and good spectral detection capabilities. Several astrophysical sources has been detected and localized, including Crab, Vela and GX 301-2. The instrument has the skill to resolve correctly sources in a field of view of [-40, +40] degrees interval, with the angular resolution of 6 arcmin, and a spectral analysis with the resolution of 8 keV. Transient events are regularly detected by SA with the aid of its temporal resolution (2 microseconds) and using signal coincidence on different portions of the instrument, with confirmation from other observatories. The SA data processing scienti\ufb01c software performing at the AGILE Ground Segment is divided in modules, grouped in a processing pipeline named SASOA. The processing steps can be summarized in \\textit{data reduction}, \\textit{photonlist building}, \\textit{sources extraction} and \\textit{sources analysis}. The software services allow orbital data processing (near real-time), daily data set integration, Temporal Data Set (TDS) processing and TDS processing with source target optimization (TDS\\_SRC). Automatic data processing monitoring and interactive data analysis is possible from an internet connected workstation, with the use of SA data processing Web services. Many solutions were implemented in order to achieve fault tolerance. Archive management and data storage are performed with the help of relational database instruments. ", "introduction": "\\begin{deluxetable}{|c|l|l|} \\tablecaption{SASOA software Data Processing Stages \\label{E.16-tbl-1}} \\startdata \\textbf{\\begin{tiny}n.\\end{tiny}} &\\textbf{\\begin{tiny}STAGE\\end{tiny}} & \\textbf{\\begin{tiny}Description\\end{tiny}} \\nl \\hline \\hline \\begin{tiny}1\\end{tiny} & \\begin{tiny}TRIGGER\\end{tiny} & \\begin{tiny}control if new data are transmitted\\end{tiny} \\nl \\hline \\begin{tiny}2\\end{tiny} & \\begin{tiny}DATA\\_DOWNLOAD\\end{tiny} & \\begin{tiny}if triggered download available data\\end{tiny} \\nl \\hline \\begin{tiny}3\\end{tiny} & \\begin{tiny}COMMANDS\\_PARSING\\end{tiny} & \\begin{tiny}parse run options\\end{tiny} \\nl \\hline \\begin{tiny}4\\end{tiny} & \\begin{tiny}LOCAL\\_SETTINGS\\end{tiny} & \\begin{tiny}set hosts and data coordinates \\end{tiny} \\nl \\hline \\begin{tiny}5\\end{tiny} & \\begin{tiny}DATA\\_FETCH\\end{tiny} & \\begin{tiny}load data input for actual run in local buffer \\end{tiny} \\nl \\hline \\begin{tiny}6\\end{tiny} & \\begin{tiny}FILE\\_NAMING\\end{tiny}& \\begin{tiny}creates file names following the standard\\end{tiny} \\nl \\hline \\begin{tiny}7\\end{tiny} & \\begin{tiny}CORRECTION\\end{tiny}& \\begin{tiny}perform data correction task\\end{tiny} \\nl \\hline \\begin{tiny}8\\end{tiny} & \\begin{tiny}ORBIT\\_REBUILDING\\end{tiny}& \\begin{tiny}rebuild data related on a given orbit\\end{tiny} \\nl \\hline \\begin{tiny}8\\end{tiny} & \\begin{tiny}DATA\\_REDUCTION\\end{tiny}& \\begin{tiny}data separation, reconstruction and equalization\\end{tiny} \\nl \\hline \\begin{tiny}9\\end{tiny} & \\begin{tiny}PHOTONLIST\\_BUILDING\\end{tiny}& \\begin{tiny}analyse data of the reduced photon list\\end{tiny} \\nl \\hline \\begin{tiny}10\\end{tiny} & \\begin{tiny}ATTITUDE\\_CORRECTION\\end{tiny}& \\begin{tiny}correction for the attitude wobbling\\end{tiny} \\nl \\hline \\begin{tiny}11\\end{tiny} & \\begin{tiny}IMAGING\\end{tiny}& \\begin{tiny}image extraction \\end{tiny} \\nl \\hline \\begin{tiny}12\\end{tiny} & \\begin{tiny}FILING\\end{tiny}& \\begin{tiny}archive produced data\\end{tiny} \\enddata \\end{deluxetable} The SASOA (Super Agile Standard Orbital Analysis) pipeline is composed by a succession of elaboration tasks applied on the SA detected eventlist, the output of a certain task is the input of the following. Macro-tasks launch other subfunctions, some of them are implemented by external scripts and batch programs. In SASOA, a daemon running an endless loop provides to check if data for the incoming contacts are available and when they are found, the trigger for contact data processing is given. Data for a given contact mostly contain the data related to measurements acquired by the AGILE instruments during the last orbit. The failure of a single task of the pipeline may block the production of the final output data for a single run but the failure of a single contact run does not stop the daemon for the next incoming triggers and related data processing run. Different kinds of triggers can be used: the first way used to trigger the sasoa pipeline is to control if a file with extension .ok is generated by the preprocessing system for the incoming contact data, i. e. if a file named \"VC1.002469.009.ok\" is present, data for the contact 2469 are correctly generated and available on the data area of the preprocessing machine. Another trigger modality is to control if a value in a database of the ASDC Data Center is changed. When the system has assured that data are present, then the data are downloaded from the \\textit{gtb} server in Bologna using the \\textit{wget} utility (see M. Trifoglio et Al, \u201c\\textit{Archiving the AGILE Level-1 telemetry data}, Astronomy and Astrophysics, 2008\u201d, ), data from ASDC are downloaded with \\textit{wget} also, but they have to be downloaded one by one (mirror and -A option are disabled) to avoid transmission band overloading. Data products (histograms, plots, statistics \\& reports) are generated at different levels, hosts coordinates, data directories paths and other configurations of the processing system are set. The most important input telemetry data used for SA instrument data analysis are eventlist data (SA TM 3905), attitude data (SA TM 3914), and ephemerides data (SA TM 3916). The file names for i/o file of all tasks are generated following the official naming convention, the correction stage performs time synchronization among input data files and other corrections. Then some actions useful to rebuild the orbital organization of the data are performed: if the contact contains more than 1 orbit, then data are divided in the respective parts and the pipeline is launched on the first part of data. After the run end, the system is triggered on the 2nd part. In the Data Reduction stage the eventlist from the previous stage is taken as input (complete orbit or orbital part), with efficiency files and then a reduced output eventlist is generated. In the Sources Extraction stage, the data are corrected removing wobbling effects, then the exposure is evaluated using informations regarding Earth Occultation and SAGA time periods. After these steps the imaging procedure is applied and sources lists are generated, giving position and flux for the detected sources. \\\\ \\begin{figure}[t] \\epsscale{0.70} \\plotone{E.16_1.eps} \\caption{TDS processing schema} \\label{E.16-fig-1} \\end{figure} \\begin{figure}[t] \\begin{center} \\epsscale{0.70} \\includegraphics[angle=-90, width=8cm]{E.16_2.eps} \\caption{TDS SRC software GUI} \\label{E.16-fig-2} \\end{center} \\end{figure} \\begin{figure}[t] \\epsscale{0.70} \\plotone{E.16_3.eps} \\caption{SA software monitor} \\label{E.16-fig-3} \\end{figure} ", "conclusions": "The SA scientific processing system is ready, enabling the access from web solve installation \\& portability problems. Processing stations will be added to fit with the actual number of data analysis software users. All SA software is portable, we have in use running versions of the whole processing system under Linux OpenSUSE (until 10.3 version, for 32 \\& 64bit architecture)." }, "0911/0911.2219_arXiv.txt": { "abstract": "While limited to low spatial resolution, the next generation low-frequency radio interferometers that target 21 cm observations during the era of reionization and prior will have instantaneous fields-of-view that are many tens of square degrees on the sky. Predictions related to various statistical measurements of the 21 cm brightness temperature must then be pursued with numerical simulations of reionization with correspondingly large volume box sizes, of order 1000 Mpc on one side. We pursue a semi-numerical scheme to simulate the 21 cm signal during and prior to Reionization by extending a hybrid approach where simulations are performed by first laying down the linear dark matter density field, accounting for the non-linear evolution of the density field based on second-order linear perturbation theory as specified by the Zel'dovich approximation, and then specifying the location and mass of collapsed dark matter halos using the excursion-set formalism. The location of ionizing sources and the time evolving distribution of ionization field is also specified using an excursion-set algorithm. We account for the brightness temperature evolution through the coupling between spin and gas temperature due to collisions, radiative coupling in the presence of Lyman-alpha photons and heating of the intergalactic medium, such as due to a background of X-ray photons. The hybrid simulation method we present is capable of producing the required large volume simulations with adequate resolution in a reasonable time so a large number of realizations can be obtained with variations in assumptions related to astrophysics and background cosmology that govern the 21 cm signal. ", "introduction": "Observations of the 21 cm line of neutral hydrogen are currently considered to be one of the most promising probes of the epoch of reionization (EoR) and possibly even the preceding period, during the so called dark ages \\citep{madau97, loeb04, gnedin04, furlanetto04a, zaldarriaga04,sethi05, bharadwaj05, morales03, loeb04}. Given the line emission, leading to a frequency selection for observations, the 21 cm data provides a tomographic view of the reionization and pre-reionization process \\citep{santos05, furlanetto04c} and can also be an exquisite cosmological probe providing complimentary information when compared to cosmic microwave background anisotropies \\citep{mao08,santos06,mcquinn06,bowman07}. New large area radio interferometers are now being developed, with the promise of measuring this signal, namely, LOFAR\\footnote{http://lofar.org} in the Netherlands, MWA\\footnote{http://web.haystack.mit.edu/arrays/MWA} in Australia and the low-frequency extension of the planned Square Kilometre Array (SKA\\footnote{http://www.skatelescope.org}). Motivated by the observational possibilities offered by the current and upcoming low frequency radio interferometers a great deal of effort has been underway in order to fully understand and generate the expected 21 cm signal that will be seen by these experiments (see \\citealt{furlanetto06c} for a review). Analytical models can be very useful to quickly generate the signal with high dynamic range, in particular the power spectrum of 21 cm brightness temperature fluctuations \\citep{furlanetto04b,furlanetto06a,sethi05,barkana07} and for testing the ability of a given experiment to constraint cosmological and astrophysical parameters \\citep{santos06,mcquinn06,mao08}. The analytical models have also been useful to understand the possible contributions to the 21 cm signal at high redshifts \\citep{barkana05,pritchard06}, although it is at low redshifts ($z\\lesssim 10$) that they seem to provide a better description of the 21 cm 2-point correlation function \\citep{santos08}. However, these models still have problems dealing with the spatial distribution of the reionization process, such as bubble overlap and ignores complicated astrophysics during reionization. Numerical simulations, on the other hand, can potentially provide an improved description of the 21 cm brightness temperature signal from first principles. They typically involve a combination of a N-body algorithm for dark matter, coupled to a prescription for baryons and star formation (often complemented by a hydrodynamical simulation), plus a full radiative transfer code to propagate the ionizing radiation \\citep{gnedin00a, razoumov02,sokasian03,ciardi03,kohler05,iliev06,zahn06,mcquinn07,shin07,trac06,baek09}. These simulations have the advantage of properly dealing with the spatial distribution of the fields (such as bubble overlapping or the non-sphericity of the bubbles) and can be used to extract more information than just the 21 cm anisotropy power spectrum. The disadvantage is that they are slow to run, being constrained in dynamical range to sizes typically smaller than 143 Mpc, and producing a large set of simulations for cosmological and astrophysical parameter estimates is unrealistic. Instead the 21 cm community makes use of hybrid approaches involving simulations and analytical techniques. \\citet{thomas09} generate faster simulations by post-processing a dark matter simulation with a 1-D radiative transfer code to create bubbles around the sources of radiation embedded in the dark matter halos. \\cite{mesinger07} have taken a hybrid approach between the analytical models \\citep{furlanetto04b,furlanetto04c} and the spatial description provided by the full numerical simulations. This is based on 3-D Monte Carlo realizations of the dark matter density field combined with an ``analytical prescription'' in order to generate the halo and ionization fields. \\citet{zahn06} also considered a similar analytical prescription but based directly on the linear density field (see also \\citealt{alvarez09}). Although much faster than previous numerical approaches, the hybrid approach of \\citet{mesinger07} is still constrained to ``small'' sizes depending on the amount of shared memory available in the computer and to low redshifts since it neglects the spin temperature evolution for the 21 cm calculation. It is useful to note that the proposed low-frequency interferometers that plan to observe the 21 cm signal from reionization have low spatial resolution capabilities but very large fields of view - 5x5 deg$^2$ or more, requiring boxes of at least 1000 Mpc in size if we want to properly simulate a single field of view and pass such a simulation through the observation pipeline in order to understand how instrumental noise and systematic effects impact the observations. At the same time we need to be able to resolve halos of the order of 10$^8 M_\\odot$ corresponding to a cooling temperature of $10^4$ K that is deemed necessary to host the ionizing sources responsible for reionization. This implies a resolution of 0.14 Mpc requiring the use of large boxes with $(7000)^3$ cells, making even a hybrid method not useful. Moreover, at high redshifts we need to take into account detailed physics that determine the 21 cm brightness temperature, such as the heating of the inter galactic medium (IGM) or the Lyman-$\\alpha$ coupling of the spin temperature to the gas temperature, which makes it a challenging exercise to implement a complete radiative transfer/gas dynamics algorithm in a large volume simulation box. In this paper, we propose a semi-numerical technique, partially following the hybrid approach prescribed in \\citet{mesinger07}, capable of quickly generating an end-to-end simulation of the 21 cm signal even at high redshifts when the effect of the spin temperature is non-negligible. Moreover, this method can be used to simulate very large volumes (e.g. 1000 Mpc), crucial to simulate the field-of-view of next generation of radio telescopes, without sacrificing the speed or requiring unpracticable amounts of computer memory. The code to generate this type of simulation will be provided publicly online\\footnote{\\url{http://www.simfast21.org}} and it will be subject to continuous improvement through calibration against full radiative transfer/hydrodynamic simulations. The large volume simulations created with this method will be part of the SKA Design Studies Simulated Skies ($S^3$) initiative and we hope this approach can prove useful to generate sky models for the future 21 cm experiments, which is crucial to test for calibration issues and foreground removal as well as to study the features of the 21 cm signal that can be probed by a variety of experiments and to check the dependence of various statistics extracted from the observations on the underlying astrophysical and cosmological parameters. In the next Section we introduce the algorithm to generate the simulation of the 21 cm signal up to high redshifts, while in Section~3 we explain how to extend the simulations to very large volumes. For results presented here, we assumed a flat $\\Lambda$CDM universe with the following cosmological parameters: $\\Omega_m=0.28$, $\\Omega_b=0.046$, $h=0.70$, $n_s=0.96$ and $\\sigma_8=0.817$ based on the results from WMAP5, BAO and SN (see \\citealt{hinshaw09} and references therein). Throughout the paper we quote all quantities in comoving units unless otherwise noted. ", "conclusions": "We presented a semi-numerical method capable of quickly generating end-to-end simulations of the 21 cm signal even at the high redshifts where the spin temperature is non-negligible. The algorithm allows to generate brightness temperature boxes with very large volumes, e.g. (1000 Mpc)$^3$, crucial to properly simulate the field of view of the next generation of radio-telescopes, without sacrificing the speed or requiring unfeasible computer resources\\footnote{During the review process of this paper another algorithm was presented by \\citet{mesinger10}.}. The corresponding code (SimFast21) implements the following prescription: \\begin{itemize} \\item Monte Carlo generation of the dark matter density and velocity fields at z=0 assuming Gaussian distribution functions. \\item Determination of the halo catalog using an excursion-set formalism on the linear density field at any given redshift. For large volumes a biased Poisson sampling is used to introduce small mass halos. \\item Adjustment of the halo and dark matter locations using the Zel'dovich approximation. \\item Simulation of $x_i({\\bf x},z)$ from FFTs of the previous boxes. \\item Extraction of the star formation rate density from the halo mass distribution. \\item Extraction of gas temperature fluctuations $T_K({\\bf x},z)$ . \\item Calculation of the Ly$_\\alpha$ coupling and collisional coupling parameters. \\item Determination of the 21 cm brightness temperature boxes using the above quantities. \\end {itemize} The computational time required for the above steps depends mainly on the volume to be covered (our bubble filtering procedure is now pratically independent of the ionization fraction due to the FFT based algorithm we are now using). This basically affects the halo determination running time, since larger volumes imply a larger number of halos and operations for overlap checking. Nevertheless, the typical time for the computation of the above steps is about 2 hours for the simulation with L=1000 Mpc, which can be considered remarkably fast. We are expecting that further optimization of the code can still reduce this time. Although much faster than hydrodynamical numerical simulations, our analysis shows that relevant quantities such as the halo mass function, the halo mass power spectrum, the star formation rate or the ionization fraction power spectrum are all consistent with the numerical simulations with which we compared our results to. In particular, the ionization fraction power spectrum is similar to the one obtained with radiative transfer codes (RT) when comparing to the same volume. We still find some differences on the largest scales when comparing to simulations with relatively small volumes (143 Mpc), which we can relate to the production of larger HII bubbles. These differences become much smaller as we increase the simulation volume and the change of size in the largest bubbles becomes less important when compared to the overall size of the simulation. The dependence on the astrophysical parameters of the simulation was encoded in three functions: the ionization efficiency, $\\zeta$, the Ly$_\\alpha$ spectral distribution function of the sources, $\\epsilon_\\alpha$ and the X-ray spectral distribution function, $\\epsilon_X$. These functions can be easily changed for a model of our choice and the code can then quickly generate new simulations of the signal (even faster if we keep the same cosmology). By combining all the unknowns into a physically meaningful small set of parameters we can easily probe the huge intrinsic parameter space available at high redshifts probed by 21 cm observations. This versatility is included in our algorithm. The possibility to generate reionization simulations from scratch without the need for supercomputers is a great advantage allowing to experiment with different models. Moreover, this allows room for improvement through calibration with full numerical simulations. The code can be a useful tool to generate sky models for future 21 cm experiments, important to test the observation pipeline, study how the foregrounds affect the observations and can be separated given a noise model, and to develop optimal estimators for signal extraction. The code will be released for public use in due course and we welcome the community participation in its updating and upgrading as well as development of additional applications on observational probes of reionization beyond the 21 cm observations." }, "0911/0911.0523.txt": { "abstract": "{}{}{}{}{} % 5 {} token are mandatory \\abstract % context heading (optional) % {} leave it empty if necessary {} % aims heading (mandatory) { We investigate the properties and the environment of radio sources with optical counterpart from the combined VLA-COSMOS and zCOSMOS samples. The advantage of this sample is the availability of optical spectroscopic information, high quality redshifts, and accurate density determination. } % methods heading (mandatory) {By comparing the star formation rates estimated from the optical spectral energy distribution with those based on the radio luminosity, we divide the radio sources in three families, passive AGN, non-passive AGN and star forming galaxies. These families occupy specific regions of the 8.0-4.5 $\\mu$m infrared color- specific star formation plane, from which we extract the corresponding control samples.} % results heading (mandatory) {Only the passive AGN have a significantly different environment distribution from their control sample. The fraction of radio-loud passive AGN increases from $\\sim 2\\% $ in underdense regions to $\\sim 15 \\% $ for overdensities (1+$\\delta$) greater than 10. This trend is also present as a function of richness of the groups hosting the radio sources. Passive AGN in overdensities tend to have higher radio luminosities than those in lower density environments. Since the black hole mass distribution is similar in both environments, we speculate that, for low radio luminosities, the radio emission is controlled (through fuel disponibility or confinement of radio jet by local gas pressure) by the interstellar medium of the host galaxy, while in other cases it is determined by the structure (group or cluster) in which the galaxy resides. } % conclusions heading (optional), leave it empty if necessary {} ", "introduction": "Radio sources have been traditionally classified into two broad classes depending upon their emission mechanism: AGN-driven emission has been supposed for early type galaxies, while a star forming origin of the radio power has been invoked for late type galaxies \\citep{Condonreview}. However, not all optical galaxies exhibit radio emission (at least at the depth of current surveys) and it is important to investigate the physical mechanism which triggers this emission. For AGN-induced activity, the difference between radio quiet and radio loud galaxies is naturally connected with the availability of fuel for the central engine, the black hole mass, and its emission efficiency \\citep[see e.g.][and references therein]{Shabala}, while for star forming objects it is closely related to the strength of the star formation episode. Heating by a central AGN is also thought to be important for the host galaxy evolution, providing the energy to stop the central cooling that drives gas to the central black hole in an intermittent feedback mechanism dubbed \"radio mode\" \\citep[see][]{Croton,Ciotti}. The same intermittent mechanism is indirectly observed as bubbles in the intra-cluster medium \\citep{Birzan2004} or in the restarted activity of radio sources \\citep{Venturi2004,Bardelli2002}. This heating emission is thought also to be able to suppress the star formation and to be an important mechanism for creating the galaxy color bimodality \\citep{Croton}. Moreover, it is a fact that many radio AGN reside in early type galaxies \\citep{LO96}, a class of objects which are preferentially found in high density environments. It has been proposed that cluster/group mergers, galaxy-galaxy mergings or at least tidal disturbance caused by close encounters between galaxies would increase radio activity. For early type galaxies, these phenomena could drive gas more efficently toward the center, while for late type galaxies, these encounters would trigger and/or enhance star formation \\citep[see e.g.][]{Bekki,Vollmer}. The radio luminosity of the AGN also depends on the interaction of the jet coming from the black hole with the surrounding gas of the host galaxy and/or of the hosting cluster \\citep[see e.g.][and references therein]{Shabala}. For these reasons, some degree of dependence of radio activity on the environment is expected. Although the problem is clear from the qualitative description, controversial results are present in the literature and for mainly two reasons. First of all, very few papers are primarly based on radio data. Usually, papers in the literature consider AGN selected on the basis of their optical (mostly estimated from line diagnostics) or X-ray activity \\citep[see][and references therein]{Kauffmann2004}. It seems that the family of optical/X-ray AGN inhabits similar environments to non-active galaxies, with a preference for lower densities at higher stellar masses \\citep{Kauffmann2004,Silverman}. However, \\cite{Best05} claim that the phenomena generating the emission lines and the radio emission are statistically independent, even if recent results indicate that radio loud emission line AGN favour higher densities than radio quiet ones \\citep{Kauffmann08} \\citep[see also ][for a study in the IR]{Caputi09}. Secondly, only specific environments have been considered, mainly clusters or groups \\citep{Miller,Hill, Giacintucci}, lacking therefore the coverage of a large dynamical range in densities. The first paper that studied in a complete statistical way, both from the radio and optical side, the environment of radio galaxies was that of \\cite{Bestenv} . This paper uses data from 2dF Galaxy Redshift Survey \\citep{Colless} and the NRAO VLA Sky Survey \\citep{NVSS} with a well defined estimator of density, i.e. the projected density of the 10$^{th}$ nearest neighbor. Moreover, groups (and clusters) are found with the standard friend-of-friends detection method. The redshift range is $0.0210$). Similarly, such trend is also found (but at a low significance level) as a function of richness of galaxy groups. Studying the radio luminosity-stellar mass plot, we found that for luminosities lower than log P(W Hz$^{-1}$)=23.5 there is a correlation between the two quantities, that disappears at higher powers. The correlation between stellar mass and radio luminosity is not unexpected considering the relation between the accretion on to a black hole in the \"radio mode\" to the black hole mass and hot gas fraction \\citep[see equation 10 of][]{Croton}. As stated by \\cite{Croton}, the hot gas fraction is approximately constant for galaxies with $v_{vir}>150$ km s$^{-1}$ and therefore the black hole luminosity scales with the black hole mass, which correlates with the stellar mass. Interestingly, half of the objects in the low radio luminosity regime reside in low densities, while the others are in overdense regions. The correlation disappears for radio luminosities higher than log L$_{Radio}$(W Hz$^{-1}$)=23.5. Almost all of these radio luminous objects reside in overdense region. On the other hand, no difference is present in the black hole mass between high and low densities and therefore it seems that in richer environments the emission mechanism is more efficient than in low density regions. Therefore, we can tentatively say that at low density the emission is determined by the host galaxy stellar mass and at high density by the host structure (group/cluster) in which the galaxy resides. The higher luminosities in denser environments could be caused by a greater fuel supply (due to the central cooling of the gas) and/or to the fact that also outside the galaxy there is a significant gas density to confine the radio jet. On the other hand, we already know that ellipticals at the center of galaxy groups are richer in hot gas than isolated and non-dominant ones \\citep{Helsdon} and this implies that a mechanism driving gas within dominant galaxies is or has been active. This can also justify a longer time spent by the central engine in the \"switched on\" status, as suggested by the increasing ratio of passive AGN to control sample as a function of density. It would be interesting to investigate whether the high density AGN have luminosities that correlate with the central cooling flow, following the relation between radio luminosity and accretion mass found by \\cite{Mittal}. Note that our findings are consistent with those of both \\cite{Bestenv}, \\cite{Mandelbaum}. Finally, the difference in overdensity between radio-loud and control samples remains constant across the studied redshift range, implying that no significant evolution in this phenomenon is present. {\\it Non-passive AGN}: This is a less well-defined family. In principle, in this class one can find not only AGN but also objects such as post--starburst galaxies. In this case, radio emission disappears on a time scale of $10^{8}$ years, which could be longer than the disappearing time of the star formation signatures in optical. However, we expect that these objects are rare because of the implied short time scales. In fact, only two of the radio detected post--starbursts of the \\cite{Verganipsb} sample and one of the upper limits lie within our AGN region. This means that post-starburst galaxies contaminate our sample by $4 \\%$. The non-passive AGN population occupies a small region of the infrared color-specific star formation rate plot, which is approximately equivalent to the green valley defined in the optical bands. These galaxies are objects with spectrophotometric type earlier than Sa-Sb and applying the standard diagnostics based on the emission line ratios we found that $50\\% $ are classified as optical AGN and the remaining are star forming galaxies. The ZEST morphologies of non-passive AGN are equally divided between early types and spirals (with a $\\sim 8 \\% $ of irregulars), while the control sample has percentages of $\\sim 60 \\%$ and $40 \\%$, respectively. The radio emitting objects of this class follow the same environment distribution of the corresponding control sample galaxies and do not show the luminosity-density behaviour. {\\it Star Forming galaxies} The composite spectra of radio emitting star forming galaxies is redder than the control sample with a smaller [OII] and higher H$\\alpha$ equivalent width. Considering that the radio detected star forming galaxies are objects where the star formation rate is on average higher than the undetected ones, this implies that the dust content is higher. This increase of dust extinction with increasing star formation is already known \\citep[see e.g.][]{Pannella} and could justify the relation between the emission lines and the radio SFR estimators shown in the lower panel of Figure \\ref{sfrsfr}. The ZEST morphological classification shows that the radio detected objects tend to be of earlier type than the control sample, as found for the spectrophotometric types. The radio emitting star forming galaxies show an apparent difference in density distribution with respect to the control sample, but we have shown that this is due to the different stellar mass distributions of the two samples. Correcting the control sample with the fraction of radio emitting objects, no differences in environment are present between radio and control sample galaxies. This is consistent with the conclusion of \\cite{Kauffmann2004} that the star formation does not depend on the density for scales larger that the Megaparsec." }, "0911/0911.0670_arXiv.txt": { "abstract": "This paper presents a numerical study over a wide parameter space of the likelihood of the dynamical bar-mode instability in differentially rotating magnetized neutron stars. The innovative aspect of this study is the incorporation of magnetic fields in such a context, which have thus far been neglected in the purely hydrodynamical simulations available in the literature. The investigation uses the \\cosmospp code which allows us to perform three-dimensional simulations on a cylindrical grid at high resolution. A sample of Newtonian magneto-hydrodynamical simulations starting from a set of models previously analyzed by other authors without magnetic fields has been performed, providing estimates of the effects of magnetic fields on the dynamical bar-mode deformation of rotating neutron stars. Overall, our results suggest that the effect of magnetic fields is not likely to be very significant in realistic configurations. Only in the most extreme cases are the magnetic fields able to suppress growth of the bar mode. ", "introduction": "\\label{sec:intro} Rotating neutron stars formed following the gravitational collapse of a massive stellar iron core or the accretion-induced collapse of a white dwarf can be subject to various nonaxisymmetric instabilities depending on the amount and degree of differential rotation. The prospects of detection of gravitational radiation from newly born rapidly rotating neutron stars by the current and future generations of kHz-frequency, ground-based gravitational wave interferometers highly motivate the investigation of such instabilities. In particular, if the rotation rate is high enough and shows a high degree of differentiation, the star is subject to the so-called $m=2$, {\\it dynamical} bar-mode instability driven by hydrodynamics and gravity, with $m$ being the order of the azimuthal nonaxisymmetric fluid mode $e^{\\pm im\\varphi}$ . On the other hand, at lower rotation rates gravitational radiation and viscosity can drive a star {\\it secularly} unstable against bar-mode deformation. These two flavors of the bar-mode instability set in when the ratio $\\beta=T/|W|$ of rotational kinetic energy $T$ to gravitational potential energy $W$ exceeds a critical value $\\beta_{\\rm c}$. Early studies with incompressible MacLaurin spheroids in Newtonian gravity showed that the onset of the instability arises when $\\beta_c\\sim 0.27$ and $0.14$ for the dynamical and secular cases, respectively~\\citep{chandra69}. Improvements on these simplified analytic models have been achieved through numerical work. Newtonian and general relativistic analyses of the dynamical bar-mode instability are available in the literature, using both simplified models based on equilibrium stellar configurations perturbed with suitable eigenfunctions, and more involved models for the core collapse scenario. Newtonian hydrodynamical simulations have shown that the value of $\\beta_{\\rm c}$ is quite independent of the stiffness of the equation of state, provided the star is not strongly differentially rotating (see~\\citet{houser94,new00,liu02} and references therein). On the other hand, relativistic simulations~\\citep{shibata00} have yielded a slightly smaller value of the dynamical instability parameter ($\\beta_{\\rm c}\\sim 0.24-0.25$). They have also shown that the dynamics of the process closely resembles that found in Newtonian theory, that is, unstable models with large enough $\\beta$ develop spiral arms following the formation of bars, ejecting mass and redistributing the angular momentum. Further relativistic simulations~\\citep{baiotti07} have shown the appearance of nonlinear mode-coupling which can limit, and even suppress, the persistence of the bar-mode deformation. It is also worth mentioning that, as the degree of differential rotation becomes higher and more extreme, Newtonian simulations~\\citep{shibata02,shibata03} have also shown that rotating stars are dynamically unstable against bar-mode deformation even for values of $\\beta$ of order 0.01. Whether the requirements for the development of the instability inferred from numerical simulations are met by the collapse progenitors remains unclear. Observations of surface velocities imply that a large fraction of progenitor cores are rapidly rotating. However, it has been shown that magnetic torques can spin down the core of the progenitor, leading to slowly rotating neutron stars at birth~\\citep{spruit98}. The most recent computations of the evolution of massive stars, which include angular momentum redistribution by magnetic torques and spin estimates of neutron stars at birth, lead to core collapse progenitors which do not seem to rotate fast enough to guarantee the unambiguous growth of the canonical bar-mode instability~\\citep{heger05,ott06}. These estimates are in agreement with observed periods of young neutron stars. However, rapidly-rotating cores might be produced by an appropriate mixture of high progenitor mass and low metallicity. Recent estimates suggest that about 1\\% of all stars with masses larger than 10 solar masses will produce rapidly rotating cores~\\citep{woosley06}. Recent relativistic simulations of a large sample of rotational core collapse models carried out by~\\citet{dimmelmeier08}, which include a state-of-the-art treatment of the relevant physics of the collapse phase and realistic precollapse rotational profiles~\\citep[see also][]{ott07}, have shown that the critical threshold of the dynamical bar-mode instability is never surpassed, at least early after core bounce. However, a large set of the models investigated by~\\citet{ott07} and~\\citet{dimmelmeier08} do show that models with sufficiently differential and rapid rotation are subject to the low$-\\beta$ instability. In addition it is also worth mentioning the simulations of~\\citet{baiotti08} which show that hypermassive differentially rotating neutron stars form following the merger of low-mass binary neutron stars, when modeled as polytropes. The resulting hypermassive star undergoes a persistent phase of bar-mode oscillations, emitting large amounts of gravitational radiation prior to its delayed collapse to a black hole. An important piece of physics that the existing numerical work has so far neglected is the presence of magnetic fields. The amplification and emergence of strong magnetic fields in neutron stars from initially weak magnetic field configurations in the pre-collapse stellar cores are currently under investigation~\\citep[see e.g.,][and references therein]{burrows07,cerda-duran08}. We note, however, that the weakest point of all existing magneto-rotational core collapse simulations to date is the fact that both the strength and distribution of the initial magnetic field in the core are unknown. In this paper, we show results from a detailed numerical study of the effects that magnetic fields may have on the dynamical bar-mode instability in differentially rotating magnetized neutron stars. In particular, we investigate how sensitive the onset and development of the instability is to the presence of magnetic fields, as well as the role played by the magnetorotational instability (MRI) and magnetic braking mechanisms to alter the angular momentum distribution in the star and possibly suppress the bar-mode instability. Our study is motivated by the potential astrophysical implications that the presence of strong magnetic fields may have for post-bounce core collapse dynamics and, in turn, for gravitational wave astronomy. The uncertainty of the strength and distribution of magnetic fields in collapse progenitors is reflected in our somewhat ad hoc parameterization of the field configuration through a large sample of equilibrium models of rapidly and highly differentially rotating neutron stars. Our sample of Newtonian magnetohydrodynamical (MHD) simulations is based upon the set of purely hydrodynamical models previously analyzed by~\\citet{new00}. The simulations are performed using the covariant (and adaptive mesh refinement) code \\cosmospp~\\citep{anninos03a,anninos03b,anninos05,fragile05} which allows us to perform three-dimensional simulations on a logarithmically scaled cylindrical grid at high resolution. A number of equilibrium models of rapidly rotating stars with different values of the rotational instability parameter ($\\beta=T/|W|$) and magnetic plasma beta ($\\beta_B=P/P_B$), and different polytropic equations of state are constructed, introducing also different configurations for the magnetic field distribution (of both poloidal and toroidal varieties) and field strengths. The equilibrium models are perturbed by seeding small random perturbations in order to initiate the onset of the bar-mode deformation. The organization of the paper is as follows. Section~\\ref{sec:methods} discusses our basic formalism, numerical methods, diagnostics, and the construction of initial data; Section~\\ref{sec:results} presents our results in two subsections, one for initially toroidal field configurations and one for poloidal. We conclude with a summary and discussion of our results in Section~\\ref{sec:conclusion}. ", "conclusions": "\\label{sec:conclusion} We have studied the growth of the dynamical bar-mode instability in differentially rotating magnetized neutron stars through a set of numerical Newtonian MHD calculations. The calculations explored both toroidal and poloidal initial field distributions of differing strengths, as well as the role of the equation of state. For our initially toroidal field configurations, field amplification always saturated at a level insufficient to strongly affect the dynamics of the bar mode. Even the most extreme case (TG53B1, $\\beta_\\mathrm{B,min}=1$), with equal initial magnetic and hydrodynamic pressures within the toroidal loop, gave only a $\\sim30$\\% change in gravitational wave properties. Otherwise evolution proceeded quite similarly to our unmagnetized reference cases. The effects were larger for the poloidal configurations. In the most extreme case considered (model PG53B10HR with $\\beta_\\mathrm{B,min}=10$), the magnetic field was sufficient to completely suppress the formation of a bar. However, in that case, we started with the magnetic field already within a factor of $\\sim$30 of being dynamically more important than thermal pressure in some portions of the star. It then only took a few dynamical times of field growth for $\\beta_B$ to approach unity, a timescale that is much shorter than the bar deformation time for azimuthal Fourier modes to reach appreciable amplitudes. Also, due to the computationally demanding nature of evolving strong poloidal fields, we cannot verify that this result will not change with increased grid resolution. For less extreme initial conditions, the effects of including poloidal field components were consistently more modest with decreasing initial field amplitude. These lower amplitude calculations are also less demanding computationally and can, like the toroidal cases, be more easily checked for convergence. The principle effects of introducing poloidal fields are in systematically delaying the onset of the bar mode and in suppressing both Fourier mode and gravitational wave amplitudes, all of which become negligible with initial field amplitudes below the threshold of $1/\\beta_\\mathrm{B,min} \\lesssim 10^{-2}$. Overall, our results suggest that the effect of magnetic fields on the emergence of the bar-mode instability in neutron stars is not likely to be very significant. Particularly considering that collapse progenitor models predict realistic field configurations that are dominantly toroidal in nature with toroidal and poloidal components of order $10^{10}$G and $10^6$G, respectively \\citep{heger05}, substantially below most of the field configurations we have considered. Thus, except in special cases where neutron stars are born very highly magnetized, we might still expect them to be good gravitational wave sources if their rotational kinetic energies exceed the critical bar-mode instability parameter." }, "0911/0911.2079_arXiv.txt": { "abstract": " ", "introduction": "The INTEGRAL (INTErnational Gamma-Ray Astrophysics Laboratory) satellite produces many observations since its launch in 2002, not only in gamma part of the spectra. The onboard OMC (Optical Monitoring Camera) was designed to obtain the observations in optical $V$ passband. These observations are in fact only a by-product of the mission, but nowadays there are many observations available. Despite the fact that the database of these measurements is freely available on internet, the analyses are still very rare. The most recent one using the OMC data is that by \\cite{2009IBVS.5881....1J} about a new $\\beta$~Cep star. This investigation is directly following our previous papers (\\citealt{Zasche2008NewA} and \\citealt{2009NewA...14..129Z}). The selection criteria used here were also the same: maximum number of data points and non-existence of any detailed light-curve analysis of the particular system. There were 33 systems selected for the present paper. ", "conclusions": "The light-curve analyses of thirty-three selected systems have been carried out. Using the light curves observed by the Optical Monitoring Camera onboard the INTEGRAL satellite, one can estimate the basic physical parameters of these systems. Despite this fact, the parameters are still only the preliminary ones, affected by relatively large errors and some of the relevant parameters were fixed at their suggested values. The detailed analysis is still needed, especially spectroscopic one, or another more detailed light curve one in different filters. Together with a prospective radial-velocity study, the final picture of these systems could be done. Particularly, the systems V1450~Aql and CY~Lac seem to be the most interesting ones. The first one is massive semi-detached system, which shows total eclipses and the second one due to its eccentric orbit." }, "0911/0911.0760_arXiv.txt": { "abstract": "{} {We aim at determining the rotational periods and the starspot properties in very young low-mass stars belonging to the Ori OB1c star forming region, contributing to the study of the angular momentum and magnetic activity evolution in these objects. } {We performed an intensive photometric monitoring of the PMS stars falling in a field of about 10$\\arcmin\\times 10\\arcmin$ in the vicinity of the Orion Nebula Cluster (ONC), also containing the BD eclipsing system 2MASS\\,J05352184-0546085. Photometric data were collected between November 2006 and January 2007 with the REM telescope in the $VRIJHK'$ bands. The largest number of observations is in the $I$ band (about 2700 images) and in $J$ and $H$ bands (about 500 images in each filter). From the observed rotational modulation, induced by the presence of surface inhomogeneities, we derived the rotation periods. The long time-baseline (nearly three months) allowed us to detect rotation periods, also for the slowest rotators, with sufficient accuracy ($\\Delta P/P<2\\%$). The analysis of the spectral energy distributions and, for some stars, of high-resolution spectra provided us with the main stellar parameters (mass, luminosity, effective temperature, mass, age, and $v\\sin i$) which are essential for the discussion of our results. Moreover, the simultaneous observations in six bands, spanning from optical to near-infrared wavelengths, enabled us to derive the starspot properties for these very young low-mass stars.} {In total, we were able to determine the rotation periods for 29 stars, spanning from about 0.6 to 20 days. Thanks to the relatively long time-baseline of our photometry, we derived periods for 16 stars and improved previous determinations for the other 13. We also report the serendipitous detection of two strong flares in two of these objects. In most cases, the light-curve amplitudes decrease progressively from the $R$ to $H$ band as expected for cool starspots, while in a few cases, they can only be modelled by the presence of hot spots, presumably ascribable to magnetospheric accretion. The application of our own spot model to the simultaneous light curves in different bands allowed us to deduce the spot parameters and particularly to disentangle the spot temperature and size effects on the observed light curves. } {} ", "introduction": "\\label{sec:Intro} The study of the evolution of angular momentum of low-mass stars in the pre-main sequence (PMS) phase has gained great advantage from high-precision photometry of star forming regions (SFR). While during the main sequence (MS) evolutionary phase the total angular momentum of a solar-mass star is lost efficiently via the magnetic-wind breaking process \\citep[e.g.,][]{Chabo95}, in the PMS phase the presence of a circumstellar disc coupled to the star through magnetic fields permits a more efficient angular momentum transfer, thus preventing the spin-up driven by the rapid mass-accretion and contraction of the star \\citep{Shu94}. An accurate knowledge of the rotation period distribution of the solar and low-mass stars in different SFR environments is of paramount importance for understanding how the disc-locking mechanism is effective in modifying the angular momentum in the PMS stage, and for testing the reliability of the theoretical models of PMS star interiors and their evolutionary tracks. Measurements of periodic light-variations with typical amplitudes in the range 0.05--0.40 mag in the $V$ and $R$ bands, due to uneven starspot distributions on the stellar photospheres, are a direct way to obtain this physical parameter. There has been a rapid increase in the number of PMS objects with measured rotation periods over the last decade \\citep[see, e.g.,][and references therein]{Mat04,Herbst07,Irwin09}. A broad, bimodal distribution exists at the earliest observable phases (about 1~Myr) for stars more massive than 0.4~M$_{\\sun}$. The fast rotators (50--60\\% of the sample) evolve to the ZAMS with little or no angular momentum loss. The slow rotators continue to lose substantial amounts of angular momentum for up to 5 Myr, creating the even broader bimodal distribution characteristic of 30--120 Myr old clusters. Accretion disc signatures, such as H$\\alpha$ emission and infrared (IR) excess, are observed more frequently among slowly rotating PMS stars, suggesting a connection between accretion and rotation \\citep{rebu06}. According to the present picture, discs appear to influence rotation only for a small percentage of the solar-type stars and only during the first 5 Myr of their life. This time interval is comparable to the maximum life-time of accretion discs derived from NIR studies and may be a useful upper limit to the time available for forming giant planets. For the sparse population of young stars (mainly weak-T Tauri stars, wTTS, with ages between 5 and 30 Myr) in the Orion complex, \\citet{Mari07} found no evidence of bimodality for masses higher than 0.7~M$_{\\sun}$, and a median rotation period of $\\sim$\\,1.5 days, suggesting that the spin-up process at this stage has become dominant over the disc-locking effect. It is then mandatory to explore the rotational properties of as many clusters/associations as possible with ages less than 5 Myr. Indeed, in this age range, the different scenarios for angular momentum evolution proposed by \\citet{lamm} and related to different time-scales of disc-locking and stellar contraction \\citep{Hart02} are expected to produce their stronger effects. There is growing evidence that the same mechanisms operating in solar-mass stars also regulate the rotation of very-low mass (VLM) stars and brown dwarfs (BDs). Indeed, very recently many BDs showing accretion discs with physical properties similar to their analogs around higher-mass stars have been discovered \\citep[e.g.,][and reference therein]{Luhman05,Alc06,Muz06,Mer07}. However, it seems that both disc interactions and stellar winds are less efficient in braking these objects, although the statistics available for this VLM regime is rather poor. In some cases, both VLM and BD objects in star forming regions exhibit photometric light curves with high amplitudes and irregular modulations \\citep{Scholz04,Scholz05} that are usually explained by hot spots caused by mass-accretion flow from the disc \\citep{Herbst94}. The same authors remark that wTTS and some classical T Tau stars (cTTS) display instead more regular light curves with smaller amplitudes ascribable to cool photospheric spots that can survive for several stellar rotations. In this paper, we report the results of an intensive photometric monitoring, in near-infrared (NIR) and optical bands, of a sample of young low-mass stellar and sub-stellar objects (YSOs) in a small area flanking the Orion Nebula Cluster (ONC). The main goal of our observations was the study of the NIR light curve of \\object{2MASS\\,J05352184-0546085}, an eclipsing system composed of two BD components recently discovered by \\citet{stass06}. The long-lasting and intensive simultaneous optical and NIR monitoring also allowed us to study the photometric variability of many YSOs in the field of view around the eclipsing BD. Therefore, in this study we report the characterization of these objects and determine their rotational periods, as well as the properties of their starspots. The present work thus contributes to the study of the early phases of angular momentum evolution and of the behaviour of magnetic activity in young, VLM objects which still needs observational and interpretation efforts. We derived the rotation periods of our targets by means of time-series analysis of our large photometric data-set gathered in the $VRIJHK'$ bands. The physical characterization (radius, luminosity, effective temperature, mass, and age) of the objects, which is essential for any discussion on angular momentum evolution and stellar activity, is based on our photometric data as well as on literature mid-IR data from the Spitzer space telescope \\citep{rebu06} through an analysis of their spectral energy distribution (SED). For some objects, accurate physical parameters and the projected rotational velocity, $v\\sin i$, were determined by means of high-resolution spectra. The use of simultaneous multi-band light curves allows an accurate analysis of the spot properties because the different amplitude of the light curves constrains the spot temperatures and filling factors. ", "conclusions": "\\label{sec:conclusions} We presented the results of an intensive photometric monitoring of a 10$\\arcmin\\times 10\\arcmin$ field flanking the Orion Nebula Cluster (ONC) conducted during three consecutive months. The main results of our work can be summarized as follows: \\begin{itemize} \\item We detected rotation periods for 29 stars, spanning from about 0.6 to 20 days, sixteen of which are new periodic variables. Thanks to the relatively long time-baseline we measured the periods with sufficient accuracy ($\\Delta P/P<2\\%$) also for the slowest rotators. We remark that none of the stars with NIR excess has a rotation period shorter than 5\\,days, in agreement with the hypothesis that the angular momentum of the stars is moderated by interaction with the disc. \\item The analysis of the spectral energy distribution and, for some stars, the high-resolution spectra provided us with $T_{\\rm eff}$ and luminosity, and that allowed us to construct the HR diagram of these stars. We could then assign a photometric membership to the objects by a comparison with PMS evolutionary tracks and derive masses and ages. The majority of the objects with detected period are classified as Orion members and result to be younger than 10\\,Myr. \\item Our spot modelling code enabled us to derive the starspot properties for five of these star, based on the simultaneous analysis of the light curves in several optical and NIR bands. For one of these, the light curves could only be modelled with hot spots, which are likely related to magnetospheric accretion. For three stars with $T_{\\rm eff}$ in the range 4100--4400\\,K and rotation periods between 4 and 8 days we found cool spots with $\\Delta T=T_{\\rm ph}-T_{\\rm sp}$ in the range 1000--1300\\,K, i.e. larger, on average, than $\\Delta T$ typically found for the sub-giant and giant components of RS~CVn systems. Conversely, the spotted area (8--10\\% of the star surface) is smaller. These differences could be due to the fully convective internal structure of PMS stars which gives rise to a different dynamo action. For the cool ($T_{\\rm eff}\\simeq 3500$\\,K) and ultra-fast ($P\\simeq 0.56$\\,days) star \\#9, a smaller spot contrast ($\\Delta T\\simeq 270$\\,K) is found. This could be due both to the lower effective temperature and the high rotation rate. For the star with the highest modulation amplitudes (star \\#11), which also displays a remarkable near- and mid-IR excess, the light curves were not consistent neither with hot nor with cool spots. We suggest that the flux modulation could be produced by inhomogeneities of the circumstellar disk. \\item A very strong flare was detected on star \\#19 (V498\\,Ori) in the $I$ band. The temporal evolution of the event was fully resolved and we evaluated the rise and decay e-folding times as $\\tau_{r}\\simeq 2.5$\\,min and $\\tau_{d}\\simeq 24.9$\\, min, respectively. We estimated an energy released in the $I$ band of nearly $3\\cdot10^{35}$ erg and a 20 times higher energy released in the optical continuum. Another strong flare, which released an energy of about $6\\cdot10^{34}$\\,erg in the $I$ band, was observed on star\\,\\#\\,11 (\\object{NR~Ori}), which is cooler and less massive than star \\#\\,19 and displays the broadest modulation and a strong infrared excess. \\end{itemize}" }, "0911/0911.5003_arXiv.txt": { "abstract": "The final composition of giant planets formed as a result of gravitational instability in the disk gas depends on their ability to capture solid material (planetesimals) during their 'pre-collapse' stage, when they are extended and cold, and contracting quasi-statically. The duration of the pre-collapse stage is inversely proportional roughly to the square of the planetary mass, so massive protoplanets have shorter pre-collapse timescales and therefore limited opportunity for planetesimal capture. The available accretion time for protoplanets with masses of 3, 5, 7, and 10 Jupiter masses is found to be 7.82$\\times 10^4$, 2.62$\\times 10^4$, 1.17$\\times 10^4$ and 5.67$\\times 10^3$ years, respectively. The total mass that can be captured by the protoplanets depends on the planetary mass, planetesimal size, the radial distance of the protoplanet from the parent star, and the local solid surface density. We consider three radial distances, 24, 38, and 68 AU, similar to the radial distances of the planets in the system HR 8799, and estimate the mass of heavy elements that can be accreted. We find that for the planetary masses usually adopted for the HR 8799 system, the amount of heavy elements accreted by the planets is small, leaving them with nearly stellar compositions. ", "introduction": "The recent discoveries of wide-orbit massive protoplanets by direct imaging (Kalas et al., 2008; Marois et al., 2008) have presented new type of gaseous planets that giant planet formation theories must be able to explain. Work regarding possible formation scenarios has already been demonstrated by several groups (e.g., Dodson-Robinson et al., 2009b; Nero and Bjorkman, 2009). Systems with massive gaseous planets at large radial distances offer a challenge for theorists mainly because they cannot be explained easily by the core accretion model, the standard model for giant planet formation (e.g., Cameron, 1978; Pollack et al., 1996). The core accretion model fails to form giant planets at large radial distances in situ due to the low surface density and the extremely long accretion timescale (Ida and Lin, 2004; Dodson-Robinson et al., 2009b). As a result, a more detailed investigation of giant planet formation scenarios seems to be required. \\par In addition, the increasing number of transiting planets provides information regarding the planetary mean densities, and therefore, their compositions. Giant planet formation theories are also required to explain the planets' observed properties, and when possible, provide predictions that can be tested with future observations and upcoming data. Since the heavy element masses of planets can be estimated for transiting planets, further investigation and more detailed understanding of the origin of the heavy elements and the cause for the large variety in heavy elements enrichments observed (e.g., Guillot et al., 2007) is desirable. \\par While the core accretion model offers an enrichment with heavy elements that is coupled to the formation process (e.g., Pollack et al., 1996; Hubickyj et al., 2005), in the gravitational instability model (Cameron, 1978; Boss, 1997) the planet's enrichment with solids is so far considered as an independent part of the actual formation mechanism. In the disk instability model giant protoplanets are initially formed with stellar compositions, and can be enriched with heavy elements after their formation by accretion of planetesimals (e.g., Helled et al., 2006). The total mass of solids that can be accreted depends on the planetary mass, its radial distance, the available solid material for capture, and the protoplanet's efficiency in accreting the solid planetesimals (Helled and Schubert, 2009). \\par A gravitationally unstable condensation of a few Jupiter masses in a protoplanetary disk evolves through three phases (DeCampli and Cameron, 1979; Bodenheimer et al., 1980). First, it contracts quasi-statically with cool internal temperatures on the order of a few hundred K, with hydrogen in molecular form, and with radii a few thousand times the present Jupiter radius. Second, once the central temperature reaches about 2000 K, the H$_2$ dissociates and initiates a dynamical collapse of the entire planet, ending only when the radius has decreased to a few times the present Jupiter radius. Third, it contracts and cools on the long time scale of $\\sim 10^9$ yr. The first phase is the crucial one for the capture of solids, and the time scale of this phase, which is about $4 \\times 10^5$ yr for 1 Jupiter mass, determines the amount of material that can be captured. As the protoplanets contract quasi-statically during this phase, planetesimals in their feeding zone can be slowed down by gas drag, and get absorbed in the planetary envelopes (e.g., Helled and Schubert, 2009). As a protoplanet contracts, fewer planetesimals pass through its envelope, and instead of being accreted they get ejected from the protoplanet's vicinity. Efficient planetesimal accretion ends once the central temperature reaches $\\sim$ 2000 K and molecular hydrogen begins to dissociate. Whether gaseous protoplanets can form as a result of a local gravitational instability in a protoplanetary disk is still a matter of debate. The majority of theoretical and numerical work suggests that the gravitational instability scenario for planet formation is rather unlikely, at least at relatively small radial distances (e.g., Rafikov, 2007; Cai et al., 2009). Other results suggest that protoplanetary disks can break into gaseous protoplanets (e.g., Boss, 1997), although this may be limited to special conditions (Mayer et al., 2007). The extrasolar giant planets observed by direct imaging (Kalas et al., 2008; Marois et al., 2008) have raised the suggestions that such massive gaseous planets at large radial distances were formed as a result of gravitational instabilities (Dodson-Robinson et al., 2009b; Nero and Bjorkman, 2009). The fact that HR 8799 host star is found to be metal-poor ([Fe/H]=-0.47) supports the idea that its planetary system was formed via disk instabilities. More advanced numerical simulations with longer integration times and further theoretical work seem to be required before a robust conclusion regarding this formation mechanism and its limitations can be made. While recent theoretical work (Boley, 2009; Rafikov, 2009; Nero and Bjorkman, 2009) suggests that gravitational instability can form planets only at distances greater than 50-100 AU, it is possible that these planets can migrate inward as a result of interactions with the disk, after formation. Thus in this paper we use the working hypothesis that massive gaseous planets presently at radial distances $>$ 20 AU could have formed as a result of gravitational instability. Below we investigate the process of heavy element enrichment via planetesimal capture for the planetary system HR 8799. ", "conclusions": "We investigate the possibility of heavy element enrichment for the planetary system HR 8799. We consider planetary masses of 3, 5, 7 and 10 M$_J$, with radial distance of 24, 38, and 68 AU. We find that planetesimal accretion is relatively inefficient for such a system due to main two reasons: First, massive protoplanets have relatively short precollapse stages, and therefore, have limited time for planetesimal capture. Second, the accretion rate is low at large radial distances due to its dependence on the solid surface density and orbital frequency (both decreasing with radial distance). Farther from the star planetesimal velocities are lower, but the accretion time-scale is substantially longer. \\par Smaller planetesimals can be accreted more efficiently, and if planetesimal sizes are of the order of $\\sim$ 1 km, solid material between a few and tens of M$_{\\oplus}$ can be captured. However, if planetesimals are formed with initial sizes of $\\sim$100 km or larger, the accreted mass is negligible. We therefore conclude that under such conditions wide-orbit gaseous planets, if formed via gravitational instability, will have nearly stellar compositions. \\par The capture cross-section used in this work was calculated taking the gravitational enhancement factor $F_g$ which corrects for the three-body effects (see Greenzweig and Lissauer 1990 for details) as unity. As a result, the accreted mass presented here should be taken as a lower bound. The total mass of solids in the protoplanet's feeding zone, which represents the maximum solid mass available for accretion, can be up to 2-3 M$_J$, increasing with radial distance. However, it is unlikely that this available mass can actually be accreted due to the low accretion rate (see Helled and Schubert, 2009 for details) unless it is found that a significant gravitational focusing occurs. Detailed calculations of the gravitational enhancement factor for extended massive protoplanets with gas drag included seem to be required before a robust estimate for the upper bound of the accreted mass can be made. \\par Our results suggest that massive protoplanets at wide orbits will have metallicity similar to their parent star. This conclusion does not depend on the stellar metallicity. Naturally, protoplanets around metal-rich stars will contain more high-Z material in their interiors due to the larger fraction of heavy elements (higher dust-to-gas ratio), but as we show here, an increase in the available solid material typically does not lead to further enrichment due to the low accretion rate. Protoplanets around stars with lower metallicity than considered here, will have a smaller fraction of heavy elements due to their initial composition which is metal-poor, and the fact that planetesimal accretion is inefficient as a result of the long accretion timescale and the low solid surface density. Our suggestion that massive gaseous planets at large radial distances will have stellar compositions is therefore independent of stellar metallicity. \\par Our conclusion that massive protoplanets ($\\ge$ 3 M$_J$) will have relatively low enrichments in heavy elements is valid under the assumption that the high-Z material is captured during the pre-collapse stage in the form of solid planetesimals. It is possible however, that another enrichment mechanism for massive gaseous planets exists. For example, it may be possible that gaseous protoplanets are enriched with heavy elements from 'birth'. Disk fragments form at the corotation of dense spiral waves (Durisen et al. 2008) which can be enhanced with heavier elements (e.g., Haghighipour and Boss 2003, Rice et al. 2004) leading to formation of enriched fragments. In that case, solid concentrations are achieved at the locations where fragmentation is most likely to occur, and the formed protoplanets will be metal-rich relatively to their host star. In addition, simulation of solids in disks show that regions in the disk can become highly enriched with solid material as a result of vortices in the disk (e.g., Klahr and Bodenheimer, 2006), although such vortices are likely to be destroyed in self-gravitating disks (Mamatsashvili and Rice, 2009). In any case, forming gaseous massive protoplanets which are enriched with heavy elements from birth is possible, and as a result massive gaseous protoplanets at large radial distances could contain a significant amount of heavy elements. It would be desirable to investigate whether enrichment from birth can be distinguished from enrichment at later stages (planetary evolution, internal structure, etc.), and we hope to address this issue in future work. An even more important issue that needs to be addressed to make future progress on this problem is a detailed study of planetesimal formation and collisional evolution in the 50--100 AU region of a gravitationally unstable disk. \\par Given the assumption that the planets around HR 8799 formed from near-stellar composition, a question arises as to the observable metallicity in the present atmospheres. Even if there is added planetesimal material it tends to be deposited deep inside the planet, and a substantial fraction of it would condense out and form a solid core (Helled et al., 2008). Also, to a lesser extent, the grains in the original composition of the planet would tend to settle out, and at least the rock component could be added to the core. The heavy element material not in the core would have been mixed by convection before the present time. These considerations suggest a slight depletion in heavy elements in the envelope of the planet as compared with the star. However numerous additional processes, such as cloud formation and possible core erosion, affect the atmosphere between the early phase studied here and the present time. Thus it is difficult to make precise predictions. In the limiting case where the planet has been completely mixed and has uniform composition at the present time, a typical case (38 AU, 10 km planetesimals, 7 M$_J$), the added solid mass is only 2.9 M$_\\oplus$ compared to a total mass of 2226 M$_\\oplus$. Thus even with some enhancement as discussed earlier, the difference between planetary and stellar metallicity would be negligible. Even in the case of the most-enriched planet, the fraction of solid material added is only 4\\% by mass." }, "0911/0911.2623_arXiv.txt": { "abstract": "{The observed spectral variation of HD\\,50138 has led different authors to classify it in a very wide range of spectral types and luminosity classes (from B5 to A0 and III to Ia) and at different evolutionary stages as either HAeBe star or classical Be. } {Based on new high-resolution optical spectroscopic data from 1999 and 2007 associated to a photometric analysis, the aim of this work is to provide a deep spectroscopic description and a new set of parameters for this unclassified southern B[e] star and its interstellar extinction. } {From our high-resolution optical spectroscopic data separated by 8 years, we perform a detailed spectral description, presenting the variations seen and discussing their possible origin. We derive the interstellar extinction to HD\\,50138 by taking the influences of the circumstellar matter in the form of dust and an ionized disk into account. Based on photometric data from the literature and the new Hipparcos distance, we obtain a revised set of parameters for HD\\,50138.} {Because of the spectral changes, we tentatively suggest that a new shell phase could have taken place prior to our observations in 2007. We find a color excess value of E(B-V) = 0.08\\,mag, and from the photometric analysis, we suggest that HD\\,50138 is a B6-7 III-V star. A discussion of the different evolutionary scenarios is also provided. } {} ", "introduction": "Stars that present the B[e] phenomenon are known to form a heterogeneous group. This group is composed of objects at different evolutionary stages, such as high- and low-mass evolved stars, intermediate-mass pre-main sequence stars, and symbiotic objects (Lamers et al. \\cite{Lamers}). However, for more than 50\\% of the confirmed galactic B[e] stars, the evolutionary stage is still unknown, so that they are gathered in the group of the unclassified B[e] stars. This problem is mainly caused by poor knowledge of their physical parameters, especially their distances. In this paper, we present our study related to the southern B[e] star \\object{HD\\,50138} (V743 Mon, MWC158, IRAS 06491-0654). The spectrum of this star was discussed for the first time by Merrill et al. (\\cite{Merrill1}). Later, the presence of spectral variability was cited by Merrill (\\cite{Merrill2}) and Merrill \\& Burwell (\\cite{Merrill3}). Since then, numerous papers have mainly considered this star as either a pre-main sequence star, more specifically as a Herbig Ae/Be star, or as a classical Be star. Jaschek et al. (\\cite{Jaschek2}) have even considered this star as a transition object between a classical Be and a B[e] star. Later Lamers et al. (\\cite{Lamers}) and Zorec (\\cite{Zorec}) include this star in the list of unclassified B[e] stars. The difficulty in obtaining the correct classification of HD\\,50138 is mainly caused by the strong spectral variability as reported in the literature. Pogodin (\\cite{Pogodin}) cites the presence of different scales of spectral variability, from days to months, especially in the line profiles of H$\\alpha$, He{\\sc i} ($\\lambda$ 5876), and Na{\\sc i} lines. Spectral variations have also been identified in the UV, by the analysis of IUE spectra (Hutsem\\'ekers \\cite{Hutsemekers}). These spectral variabilities have been explained by an outburst that possibly happened in 1978-1979 (Hutsem\\'ekers \\cite{Hutsemekers}) and by a shell phase in 1990-1991 (Andrillat \\& Houziaux \\cite{Andrillat}). On the other hand, HD\\,50138 had presented small and non-periodic photometric variations that did not seem to be associated to the spectral changes related to this shell phase (Halbedel \\cite{Halbedel}). In addition, studies based on polarimetry and spectro-polarimetry have identified an intrinsic polarization that seems to be linked to nonspherical symmetry of matter around this object, probably a circumstellar disk (Vaidya et al. \\cite{Vaidya}; Bjorkman et al. \\cite{Bjorkman}; Oudmaijer \\& Drew \\cite{Oudmaijer}; Harrington \\& Kuhn \\cite{Harrington}). A disk-scattering effect associated to an outflow has also recently been suggested by Harrington \\& Kuhn (\\cite{Harrington2}). On the other hand, Cidale et al. (\\cite{Cidale}) propose that this object could be a binary system. Based on spectro-astrometry, Baines et al. (\\cite{Baines}) suggest the same possibility, where the companion would be separated by 0.5$\\arcsec$ - 3.0$\\arcsec$. However, up to now, no direct evidence of binarity has been found. HD\\,50138 was further observed during the Hipparcos mission, and its distance of 290$\\pm$71\\,pc (Perryman et al. \\cite{Perryman}) turned out with 500$\\pm$150\\,pc to be almost doubled according to the new data reduction procedure performed by van Leeuwen (\\cite{vanLeeuwen}). The newly determined distance and the large uncertainties in the previously published stellar classification attempt request and warrant detailed investigation and revision of the stellar parameters of HD\\,50138. In this study, we present our new optical spectroscopic observations and perform a photometric and spectroscopic analysis of HD\\,50138, aimed on the one hand at describing the observed spectral variations and, on the other, at better constraining the stellar parameters, needed for an improved discussion about the nature of this object. The paper has the following structure. In Sect.\\,\\ref{obs} we describe our observations. In Sect.\\,\\ref{results}, we present our results. In Sect.\\,\\ref{spectral_descr}, we describe our high-resolution spectra taken on different dates. In Sect.\\,\\ref{SpType}, we derive from the analysis of our spectroscopic data and public photometric measurements, the insterstellar and circumstellar extinction, hence the stellar parameters and the spectral type of this object. In Sect.\\,\\ref{discussion}, we discuss the possible scenarios for explaining the nature of this curious star, and finally in Sect.\\,\\ref{conclusion}, we present our conclusions. ", "conclusions": "HD\\,50138 is a very curious star that displays strong spectroscopic variations. Based on analysis of new high-resolution data, we present a detailed description of these variations. Our analysis of the photometric data suggests that HD\\,50138 is a B6-7 III-V star, whose luminosity was tentatively obtained from a careful study of the influence of the possible circumstellar extinction sources and based on the new Hipparcos distance. A new value for the color excess, $E(B-V) = 0.08\\pm 0.01$\\,mag, was derived. In addition, we suggest that a new-shell phase or the formation of one-armed spiral could have taken place before 2007. Based on our results, a pre-main sequence star or a transition object between Be and B[e] stars, close to or just at the turn-off from the main sequence, or a binary scenario, can be neither confirmed nor discarded; however, an observational campaign, based on photometry, to derive a detailed light-curve and high-resolution spectroscopy associated to a detailed analysis in terms of the line profile appearances and variations have to be performed to confirm the possible shell phases and the slow-down of this star. A careful interferometric analysis, associated to a SED modeling considering different scenarios for the circumstellar dust is in progress (Borges Fernandes et al., in preparation) and will certainly provide better constraints for the circumstellar geometry and the nature of this curious star." }, "0911/0911.4233_arXiv.txt": { "abstract": "{ We present the results of soft X-ray studies of the classical nova V2491 Cygni using the Suzaku observatory. On day 29 after outburst, a soft X-ray component with a peak at $\\sim$0.5~keV has appeared, which is tantalising evidence for the beginning of the super-soft X-ray emission phase. We show that an absorbed blackbody model can describe the observed spectra, yielding a temperature of 57~eV, neutral hydrogen column density of 2$\\times$10$^{21}$~cm$^{-2}$, and a bolometric luminosity of $\\sim$10$^{36}$~erg~s$^{-1}$. However, at the same time, we also found a good fit with an absorbed thin-thermal plasma model, yielding a temperature of 0.1~keV, neutral hydrogen column density of 4$\\times$10$^{21}$~cm$^{-2}$, and a volume emission measure of $\\sim$10$^{58}$~cm$^{-3}$. Owing to low spectral resolution and low signal-to-noise ratio below 0.6~keV, the statistical parameter uncertainties are large, but the ambiguity of the two very different models demonstrates that the systematic errors are the main point of concern. The thin-thermal plasma model implies that the soft emission originates from optically thin ejecta, while the blackbody model suggests that we are seeing optically thick emission from the white dwarf. } ", "introduction": "Classical novae are a class of cataclysmic variables, which occur in accreting binaries with a white dwarf and a late-type companion. Hydrogen-rich material is accumulated on the white dwarf surface. The surface temperature increases as the material accumulates. When the accreted material reaches a critical mass for nuclear fusion, an thermonuclear runaway process ignites on the white dwarf surface. A review of the evolution of classical novae can be found in e.g., \\cite{starrfield2008}. After outburst, the system is expected to display X-ray emission via different mechanisms: (1) Non-thermal emission has recently been reported from the classical nova V2491 Cygni by the Suzaku satellite \\citep{takei2009a}. (2) Thin-thermal plasma emission from adiabatic shocks has been reported in some classical novae, mainly in data taken with the Swift satellite (e.g., \\citealt{bode2006,ness2009a}). (3) Photospheric emission from a hot layer of a white dwarf surface, which is characterised by bright, soft blackbody-like continuum emission and absorption lines, similar to the class of Super-Soft X-ray Sources (SSS; \\citealt{kahabka1997}). Following from various arguments, the time scale of SSS emission is a function of the white dwarf mass and the chemical composition of the post-outburst envelope (e.g., \\citealt{sala2005,hachisu2006}). The optically thick wind theory is an approach to model SSS light curves with some simple assumptions (e.g., \\citealt{hachisu2005,hachisu2006}). In this theory, the onset of the SSS phase is assumed to occur after the wind stops, and the SSS phase fades out once the nuclear fuel on the white dwarf surface is consumed. Meanwhile, recent observations have shown that significant wind velocities can still be detected during the SSS phase (see contribution by Ness in this same journal). It is difficult to estimate the effects of the expansion on model light curves, and, ultimately, the white dwarf mass that would be derived from models that account for the continued expansion during the SSS phase. More advanced models are clearly needed, nevertheless, the duration of the SSS phase is an important observational quantity that will always be needed to determine the white dwarf mass. In order to determine the duration of the SSS phase, we have to estimate the turn-on and turn-off time of SSS emission. For both quantities, dense observations in time are needed, and intensive monitoring campaigns are currently being performed with Swift (see, e.g., the contributions by Beardmore and by Osborne in this same journal). However, the turn-on time is particularly difficult to be estimated because we need to discriminate SSS emission from other sources of X-ray emission. For example, in RS\\,Oph, significant shock emission was present during the early times and faded on a time scale of $\\sim$30~days. When the SSS phase started around day 30 after outburst, some shock emission was still present that may contaminate the SSS emission (e.g., \\citealt{ness2009b}). \\cite{rohrbach2009} analysed three Chandra ACIS spectra of V1494 Aql obtained on days 134, 187, and 248 after outburst, and interpreted a soft excess in the last observation as SSS emission. However, in the two earlier ACIS spectra, emission lines from N were present at the same energies. They emphasised that the soft component in all three ACIS spectra could equally well be fitted by a blackbody or a thin-thermal plasma model with high N or O abundance. The conclusion of SSS emission in a CCD-type spectrum has thus to be treated with care, because the energy resolution is not high enough to compare these models. \\cite{rohrbach2009} had additional high-resolution Chandra LETGS spectra at their disposal that clearly showed that the soft component was atmospheric rather than thin-thermal emission. For these purposes, it is thus mandatory to observe with high energy resolution with sufficient sensitivity at soft X-ray energies so that different models can be distinguished. In this paper, we present the results of soft X-ray studies of the classical nova V2491 Cygni using the Suzaku observatory. Suzaku has taken two observations during the early phase, and the X-ray CCDs aboard Suzaku provided X-ray spectra of this nova \\citep{takei2009a,takei2009b}. A tantalising hint of SSS emission was seen on day 29, although it provides several different interpretations at the same time. We discuss different possibilities of interpreting the soft component. ", "conclusions": "The fast nova V2491 Cygni has been observed on days 9 and 29 after outburst with Suzaku. The first observation showed a non-thermal component and an optically thin thermal plasma component with a temperature of $\\sim$2.9~keV \\citep{takei2009a}. Later, on day 29, the non-thermal component has disappeared, and the thin-thermal component has a softer spectrum, yielding a lower temperature \\citep{takei2009b}. In addition, a super-soft component with a peak at 0.5~keV was present. The origin of this component can either be first light of the SSS phase, or it could be the soft extension of the thin-thermal plasma component. With the given combination of limited spectral resolution and soft sensitivity, we can not distinguish between these two possibilities. Technically, both scenarios are possible and plausible. Depending on the correct interpretation, the duration of the SSS phase could be different by about a week. According to \\cite{page2009}, the SSS phase started between days 36 and 42. If the soft component detected by Suzaku is indeed SSS emission, then the start of the SSS phase would be at least one week earlier. Central to addressing such questions is the spectral resolution. In high resolution, one could clearly distinguish between SSS and a thin-thermal plasma emission. The results of this work are one example that demonstrates the need for high spectral resolution with sufficient sensitivity at soft energies that would have to be implemented in future X-ray observatories like the International X-ray Observatory." }, "0911/0911.4719_arXiv.txt": { "abstract": "We present spectroscopic and photometric observations of the Type IIn supernova (SN) 2008iy. \\sn\\ showed an unprecedentedly long rise time of $\\sim$400 days, making it the first known SN to take significantly longer than 100 days to reach peak optical luminosity. The peak absolute magnitude of \\sn\\ was $M_r \\approx -19.1$ mag, and the total radiated energy over the first $\\sim$700 days was $\\sim$2 $\\times 10^{50}$ erg. Spectroscopically, \\sn\\ is very similar to the Type IIn SN~1988Z at late times, and, like SN~1988Z, it is a luminous X-ray source (both supernovae had an X-ray luminosity $L_{\\rm X} > 10^{41}$ erg s$^{-1}$). \\sn\\ has a growing near-infrared excess at late times similar to several other SNe~IIn. The H$\\alpha$ emission-line profile of \\sn\\ shows a narrow P Cygni absorption component, implying a pre-SN wind speed of $\\sim$100 \\kms. We argue that the luminosity of \\sn\\ is powered via the interaction of the SN ejecta with a dense, clumpy circumstellar medium. The $\\sim$400 day rise time can be understood if the number density of clumps increases with distance over a radius $\\sim$1.7 $\\times 10^{16}$ cm from the progenitor. This scenario is possible if the progenitor experienced an episodic phase of enhanced mass loss $<$ 1 century prior to explosion or if the progenitor wind speed increased during the decades before core collapse. We favour the former scenario, which is reminiscent of the eruptive mass-loss episodes observed for luminous blue variable (LBV) stars. The progenitor wind speed and increased mass-loss rates serve as further evidence that at least some, and perhaps all, Type IIn supernovae experience LBV-like eruptions shortly before core collapse. We also discuss the host galaxy of \\sn, a subluminous dwarf galaxy, and offer a few reasons why the recent suggestion that unusual, luminous supernovae preferentially occur in dwarf galaxies may be the result of observational biases. ", "introduction": "The recent development of synoptic, wide-field imaging has revealed an unexpected diversity of transient phenomena. One such example is the discovery of a new subclass of very luminous supernovae (VLSNe; e.g., \\citealt{ofek06gy,smith07-2006gy,quimby05ap}). While events of this nature are rare \\citep{miller08es,quimby2009}, each new discovery serves to bracket our understanding of the physical origin of well-established classes of SNe and does, in principle, demand an increased clarity in our understanding of the post-main-sequence evolution of massive stars. Yet another unusual transient was discovered by the Catalina Real-Time Transient Survey (CRTS; \\citealt{drake2009}), which they announced via an ATel as CSS080928:160837+041626 on 2008 Oct.\\ 07 UT\\footnote{UT dates are used throughout this paper unless otherwise noted.} \\citep{drake08-2008iy}. The transient was classified as a Type IIn SN with a spectrum taken on 2009 Mar.\\ 27 (\\citealt{mahabal2009}; see \\citealt{schlegel96} for a definition of the SN~IIn subclass, and \\citealt{filippenko1997} for a review of the spectral properties of SNe). The SN was later given the IAU designation \\sn\\ \\citep{catelan2009}. \\citet{mahabal2009} noted that the transient was present on CRTS images dating back to 2007 Sep.\\ 13; however, it went undetected by the CRTS automated transient detection software until 2008 because it was blended with a non-saturated, nearby ($\\sim$11\\arcsec\\ separation) star. Here, we present our observations and analysis of \\sn, which peaked around 2008 Oct.\\ 29 \\citep{catelan2009} and had a rise time of $\\sim$400 days. This implies that \\sn\\ took longer to reach peak optical brightness than any other known SN. Type II SN rise times are typically $\\la$ 1 week (e.g., SNe~2004et and 2006bp, \\citealt{wli-2004et, quimby06bp}; see also \\citealt{patat1993,wli-rates-LF}), and have never previously been observed to rise $\\ga$ 100 days, let alone 400, making \\sn\\ another rare example of the possible outcomes for the end of the stellar life cycle. In addition to an extreme rise time, \\sn\\ is of great interest because the unique circumstellar medium (CSM) in which it exploded may provide a link to very long-lived SNe, such as SN 1988Z, and thus provide clues into the nature of their progenitors. This paper is organised as follows. Section 2 presents the observations. The data are analysed in Section 3, and the results are discussed in Section 4. We give our conclusions, as well as predictions for the future behaviour of \\sn, in Section 5. ", "conclusions": "We have reported on observations of the Type IIn \\sn, which took $\\sim$400 days to reach peak optical output. There are few known SNe with optical rise times $\\ga$ 50 days, and \\sn\\ is the first with a rise time $>$ 1 year. We argue that this long rise to peak is caused by the interaction of the SN ejecta with a dense CSM; radioactivity is unlikely to drive a 400-day rise. Furthermore, the late-time optical decay, $\\sim$0.4 mag (100 days)$^{-1}$, is slower than that of $^{56}$Co, which provides further evidence that radioactive heating is not a dominant energy source for \\sn. Spectroscopically, \\sn\\ is very similar to SN~1988Z at late times. SN~1988Z is understood to have exploded in a dense, clumpy CSM \\citep{chugai1994}, which we argue was also the case for \\sn. We detect \\sn\\ in X-rays with a total luminosity $L_{\\rm X} = (3.7 \\pm 1.2) \\times 10^{41}$ erg s$^{-1}$, which is similar to that of the Type IIn SNe~1988Z and 1995N \\citep{fox2000}. Similar to other SNe~IIn, \\sn\\ has a growing NIR-excess at late times. \\sn\\ had a peak absolute magnitude of $M_r \\approx -19.1$ mag and a total radiated energy of $\\sim 2 \\times 10^{50}$ erg in the optical (assuming no bolometric correction). The steady increase in optical luminosity over a $\\sim$400-day period means that the wind-density parameter, $w$, increased over a distance of $\\sim$1.7 $\\times 10^{16}$ cm from the SN. We propose two possible scenarios to explain this increase in $w$: (i) the progenitor experienced an episode of LBV-like, eruptive mass-loss $\\sim$55 years prior to the SN, or (ii) the wind speed of the progenitor was increasing during the years leading up to core collapse. We prefer the former scenario, as the later adds unnecessarily complicated pieces to the puzzle without providing a unique solution. Our favoured scenario provides yet another piece of evidence that some SNe~IIn are connected to LBV-like progenitors (see \\citealt{gal-yam2007}, and references therein). We find that the host of \\sn\\ is a subluminous dwarf galaxy, though we caution against premature conclusions that unusual SNe, specifically those that are very luminous or have very long rise times, preferentially occur in low-mass dwarf galaxies. Finally, we close with some predictions for the late-time behaviour of \\sn. There are examples of SNe~IIn whose luminosity dramatically drops after the SN ejecta overtake the dense CSM (e.g., SN~1994W; \\citealt{sollerman94w}). However, the similarities to SN~1988Z suggest that \\sn\\ could continue interacting, and thus remain luminous, for several years. This would allow long-term monitoring in the radio, X-ray, and optical, like SN~1988Z \\citep{williams2002, schlegel2006, aretxaga88z}. We predict that \\sn\\ is a luminous radio source, like SN~1988Z \\citep{vandyk88z}, though we note that at $z = 0.0411$ deep observations may be necessary for detection. The increasing $K_s$-band luminosity from day 560 to 710 is likely due to the presence of dust, and we predict that \\sn\\ will also be luminous in the mid-IR ($\\sim$3$-$5 $\\mu$m). The data are insufficient to distinguish between newly formed dust or dust that was present prior to the SN, but future medium- to high-resolution spectroscopy could distinguish between these two cases. If new dust is being formed, it should result in a systematic blueshift in the line profiles, as the dust creates an optically thick barrier to radiation from the receding SN ejecta. \\bigskip A.A.M. would like to thank D. Poznanski for useful discussions that helped improve this paper and M. Modjaz for discussions concerning the metallicity of the host. We thank Neil Gehrels for approving the {\\it Swift} ToO request for \\sn, which was submitted by Dave Pooley, and the {\\it Swift} team for scheduling and obtaining those observations. We acknowledge the use of public data from the {\\it Swift} archive. Cullen Blake, Dan Starr, and Emilo Falco assisted with the operation of PAIRITEL. We wish to thank the following Nickel observers: P. Thrasher, M. Kislak, J. Rex, J. Choi, I. Kleiser, J. Kong, M. Kandrashoff, and A. Morton. We thank the referee, Nikolai Chugai, for suggestions that helped to improve this paper. A.A.M. is supported by the NSF Graduate Research Fellowship Program and DOE grant \\#DE-FC02-06ER41453. J.S.B.'s group is partially supported by NASA/{\\it Swift} grant \\#NNX08AN84G. A.V.F.'s group is grateful for the financial assistance of NSF grant AST--0908886 and the TABASGO Foundation. PAIRITEL is operated by the Smithsonian Astrophysical Observatory (SAO) and was made possible by a grant from the Harvard University Milton Fund, the camera loan from the University of Virginia, and the continued support of the SAO and UC Berkeley. The PAIRITEL project is partially supported by NASA/{\\it Swift} Guest Investigator Grant \\#NNX08AN84G. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration (NASA) and the National Science Foundation (NSF). Some of the data presented herein were obtained at the W. M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California, and NASA; it was made possible by the generous financial support of the W. M. Keck Foundation. The authors wish to recognise and acknowledge the very significant cultural role and reverence that the summit of Mauna Kea has always had within the indigenous Hawaiian community; we are most fortunate to have the opportunity to conduct observations from this mountain." }, "0911/0911.3763_arXiv.txt": { "abstract": "Radial metallicity gradients are observed in the disks of the Milky Way and in several other spiral galaxies. In the case of the Milky Way, many objects can be used to determine the gradients, such as HII regions, B stars, Cepheids, open clusters and planetary nebulae. Several elements can be studied, such as oxygen, sulphur, neon, and argon in photoionized nebulae, and iron and other elements in cepheids, open clusters and stars. As a consequence, the number of observational characteristics inferred from the study of abundance gradients is very large, so that in the past few years they have become one of the main observational constraints of chemical evolution models. In this paper, we present some recent observational evidences of abundance gradients based on several classes of objects. We will focus on (i) the magnitude of the gradients, (ii) the space variations, and (iii) the evidences of a time variation of the abundance gradients. Some comments on recent theoretical models are also given, in an effort to highlight their predictions concerning abundance gradients and their variations. ", "introduction": "Radial metallicity gradients are observed in the disks of many galaxies, including the Milky Way, galaxies of the Local Group and other objects. Current research topics include the determination of (i) the magnitude of the gradients, (ii) any space variations along the disk and (iii) possible time variations during the evolution of the host galaxy. In this paper, we present some recent observational evidences of radial abundance gradients. Our main focus is the Milky Way, but it will be shown that the analysis of some objects in the Local Group, particularly M33, is useful in order to study the main properties of the gradients in our own Galaxy. A brief discussion of some recent theoretical models is also given, in an effort to highlight their predictions concerning the radial abundance gradients and their variations. Some recent reviews and general papers on abundance gradients include: Freeman (\\cite{freeman}), Maciel \\& Costa (\\cite{mc2009}), Rudolph et al. (\\cite{rudolph}), and Stasi\\'nska (\\cite{stasinska04}). Theoretical models are discussed by a number of people, including Fu et al. (\\cite{fu}), Magrini et al. (\\cite{magrini09a}), Cescutti et al. (\\cite{cescutti}), Moll\\'a and D\\'\\i az (\\cite{molla}), Chiappini et al. (\\cite{chiappini03}, \\cite{chiappini01}, \\cite{chiappini97}), Hou et al. (\\cite{hou}), and Gensler (these proceedings). Recent discussions on azimuthal and vertical gradients, not treated here, can be found in Davies et al. (\\cite{davies}) and Ivezi\\'c et al. (\\cite{ivezic}). ", "conclusions": "From all objects considered, some tentative conclusions may be drawn: (1) Average abundance gradients are generally between $-0.03$ dex/kpc and $-0.10$ dex/kpc, but a single value for the whole disk may be misleading. (2) Most evidences point to a flattening out of the gradients at large galactocentric distances. (3) There are some clear evidences of a flattening of the gradients near the galactic bulge. (4) The change of slope in the outer Galaxy occurs in the region around $R \\simeq 10$ kpc. (5) Any further change of the slope needs better data than presently available. Cepheids may be an exception. (6) There are no evidences of a steepening of the gradients at large galactocentric distances, as suggested by some theoretical models. (7) Either the gradients do not change appreciably during galactic evolution, or they flatten out at a moderate rate. (8) There are no clear evidences of a steepening of the gradients with time, as suggested by some theoretical models. \\bigskip\\noindent {\\it Acknowledgements. This work was partially supported by FAPESP and CNPq.}" }, "0911/0911.4972_arXiv.txt": { "abstract": "Weak gravitational lensing has been used extensively in the past decade to constrain the masses of galaxy clusters, and is the most promising observational technique for providing the mass calibration necessary for precision cosmology with clusters. There are several challenges in estimating cluster masses, particularly (a) the sensitivity to astrophysical effects and observational systematics that modify the signal relative to the theoretical expectations, and (b) biases that can arise due to assumptions in the mass estimation method, such as the assumed radial profile of the cluster. All of these challenges are more problematic in the inner regions of the cluster, suggesting that their influence would ideally be suppressed for the purpose of mass estimation. However, at any given radius the differential surface density measured by lensing is sensitive to all mass within that radius, and the corrupted signal from the inner parts is spread out to all scales. We develop a new statistic \\ups\\ that is ideal for estimation of cluster masses because it completely eliminates mass contributions below a chosen scale (which we suggest should be about 20 per cent of the virial radius), and thus reduces sensitivity to systematic and astrophysical effects. We use simulated and analytical profiles including shape noise to quantify systematic biases on the estimated masses for several standard methods of mass estimation, finding that these can lead to significant mass biases that range from ten to over fifty per cent. The mass uncertainties when using the new statistic \\ups\\ are reduced by up to a factor of ten relative to the standard methods, while only moderately increasing the statistical errors. This new method of mass estimation will enable a higher level of precision in future science work with weak lensing mass estimates for galaxy clusters. % ", "introduction": "\\label{S:introduction} Many scientific applications require robust measurements of the mass in galaxy clusters. One such application is the use of the dark matter halo mass function to constrain cosmological model parameters, including the amplitude of matter density perturbations, the average matter density, and even the equation of state of dark energy \\citep[e.g., most recently, ][]{2007ApJ...657..183R,2008MNRAS.387.1179M,2009ApJ...692.1060V,2010ApJ...708..645R}. Another example is validation and refinement of models of cluster formation and evolution, which predict relations between the more easily measured optical and X-ray emission, and the underlying dark matter halo \\citep{2006ApJ...650..128K,2007ApJ...668....1N,2008A&A...482..451Z,2009arXiv0906.4370B}. Currently, there are thousands of known clusters selected in various ways that can be used for these applications. Future surveys such as the Dark Energy Survey (DES)\\footnote{\\texttt{http://www.darkenergysurvey.org/}}, Pan-STARRS\\footnote{\\texttt{http://pan-starrs.ifa.hawaii.edu/public/}}, and the Large Synoptic Survey Telescope (LSST)\\footnote{\\texttt{http://www.lsst.org/lsst}} will provide even larger and deeper samples that can be used for this purpose, requiring greater systematic robustness in the mass measures to complement the smaller statistical errors. Many different methods have been used to measure the halo profile of clusters and thereby estimate their masses. Kinematic tracers such as satellite galaxies, in combination with a Jeans analysis or caustics analysis, can give information over a wide range of physical scales and halo masses. While the issues of relaxation, velocity bias, anisotropy of the orbits and interlopers need to be carefully addressed, recent results suggest a good agreement with theoretical predictions for the form of the density profile \\citep{2003ApJ...585..205B,2004ApJ...600..657K,2003AJ....126.2152R,2005ApJ...628L..97D,2006AJ....132.1275R,2007MNRAS.378...41S}. Hydrostatic analyses of X-ray intensity profiles of clusters % use X-ray intensity and temperature as a function of radius to reconstruct the density profile and estimate a halo mass. The advantage of thermal gas pressure being isotropic in partially lost due to the possible presence of other sources of pressure support, such as turbulence, cosmic rays or magnetic fields. These extra sources of pressure support cannot be strongly constrained for typical clusters with present X-ray data \\citep{2004A&A...426..387S}, but could modify the hydrostatic equilibrium and affect the conclusions of such analyses. Recent results are encouraging and are in a broad agreement with predictions, although most require concentrations that are higher than those predicted by a concordance cosmology \\citep{2006ApJ...640..691V,2007ApJ...664..123B,2007MNRAS.379..209S}. While the above-mentioned systematic biases cannot be excluded, the small discrepancy could also be due to baryonic effects in the central regions, due to selection of relaxed clusters that may be more concentrated than average \\citep{2006ApJ...640..691V}, or due to the fact that at a given X-ray flux limit, the more concentrated clusters near the limiting mass are more likely to be included in the sample \\citep{2007A&A...473..715F}. Gravitational lensing is by definition sensitive to the total mass, and is therefore one of the most promising methods to measure the mass profile independent of the dynamical state of the clusters. Many previous weak lensing analyses have focused on individual clusters (for example, \\citealt{2007MNRAS.379..317H,2007ApJ...667...26P,2009ApJ...702..603A,2009arXiv0903.1103O}). Measuring the matter distribution of individual clusters allows a comparison with the combined baryonic (light and gas) distribution on an individual basis, and so can constrain models that relate the two, such as MOND versus CDM \\citep{2006ApJ...648L.109C}. However, these measurements can be quite noisy for individual clusters. Stacking the signal from many clusters can ameliorate this problem, since shape noise and the signal due to correlated structures will be averaged out. Such a statistical approach is thus advantageous if one is to compare the observations to theoretical predictions, which also average over a large number of halos in simulations. A final advantage of stacking is that it allows for the lensing measurement of lower-mass halos, where individual detection is impossible due to their lower shears relative to more massive clusters. Individual high signal-to-noise cluster observations and those based on stacked analysis of many clusters are thus complementary to each other at the high mass end, with the stacked analysis drastically increasing the available baseline in mass. Extraction of cluster dark matter halo masses from the weak lensing signal is subject to a number of uncertainties, which we discuss in this paper in detail, including the ways that the uncertainties differ for individual versus stacked cluster lensing analyses. In brief, these uncertainties are: (i) biased calibration of the lensing signal; (ii) modification of the lensing profile in the inner cluster regions due to accidental inclusion of cluster member galaxies in the source sample, intrinsic alignments of those galaxies, non-weak shear, magnification, baryonic effects that modify the initial cluster dark matter halo density profile, and cluster centroiding errors; (iii) contributions to the lensing signal from nonvirialized local structures and large-scale structure (LSS). Furthermore, parametric modeling of the mass requires the assumption of a form for the dark matter halo profile, which may differ from the intrinsic profile and/or have poorly constrained parameters. Non-parametric modeling, while not subject to this weakness, results in projected masses that must be converted to three-dimensional enclosed masses to be compared against the theory predictions, all of which are currently phrased in terms of 3d masses. We quantify the degree to which this conversion depends on assumptions about the density profile. Generally, we show the effects of many of these uncertainties on the estimated masses from cluster weak lensing analyses, both in the stacked and individual cases, using parametric and non-parametric mass modeling. Effects that modify the cluster density profile in the inner regions ($\\lesssim 0.5$\\hmpc), are particularly problematic given that the weak lensing signal \\dsr\\ is sensitive to the density profile not just at a projected separation $R$, but also at all smaller separations. We propose a modified statistic, denoted \\ups, that removes the dependence on the projected density between $R=0$ and $R=R_0$, with $R_0$ chosen to avoid scales with systematic uncertainties. The decrease in systematic errors that results from removing scales below $R_0$ comes at the expense of somewhat increased statistical errors. We explore the optimal choice of $R_0$, and quantify the degree to which our use of this new statistic to estimate cluster masses lessens systematic biases and increases statistical errors. Our tools for this investigation include simple, idealised cluster density profiles; more complex and realistic density profiles from $N$-body simulations; and finally, real cluster lensing data from the Sloan Digital Sky Survey (SDSS, \\citealt{2000AJ....120.1579Y}) that was previously analysed by \\cite{2008JCAP...08..006M}. We begin in Section~\\ref{S:theory} with a discussion of the theoretical aspects of cluster-galaxy weak lensing, including a detailed discussion of the challenges of mass determination, and a summary of typical approaches to parametric and non-parametric mass estimation, with the introduction of a new statistic from which to derive parametric mass estimates. In Section~\\ref{S:simulations}, we describe the $N$-body simulations that we use to provide sample cluster density profiles. Section~\\ref{S:data} has a description of the SDSS cluster lensing data we use to test for some of the effects that we find using the simulations. Results for both the theoretical profiles and the real data are presented in Section~\\ref{S:results}. We conclude with a discussion of our findings and their implications in Section~\\ref{S:conclusions}. ", "conclusions": "\\label{S:conclusions} In this paper, we have assessed the degree to which certain systematic errors in lensing measurement and methods of mass estimation can bias weak lensing cluster mass estimates. In brief, the challenges we considered included the following. \\begin{itemize} \\item Lensing calibration bias, which leads to changes in the mass $\\propto \\ds^{\\eta}$ for $\\eta$ typically in the range $1.4$--$2$ depending on the radial range and fit method used for the parametric NFW mass fits (lower for \\dsr\\ than for \\ups), or $\\propto \\ds$ for the non-parametric mass estimates within a fixed physical aperture (or a steeper scaling when estimated the mass within some spherical over-density radius) using \\zetac\\ (Section~\\ref{SSS:calibration},~\\ref{SSS:resnonpara}, and~\\ref{SSS:ressimpara}). \\item Offsets of the identified BCG from the minimum of the cluster potential well (Section~\\ref{SSS:offsets}) were incorporated into the lensing profiles using a model from mock catalogues presented in \\cite{2007arXiv0709.1159J}. This model includes the effects of photometric errors in selecting the wrong BCG, and is therefore an overly conservative estimate in cases where the BCG can be unambiguously identified for all clusters or where X-ray data can precisely locate the cluster centre. \\item The effect of differences between an assumed NFW concentration and the true NFW concentration were studied using pure NFW lensing signals. \\item Differences in the halo profile relative to a pure NFW profile were studied using fits to density profiles from $N$-body simulations. \\item The effects of mass from structures other than the cluster itself on the lensing signal were also studied using the signal from simulations, since we have not included only the mass that is virialized when computing \\ds\\ in the simulations. \\item Contamination of the source sample by cluster member galaxies, intrinsic alignments of those member galaxies, and baryonic effects on the halo density profile were considered to be included among the previous tests, namely changes in NFW concentration (in Section~\\ref{SS:theorresults}), changes in the inner region of the profile using variations of the $N$-body simulation outputs (in Section~\\ref{SS:simresults}), and centroid offsets that modify the signal only in the inner regions of the cluster. \\end{itemize} When fitting a parametric model (in our case, the NFW profile) to \\dsr, with fixed concentration, we find that the uncertainties due to unknown true concentration plus changes in the lensing profile due to small-scale systematics yield systematic errors that range from a factor of two in mass (when only using small scales in the fits, e.g. $0.1$--$1$\\hmpc) to tens of per cent (when using $R>0.5$\\hmpc) to several per cent (for $R>2$\\hmpc, which yields stable mass estimates but large statistical errors, and which may not be available for individual cluster lensing analyses due to limited telescope FOV). This level of systematic error occurred when allowing a relatively broad variation in concentration ($4<\\cvir <7$), given the disagreement between simulations on the concentration-mass relation at high masses, the large lognormal scatter in this relation, and other systematics such as baryonic effects discussed in Sections~\\ref{SS:challenges-theory}, \\ref{SS:challenges-obs}, and~\\ref{SS:paramodel}. When using a narrower range in concentration, the systematic errors decreased comparably, but are still unacceptably large relative to what is needed for precise cosmological parameter constraints. The addition of centroiding errors to the list of systematics we considered led to uniform suppression of the mass estimates of order tens of per cent (for $\\rmin=0.1$\\hmpc). To completely avoid this suppression while fitting to \\dsr\\ and ignoring the possibility of centroiding errors, we found it necessary to restrict the fits to $\\rmin>1$\\hmpc. Generally, the addition of larger scales, out to $\\sim 2$\\rvir, is useful in minimising the effects of small-scale systematics; going beyond that can lead to excessive contribution from large-scale structure, which will bias the mass estimates if it is not modelled accurately. Allowing a variation in concentration in the fits is another way to reduce systematic error, at the expense of statistical errors that are increased by 45 per cent, but this scheme is not helpful when dealing with systematics that have a radial profile that does not mimic a change in concentration. \\ups\\ is still more reliable at removing the impact of small-scale systematics on the lensing signal. The aperture mass statistic \\zetac\\ led to accurate estimates of projected masses, provided that either (a) the mass in the outer annulus was estimated rather than ignored, or (b) the mass in the outer annulus was ignored, but $R_{o1}\\gg R_1$ (i.e. a large range of transverse separations was included in the first integral in Eq.~\\eqref{E:zetac}). For many applications, such as the halo mass function, the quantity of interest is the 3d virial mass, for which a density profile must be assumed to do the conversion from the 2d projected mass within $R_1$. We found that uncertainty in the true density profile led to tens of per cent level biases in the 3d virial masses. The effect of centroiding errors was to uniformly suppress the aperture masses by $\\sim 10$--$20$ per cent depending on the halo mass, degree of centroiding errors, and transverse separations used for the analysis; these biases were then propagated into the 3d enclosed mass estimates. The aperture mass-based estimates of the cluster virial mass were substantially noisier than fits to \\dsr\\ using the same range of scales. Finally, the new statistic we introduce here, \\ups, removes the effect of small scales from the lensing signal, gave superior performance over \\dsr\\ when fitting an NFW profile to the cluster lensing signal. This statement is true not just for the basic tests with pure NFW profiles and profiles from simulations, but also when including the effects of centroiding errors. The increases in statistical error on the mass can be $\\sim 40$ per cent relative to fitting to \\dsr\\ over the same scales. % The residual systematic uncertainties after removal of an overall offset in the masses is of order several per cent, when fitting from $0.5}0.5$ \\citep{Gladders02,Inada03,Borys04,Sharon05,Swinbank07,Ofek08}. The Massive Cluster Survey \\citep[MACS;][]{Ebeling01} offers an unprecedented opportunity to expand the available sample of strong lensing clusters at $0.3\\le z\\ls0.7$. We present new results from this search: MACS\\,J1149.5$+$2223 (hereafter MACS\\,J1149; 11:49:34.3 $+$22:23:42.5 [J2000]) at $z=0.544$, one of a complete subsample of 12 MACS clusters at $z>0.5$ \\citep{Ebeling07}. In \\S\\ref{sec:obs} we describe the data; modeling and results are presented in \\S\\ref{sec:results}, and summarized in \\S\\ref{sec:summary}. We assume $\\Ho=70\\kms\\Mpc^{-1}$, $\\oM=0.3$ and $\\oL=0.7$; at $z=0.544$ $1''$ corresponds to $6.35\\kpc$. All uncertainties and upper/lower limits are stated and/or plotted at $95\\%$ confidence. ", "conclusions": "We have presented new \\emph{HST}/ACS and Keck I/LRIS observations of MACS\\,J1149, a massive X-ray selected galaxy cluster at $z=0.544$ discovered in the Massive Cluster Survey. These data reveal seven robustly identified multiply-imaged galaxies, three of which we have confirmed spectroscopically. The most spectacular system is a multiply-imaged face-on disk galaxy at $z=1.491$ that we identify as an $L^\\star$ ($M_B\\simeq-20.7$) late-type ($B/T\\ls0.5$) galaxy with an ongoing star formation rate of $\\sim6\\Msolpyr$; the brightest images of this galaxy are magnified by $\\mu=23$. Future observations using integral field spectrographs should probe its properties in exquisite detail, thanks to the combination of lens magnification and fortuitous viewing angle. We use the positions and redshifts of robustly identified multiply-imaged galaxies to constrain a detailed model of the mass and structure of the cluster core. Our fiducial model contains the main cluster halo plus three group-scale halos; the probability of a model this complex, relative to less complex models is ${\\rm P}(N_{\\rm halo}=4)/{\\rm P}(N_{\\rm halo}<4)\\ge10^{12}$ where $N_{\\rm halo}$ is the number of cluster/group-scale halos. We measure the mass and fraction of mass residing in substructures to be $M(\\le500\\kpc)=6.7\\pm0.4\\times10^{14}\\Msol$ and $\\fsub(\\le500\\kpc)=0.25\\pm0.12$ respectively. In summary, MACS\\,J1149 is the most complex strong-lensing cluster core studied to date, its relatively dis-assembled nature being qualitatively consistent with the expectation that clusters at high redshifts are on average less mature than those at lower redshifts. A more complete view will emerge from our analysis of the full sample of MACS clusters at $z>0.5$ (Smith et al., in prep.). We also obtain a power law density profile slope of $\\gamma=-0.3\\pm0.05$ ($95\\%$ confidence error bars) on scales of $r\\sim3-30\\arcsec$, thereby ruling out density profile slopes as flat as those recently proposed by \\cite{Zitrin09b} at $\\gs7\\sigma$ confidence. In summary, \\citeauthor{Zitrin09b}'s result can be explained by an absence of multiple-image redshifts of any form in their study, and by them not treating the unknown redshifts as free parameters in their model. These issues are probably compounded by them mis-identifying some multiple-image systems. Overall, this underlines the critical importance of measuring spectroscopic redshifts of multiply-imaged galaxies for reliable lens models of strong lensing clusters." }, "0911/0911.3143_arXiv.txt": { "abstract": "{The discovery of \\epsba, Bb, a binary brown dwarf system very close to the Sun, makes possible a concerted campaign to characterise the physical parameters of two T dwarfs. Recent observations suggest substellar atmospheric and evolutionary models may be inconsistent with observations, but there have been few conclusive tests to date. We therefore aim to characterise these benchmark brown dwarfs to place constraints on such models. We have obtained high angular resolution optical, near-infrared, and thermal-infrared imaging and medium-resolution (up to R$\\sim$5\\,000) spectroscopy of \\epsba, Bb with the ESO VLT and present $VRIzJHKL'M'$ broad-band photometry and 0.63--5.1\\micron\\ spectroscopy of the individual components. The photometry and spectroscopy of the two partially blended sources were extracted with a custom algorithm. Furthermore, we use deep AO-imaging to place upper limits on the (model-dependent) mass of any further system members. We derive luminosities of log L/L$_{\\sun}$=$-4.699\\pm0.017$ and $-5.232\\pm0.020$ for \\epsba, Bb, respectively, and using the dynamical system mass and COND03 evolutionary models predict a system age of 3.7--4.3\\,Gyr, in excess of previous estimates and recent predictions from observations of these brown dwarfs. Moreover, the effective temperatures of 1352--1385\\,K and 976--1011\\,K predicted from the COND03 evolutionary models, for \\epsba\\ and Bb respectively, are in disagreement with those derived from the comparison of our data with the BT-Settl atmospheric models where we find effective temperatures of 1300--1340\\,K and 880--940\\,K, for \\epsba\\ and Bb respectively, with surface gravities of log g=5.25 and 5.50. Finally, we show that spectroscopically determined effective temperatures and surface gravities for ultra-cool dwarfs can lead to underestimated masses even where precise luminosity constraints are available.\\thanks{The full resolution spectra of both brown dwarfs are available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr}} ", "introduction": "\\label{sec:intro} The characterisation of low-mass stars and brown dwarfs is important for studies of substellar and planetary atmospheres, the reliable application of low-mass evolutionary models, and the derivation of the full initial mass function. With over five hundred L and over one hundred T dwarfs now known\\footnote{http://dwarfarchives.org - the M, L, and T dwarf compendium maintained by Chris Gelino, Davy Kirkpatrick, and Adam Burgasser.}, statistical studies of global properties and detailed studies of the closest objects are now possible. Binary systems have an important role to play. They allow the determination of dynamical masses, provide a laboratory in which objects with the same age and chemical composition may be compared, and, where they have main-sequence companions, provide external constraints of metallicity and age which isolated objects lack, breaking the substellar mass-luminosity-age degeneracy. To fully constrain the evolutionary models of substellar objects \\citep[e.g.][]{Burrows:1997, Baraffe:2003, Saumon:2008}, it would be most useful to determine the bolometric luminosity, radius, mass, and age of a range of such objects. Bolometric luminosities can be determined from photometric and spectroscopic observations across a large wavelength range. Masses can be determined in systems where an orbit may be monitored, and finally, the age and metallicity can be inferred from better characterised stars in the same system. To constrain the atmospheric models of brown dwarfs, it is necessary to acquire high signal-to-noise spectra over as wide a wavelength range as possible, allowing robust estimates of the effective temperature and surface gravity to be made. The discovery of a distant companion (projected separation $\\sim$1500\\,AU) to the high proper-motion ($\\sim$4.7\\,arcsec/yr) K4.5V star, $\\varepsilon$ Indi, was reported by \\citet{Scholz:2003}. One of our nearest neighbours, \\epsa\\ has a well-constrained parallax from HIPPARCOS \\citep{Perryman:1997} as refined by \\citet{vanleeuwen:2007}, putting the system at a distance of 3.6224$\\pm$0.0037\\,pc. This was followed by the discovery of the companion's binary nature \\citep{McCaughrean:2004}. The proximity of \\epsba, Bb to the Earth means Ba is more than a magnitude brighter than any other known T dwarf, and allows unprecedented, detailed spectroscopic studies of these important template objects. \\epsba, Bb are uniquely suited to provide key insights into the physics, chemistry, and evolution of substellar sources. Although there are a number of other T dwarfs in binary systems, such as the M4/T8.5 binary Wolf\\,940 \\citep{Burningham:2009} and the T5/T5.5 binary 2MASS\\,1534$-$2952 \\citep{Liu:2008}, \\epsba, Bb has a very well-determined distance, a main-sequence primary star with which to constrain age and metallicity, and a short enough orbit \\citep[nominally $\\sim$15\\,years, ][]{McCaughrean:2004} such that the system and individual dynamical masses can soon be determined \\citep[][in prep.]{McCaughrean:2009, Cardoso:2010}. They are also relatively bright, close enough, and sufficiently separated to allow detailed photometric and spectroscopic studies of both components. Importantly, these two objects roughly straddle the L to T transition \\citep[cf.][]{Burgasser:2009} where the atmospheres of substellar objects alter dramatically. The study of these two coeval objects on either side of the transition will help in understanding the processes effecting the change from cloudy to cloud-free atmospheres. Characterisation of this system allows the mass-luminosity-age relation at low masses and intermediate age to be tested, investigation of the atmospheric chemistry, including vertical up-mixing, and detailed investigation of the species in the atmosphere. To date, spectroscopic observations of T dwarfs have predominantly been either at low-resolution \\citep[e.g., ][]{Burgasser:2002,Chiu:2006}, which allows spectral classification and overall spectral energy distribution modelling to determine luminosities, or high-resolution studies of relatively small wavelength regions to investigate gravity and effective temperature-sensitive features. For example, \\citet{McLean:2003} presented near-IR spectra at a spectral resolution of R$\\sim$2000 of objects spanning spectral types M6 to T8, and discussed broad changes in spectral morphology and dominant absorbers through the spectral sequence. This was complemented by R$\\sim$20\\,000 $J$-band spectra presented in \\citet{McLean:2007} where many H$_2$O and FeH features were identified and the progression of the $J$-band potassium doublet from M to T dwarfs charted. Previous studies of ultra-cool dwarfs attempting to constrain low-mass evolutionary models have been hampered by ambiguous ages, possible unresolved binarity, and the difficulty associated with acquiring observations of close, faint companions. For example, observations of AB\\,Dor\\,C have roused some controversy over the applicability of current low-mass evolutionary models, with the assumed age of the system being a major source of disagreement. \\citet{Close:2005} determined a dynamical mass for AB\\,Dor\\,C and, using an assumed age of 30--100\\,Myr, argued that evolutionary models predicted a higher luminosity than was observed. However, \\citet{Luhman:2005} countered with an analysis based on an age of 75--150\\,Myr, finding no significant discrepancy between the observations and models. \\citet{Nielsen:2005} further argued that using their slightly revised age of 70$\\pm$30\\,Myr, the models still under-estimated the mass of this object. Again, this was disputed by \\citet{Luhman:2006} after a re-reduction of the same data used by \\citet{Close:2005}, and then \\citet{Close:2007} concluded that, based on newly acquired spectra, there was no discrepancy between the observations and models. Despite this apparent rapprochement, the situation may nevertheless be further complicated by the suggestion of \\citet{Marois:2005} that AB Dor C may itself be an unresolved binary. More recently, \\citet{Dupuy:2009} presented a dynamical mass for the binary L dwarf system HD\\,130948BC which, along with an age estimated from the rotation of the main-sequence parent star, suggests that the evolutionary models predict luminosities 2--3 times higher than those observed. \\citet{Leggett:2008} used 0.8--4.0\\micron\\ spectra at R$\\sim$100--460 and near- to mid-IR photometry of HN Peg B, a T2.5 dwarf companion to a nearby G0V star, to investigate physical properties including dust grain properties and vertical mixing. In the near-IR, the resolution was too low to study spectral lines in detail. However, by fitting the overall spectral morphology and making use of the longer-wavelength data, they were able to place important constraints on vertical mixing and sedimentation. \\citet{Leggett:2009} also reported the physical properties of four T8--9 dwarfs from fitting observed near- to mid-IR spectral energy distributions with the atmospheric and evolutionary models of \\citet{Saumon:2008}. They discussed the effects of vertical transport of CO and N$_2$ and demonstrated the complementary effects of increasing metallicity and surface gravity. \\citet{Reiners:2007} analysed high-resolution (R$\\sim$33\\,000) optical spectra of three L dwarfs and the combined \\epsba, Bb system, concluding that although some individual features are not well-matched by the model atmospheres and that significant differences remain for some molecular species and alkali metal features, general features are reproduced. \\citet{Smith:2003} also acquired high resolution (R$\\sim$50\\,000) near-IR spectra of (only) \\epsba\\ in the wavelength ranges 1.553--1.559\\micron\\ and 2.308--2.317\\micron. These were fit with the unified cloud models of \\citet{Tsuji:2002} and effective temperatures of 1400\\,K and 1600\\,K were derived for the two spectral regions. Mid-IR spectroscopy of the unresolved \\epsba, Bb system was also acquired by \\citet{Roellig:2004}, who use evolutionary models along with the luminosities of \\citet{McCaughrean:2004} and an assumed age of 0.8--2.0\\,Gyr to derive effective temperatures and surface gravities and then compared composite spectral models to the observed spectrum. Their predictions were revised by \\citet{Mainzer:2007} who derived effective temperatures of 1210--1250\\,K and 840\\,K, for \\epsba\\ and Bb respectively, under the assumption of a system age of $\\sim$1\\,Gyr. Finally, \\citet{Kasper:2009} presented R$\\sim$400 near-IR NACO/VLT spectroscopy of \\epsba, Bb which were compared to the evolutionary models of \\citet{Burrows:1997} and the atmospheric models of \\citet{Burrows:2006}. They derived effective temperatures of 1250--1300\\,K and 875--925\\,K, and surface gravities of log g (cm\\,s$^{-1}$) 5.2--5.3 and 4.9--5.1, for \\epsba\\ and Bb respectively, by comparing their observed spectra with their spectral models scaled using the distance and a radius predicted by their evolutionary models. We will discuss these results further in Sect.\\,\\ref{sec:model_diffs} in contrast to our new data. In this paper we present high signal-to-noise photometry from the $V$- to $M'$-band (0.5--4.9\\micron) and medium resolution spectroscopy from 0.6--5.1\\micron\\ of the individual components of the \\epsba, Bb system. In Sect.\\,\\ref{sec:obs}, we describe the observations and data reduction, including the routines employed to extract the partially-blended photometry and spectroscopy. We re-derive the spectral types of both objects according to the updated classification scheme of \\citet{Burgasser:2006} in Sect.\\,\\ref{sec:spec_class} and discuss constraints imposed by the parent main-sequence star in Sect.\\,\\ref{sec:EpsA_constrain}. We derive the luminosities of both sources in Sect.\\,\\ref{sec:lum} and discuss the preliminary dynamical mass measurement of \\citet[][in prep.]{McCaughrean:2009} in Sect.\\,\\ref{sec:mass}. Our observations are compared to evolutionary models in Sect.\\,\\ref{sec:evo_model_comp} and to atmospheric models in Sect.\\,\\ref{sec:atm_model_comp}. We then put limits on the masses of lower-mass companions in Sect.\\,\\ref{sec:Bc_mass_limits}, and finally the predictions of evolutionary and atmospheric models and previous determinations are compared in Sect.\\,\\ref{sec:model_diffs}. ", "conclusions": "\\label{sec:conc} We have presented the results of a comprehensive photometric and spectroscopic study of the individual components of the nearest known binary brown dwarf system, \\epsba, Bb. The relative proximity of these T1 and T6 dwarfs to the Earth resulted in very high quality data, while archival results for the well-studied parent star, \\epsa, provide invaluable additional information. We find the spectra of these brown dwarfs are best matched by the BT-Settl spectral models with T$_{\\rm{eff}}$=1300--1340\\,K and log g=5.50 for \\epsba\\ and 880--940\\,K and 5.25 for \\epsbb, both with a metallicity of [M/H]=$-0.2$. COND03 evolutionary model predictions for the masses are significantly inconsistent with the measured system mass if the young age range of 0.8--2.0\\,Gyr suggested by \\citet{Lachaume:1999} is used. We find that a system age of 3.7--4.3\\,Gyr is necessary for the COND03 evolutionary models to be consistent with the measured system mass at the observed luminosities, and a review of the literature finds evidence supporting an age of $\\sim$5\\,Gyr for \\eps\\ A\\@. In the age range 3.7--4.3\\,Gyr, the COND03 models predict effective temperatures in the range 1352--1385\\,K and 976--1011\\,K, for Ba and Bb, respectively. It is clear that there are several areas in which the atmospheric models currently do not reproduce observations and a more detailed analysis of these issues will be the subject of future work. They include the strength and shape of the wide absorption by \\ion{K}{I} and \\ion{Na}{I} in the optical, the possible formation of feldspars in mid-late T dwarfs, and the reaction rates of CO and CH$_{4}$. In addition, the spectral shape in the $L$-band caused by CH$_4$ absorption is poorly reproduced, as is also the case for CH$_4$ absorption at $\\sim$1.6\\micron. The $M$-band spectra, although low resolution, also show that the atmospheric models significantly over-estimate the flux in this region. While the flux levels of the near-IR peaks can be reasonably reproduced, the level of absorption between the peaks tends to be problematic. In particular, we find a feature at 1.35--1.40\\micron\\ in both our object spectra which is not predicted in the atmospheric models. Neither source has detectable atomic \\ion{Li}{I} absorption at 6708\\,\\AA. The absence of lithium in the more massive component is consistent with the revised, higher age estimates coupled with its probable dynamical mass, while the lack of absorption in the cooler source is expected from its low effective temperature, where lithium is incorporated into molecules. Although there is significant room for improvement in the atmospheric models, the current match to \\epsba\\ and Bb is nevertheless impressive. When new data on methane opacities become available, we will be able to better reproduce the observed spectra and more reliably compare these spectral models to spectra of objects with less well-constrained physical parameters. Additionally, when these updated atmospheric models are incorporated into the evolutionary models, a fully self-consistent comparison will be possible. Finally, when the individual dynamical masses become available and if we can obtain a reliable estimate of the age of this system, based on asteroseismological observations of the parent star \\epsa, then \\epsba\\ and Bb will become invaluable benchmark objects with a full set of physical parameters which newer models will have to reproduce, making them more reliable for analysing the properties of isolated ultra-cool field dwarfs. The predictions of the evolutionary models using luminosity and mass constraints are somewhat different to the derived effective temperature and surface gravity from fitting atmospheric models to observed spectra. These differences may be resolved when the newer atmosphere models are incorporated into the evolutionary models. However, it seems that derivations of the mass of cool brown dwarfs are uncertain even where estimates of the effective temperature, surface gravity, and luminosity exist. We therefore caution against the over-analysis of predicted brown dwarf masses at this time." }, "0911/0911.4887_arXiv.txt": { "abstract": "{} {We intended to measure the radial velocity curve of the supergiant companion to the eclipsing high mass X-ray binary pulsar {EXO~1722--363} and hence determine the stellar masses of the components.} {We used a set of archival K$_{\\rm s}$-band infrared spectra of the counterpart to {EXO~1722--363} obtained using ISAAC on the VLT, and cross-correlated them in order to measure the radial velocity of the star.} {The resulting radial velocity curve has a semi-amplitude of $24.5 \\pm 5.0$~km~s$^{-1}$. When combined with other measured parameters of the system, this yields masses in the range 1.5 $\\pm$ 0.4 - 1.6 $\\pm$ 0.4 ~M$_{\\odot}$ for the neutron star and 13.6 $\\pm$ 1.6 - 15.2 $\\pm$ 1.9 ~M$_{\\odot}$ for the B0--1 Ia supergiant companion. These lower and upper limits were obtained under the assumption that the system is viewed edge-on (i = 90$^\\circ$) for the lower limit and the supergiant fills its Roche lobe ($\\beta = 1$) for the upper limit respectively. The system inclination is constrained to $i>75^{\\circ}$ and the Roche lobe-filling factor of the supergiant is $\\beta>0.9$. Additionally we were able to further constrain our distance determination to be 7.1 $\\le$ d $\\le$ 7.9 kpc for {EXO~1722--363}. The X-ray luminosity for this distance range is 4.7 $\\times$ 10$^{35}$ $\\le$ L$_{\\rm X}$ $\\le$ 9.2 $\\times$ 10$^{36}$ erg s$^{-1}$.} { {EXO~1722--363} therefore becomes the seventh of the ten known eclipsing X-ray binary pulsars for which a dynamical neutron star mass solution has been determined. Additionally {EXO~1722--363} is the first such system to have a neutron star mass measurement made utilising near-infrared spectroscopy.} {} ", "introduction": "The precise form of the neutron star (NS) equation of state is still unknown. Despite much theoretical work aimed at determining this fundamental aspect of astrophysics, to eliminate some of the contending theories we must turn to observational data. Presently the only means of determining the mass of neutron stars in accretion driven systems is by observing eclipsing X-ray binary pulsars. Unfortunately only 10 such systems are currently known, and only 6 of these have previously had mass measurements made (e.g. \\cite{ash99}, \\cite{quaintrell03}, \\cite{valbaker05}). In this paper we present the preliminary results from our on-going work on the mass of the High Mass X-ray Binary (HMXB) accretion driven pulsar {EXO 1772--363}. The counterpart star within this HMXB is heavily obscured and reddened, necessitating for the first time the utilisation of near-infrared spectroscopy to construct the Radial Velocity (RV) curve in order to obtain an accurate mass solution. Observations made of {EXO~1722--363} (alternatively designated IGR J17252--3616), in 1987 by the \\emph{Ginga} X-ray satellite were the first to detect pulsations. These pulsations were found to have a $413.9 \\pm 0.2$s period \\citep{tawara89}. Over an 8 hour period the source appeared to vary substantially in flux, decreasing from 2~mcrab to 0.2 -- 0.3~mcrab within the 6--21~keV band. It was subsequently found that the X-ray flux remained persistent from 20 -- 60~keV, however a cutoff point was found above this flux level in which the source was undetectable. {EXO~1722--363} at maximum flux and assuming a distance of 10 kpc, had a luminosity calculated as $5 \\times 10^{36}$~erg~s$^{-1}$ \\citep{tawara89}. Later observations by the \\emph{Rossi X-ray Timing Explorer} (RXTE) revealed the eclipsing nature of this system with the eclipse duration determined as $1.7 \\pm 0.1$~days \\citep{corbet05}. Subsequent observations by \\emph{INTEGRAL} followed up by \\emph{XMM-Newton} in 2004 led to a further refinement of the spin and orbital periods to $413.851\\pm0.004$~s and $9.7403\\pm0.0004$~days respectively \\citep{thompson07}. \\emph{XMM-Newton} observations allowed the source position to be determined more precisely (with an uncertainty of $4^{\\prime\\prime}$) at RA(2000.0) = $17^h 25^m 11.4^s$ and Dec = $-36^\\circ 16 ^\\prime 58.6 ^{\\prime\\prime}$. {EXO~1722--363} lies within the Galactic plane and as it is heavily reddened, unsurprisingly the counterpart star could not be detected optically. An infrared counterpart was found lying $1^{\\prime\\prime}$ from the X-ray source position \\citep{zurita06} with a corresponding entry in the 2MASS catalogue, 2MASS~J17251139--3616575 (with JHK magnitudes J = 14.2, H = 11.8 and K$_s$ = 10.7). Examination of near infrared K-band spectra obtained with the ESO \\emph{ISAAC} instrument led to our determination of the spectral classification of the mass donor as B0 -- B1 Ia \\citep{mason09}. ", "conclusions": "Although the 11 spectra reported here are of relatively low quality, and few in number, they still allow us to make a preliminary determination of the orbit of the supergiant in {EXO~1722--363} and make a first measurement of the dynamical masses of the stellar components. The results are encouraging for a number of reasons. First, the resulting neutron star mass is consistent with the canonical mass of 1.4~M$_{\\odot}$ measured in most other eclipsing HMXBs, except for that in Vela X-1, \\citep{quaintrell03}. Second, the measured mass and radius of the supergiant, $M \\sim 13 - 15$~M$_{\\odot}$ and $R \\sim 25 - 28$~R$_{\\odot}$, support the B0-1 Ia spectral classification that we have previously determined \\citep{mason09}. This is illustrated by the Hertzsprung-Russell diagram plotted in Fig. \\ref{evo_track}, which shows a close correspondence between the system primary and the properties of other galactic field BSGs \\citep{searle08}. While the similarity in temperature is to be expected - the value for the primary was adopted on the basis of its spectral type, which in turn has been calibrated by the analysis of \\cite{searle08}, the radii for the field BSGs were determined via non-LTE model atmosphere analysis, while that for the primary is instead determined dynamically. The measurement of the stellar radius and hence bolometric luminosity, has in turn has allowed a more precise determination of the distance to the system by comparison to its observed photometric magnitude and reddening. The refined distance to {EXO~1722--363} of 7.1 - 7.9 kpc results in an X-ray luminosity ranging from {L$_{\\rm {X_{min}}}$ = 0.47} $\\times$ ~10$^{36}$ to L$_{\\rm {X_{max}}}$ = 9.2 $\\times$ ~10$^{36}$ erg s$^{-1}$. However, due to the nature of the archive observations used for this work, the uncertainties on the mass and radius parameters are still rather large; it is our intention in the near future to propose and obtain more accurate VLT/ISAAC observations to further constrain the orbital solution parameters for this HMXB system. Finally, comparison to evolutionary tracks in Fig. \\ref{evo_track} might suggest the primary had an initial progenitor mass of $\\sim$ 35 - 40M$_{\\odot}$ and hence the neutron star originated in a more massive star. However we caution that the binary is highly likely to have undergone at least one episode of mass transfer in the past, rendering such a conclusion highly uncertain. As an exemplar we cite {GX 301-2}, a HMXB composed of a NS and a B hypergiant with a spectroscopic mass of 43$\\pm$10M$_{\\odot}$ \\citep{kaper06}. However \\citep{wellstein99} propose a formation scenario in which two stars of comparable initial masses evolved via quasi conservative mass transfer into the current configuration post supernova; hence determining progenitor masses for both primary and neutron star based on the current system parameters is non trivial. \\begin{figure} \\includegraphics[width=9cm]{fig3_res.eps} \\begin{center} \\caption{Position of {EXO 1722-363} on the Hertzsprung-Russell \\citep{searle08} diagram alongside a sample of O and B supergiants \tfrom differing locations, Galactic B supergiants, \\cite{crowther06}, SMC B supergiants (\\cite{trundle04}; \\cite{trundle05}) and Galactic O stars, \\cite{repolust04}. These are overplotted together with solar metallicity evolutionary tracks from \\cite{meynet00}. Also shown is the lower and upper limits on the luminosity of {EXO 1772-363}.} \\label{evo_track} \\end{center} \\end{figure}" }, "0911/0911.3966_arXiv.txt": { "abstract": "We investigate non-axisymmetric low frequency modes of a rotating and magnetized neutron star, assuming that the star is threaded by a dipole magnetic field whose strength at the stellar surface, $B_0$, is less than $\\sim 10^{12}$G, and whose magnetic axis is aligned with the rotation axis. For modal analysis, we use a neutron star model composed of a fluid ocean, a solid crust, and a fluid core, where we treat the core as being non-magnetic assuming that the magnetic pressure is much smaller than the gas pressure in the core. For non-axisymmetric modes, spheroidal modes and toroidal modes are coupled in the presence of a magnetic field even for a non-rotating star. Here, we are interested in low frequency modes of a rotating and magnetized neutron star whose oscillation frequencies are similar to those of toroidal crust modes of low spherical harmonic degree and low radial order. For a magnetic field of $B_0\\sim 10^7$G, we find Alfv\\'en waves in the ocean have similar frequencies to the toroidal crust modes, and we find no $r$-modes confined in the ocean for this strength of the field. We calculate the toroidal crustal modes, the interfacial modes peaking at the crust/core interface, and the core inertial modes and $r$-modes, and all these modes are found to be insensitive to the magnetic field of strength $B_0\\ltsim10^{12}$G. We find the displacement vector of the core $l^\\prime=|m|$ $r$-modes have large amplitudes around the rotation axis at the stellar surface even in the presence of a surface magnetic field $B_0\\sim10^{10}$G, where $l^\\prime$ and $m$ are the spherical harmonic degree and the azimuthal wave number of the $r$-modes, respectively. We suggest that millisecond X-ray variations of accretion powered X-ray millisecond pulsars can be used as a probe into the core $r$-modes destabilized by gravitational wave radiation. If the $l^\\prime=|m|=2$ $r$-mode is excited, we will have the pulsation of the frequency $\\sim4\\Omega/3$ with $\\Omega$ being the spin frequency of the star. ", "introduction": "Oscillation of strongly magnetized neutron stars has attracted an intense attention in recent years, particularly motivated by the discovery of quasi-periodic oscillations (QPOs) of magnetar candidates (e.g., Woods \\& Thompson 2006), which are believed to be one of the observational manifestations of global oscillations of neutron stars that have a strong global magnetic field of order of $B_0\\sim 10^{15}{\\rm G}$ at the stellar surface (e.g., Duncun 1998, Israel et al 2005; Strohmayer \\& Watts 2005, 2006; Watts \\& Strohmayer 2006). Thus, recent theoretical studies of the oscillations of magnetized neutron stars have been mainly concerned with the stars having extremely strong surface magnetic fields $B_0\\gtsim 10^{15}$G, and these studies have suggested that the QPOs observed in the magnetar candidates are attributable to the toroidal crust modes and Alfv\\'en modes of the neutron stars (e.g., Piro 2005; Glampedakis, Samuelsson \\& Andersson 2006; Lee 2007, 2008; Sotani, Kokkotas \\& Stergioulas 2008; Cerd\\'a-Dur\\'an, Stergioulas \\& Font 2009; Colaiuda, Beyer \\& Kokkotas 2009; Bastrukov et al 2009; Sotani \\& Kokkotas 2009). For toroidal modes of strongly magnetized neutron stars, however, it is intriguing, from the theoretical point of view, that Lee (2008) obtained discrete frequency spectra of the magnetic modes, but Sotani, Kokkotas \\& Stergioulas (2008), Cerd\\'a-Dur\\'an, Stergioulas \\& Font (2009), and Colaiuda, Beyer \\& Kokkotas (2009), using a different numerical method from that used by Lee (2008), suggested the existence of continuum frequency spectra of the modes, as originally discussed by Levin (2006, 2007). The burst oscillation observed in low mass X-ray binaries (LMXBs) can be another example in which a magnetic field plays an important role in the oscillations of neutron stars, although the strength of the field at the surface of the neutron stars in LMXBs is thought to be less than $\\sim10^{10}$G, much weaker than that for the magnetar candidates. For the burst oscillation, the hot spot model (e.g., Strohmayer et al 1997; Cumming \\& Bildsten 2001; Cumming et al 2002) and the Rossby wave model (Heyl 2004; Lee 2004) have been proposed, but it seems none of the models is accepted as the one that fully explains the observational properties of the burst oscillation. In the Rossby wave model, Heyl (2004, 2005) , Lee (2004), and Lee \\& Strohmayer (2005) have examined the possibility that the burst oscillation is produced by low frequency Rossby waves, called $r$-modes in the astrophysical literature, propagating in the surface fluid region (ocean), disregarding the effects of a magnetic field on the low frequency waves. We note that Bildsten \\& Cutler (1995) discussed the modal properties of $g$ modes propagating in the fluid ocean above the solid crust of mass accreting neutron stars as a possible mechanism responsible for the $\\sim 6$Hz QPOs observed in LMXBs. In their paper, on the assumption that the magnetic pressure, $p_B$, is much smaller than the gas pressure, $p$, the critical strength of a magnetic field, $B_c$, below which the magnetic field has no significant effects on the $g$-modes, was estimated to be $B_c\\sim2\\times 10^{10}$G for the accreted envelopes composed of carbon (see also Piro \\& Bildsten 2005). Since their argument is based on a local analysis and on the assumption $p_B\\ll p$, we think it useful to carry out a global modal analysis of low frequency waves propagating in the magnetized fluid ocean of a neutron star. Accretion powered millisecond X-ray pulsars in LMXBs show small amplitude, almost sinusoidal X-ray time variations, the dominant period of which is thought to be the spin period of the underlying neutron stars. Lamb et al (2009) argued that the millisecond X-ray variations are produced by an X-ray emitting hot spot located at a magnetic pole rotating with the star, and that so long as the symmetry center of the hot spot is only slightly off the rotation axis the X-ray variations produced have small amplitudes and become almost sinusoidal. They also suggested that if the hot spot is located close to the rotation axis, a slight drift of the hot spot away from the rotation axis leads to appreciable changes in the amplitudes and phases of the X-ray variations. Lamb et al (2009) pointed out that a temporal change in mass accretion rates and hence the radius of the magnetosphere, for example, can cause such a drift of the hot spot. We think it is also interesting to consider the possibility that global oscillations of the neutron stars work as a mechanism that perturbs the hot spot periodically. In this paper, to calculate global oscillations of a rotating and magnetized neutron star, we use the method of series expansion, in which the angular dependence of the perturbations is represented by series expansion in terms of spherical harmonic functions with different spherical harmonic degrees $l$ for a given azimuthal wave number $m$ (e.g., Lee \\& Strohmayer 1996, Lee 2007). We calculate low $m$, low frequency modes of a neutron star composed of a fluid ocean, a solid crust, and a fluid core, where the star is assumed to be threaded by a dipole magnetic field, the strength $B_0$ of which at the surface is smaller than $\\sim10^{12}$G. The method of calculation we employ is presented in \\S 2, and the numerical results are described in \\S 3, and we discuss a local analysis for low frequency modes in the magnetized fluid ocean in \\S 4, and we give discussion and conclusion in \\S 5. ", "conclusions": "Lamb et al (2009) proposed that the small amplitude, almost sinusoidal millisecond X-ray pulsation observed in accretion powered millisecond X-ray pulsars may be well explained by the hot spot model, in which the hot spots are assumed to be located at the magnetic poles, which are nearly aligned with the rotation axis. As discussed by Lamb et al (2009), even a small drift of the hot spot could produce appreciable changes in pulsation amplitudes of the X-ray pulsation. If this proposition is correct, it is interesting to pursue a possibility of using the millisecond X-ray pulsation to probe the core $r$-modes excited by gravitational wave radiation (Andersson 1998; Friedman \\& Morsink 1998). In fact, if the core $r$-modes with $|m|\\ge2$ are excited by the emission of gravitational wave, since the $l^\\prime=m=2$ $r$-mode, which is the most strongly destabilized mode among the $r$-modes (e.g., Lockitch \\& Friedman 1999; Yoshida \\& Lee 2000a), produces the surface displacement vector $\\pmb{\\xi}$ whose horizontal and toroidal components at the surface have large amplitudes around the rotation axis as shown by Figure 6, the hot spot could suffer periodic disturbance from the $r$-mode. Note that the $r$-mode induced temperature perturbation, the surface pattern of which may be proportional to $X^r\\left(\\theta\\right)e^{{\\rm i}m\\phi}$, might generate X-ray variations, the amplitudes of which should be very small. If we write the oscillation frequency of $r$-modes as \\be {\\omega/\\Omega}=\\kappa_0+\\kappa_2\\bar\\Omega^2+O\\left(\\bar\\Omega^4\\right), \\ee the coefficient $\\kappa_0$ for the modes is given by \\be \\kappa_0=2m/ \\left[l^\\prime\\left(l^\\prime+1\\right)\\right], \\ee and the coefficient $\\kappa_2$ may depend on the equation of state and the deviation from the isentropic stratification in the core (e.g., Yoshida \\& Lee 2000a,b), where $\\bar\\Omega=\\Omega/\\Omega_0$. Since the neutron star core is nearly isentropic such that $N^2\\sim0$, we only have to consider the $l^\\prime=|m|$ $r$-modes, for which we have $\\omega\\approx\\kappa_0\\Omega=2\\Omega/\\left(|m|+1\\right)$, and we obtain the frequency $\\omega\\approx 2\\Omega/3$ for $m=2$ in the corotating frame of the star and the frequency $\\sigma\\equiv\\omega-m\\Omega\\approx -4\\Omega/3$ in an inertial frame. It may be interesting to point out that if the $r$-mode of $l^\\prime=m=1$ is also excited by some mechanism, this $r$-mode can produce long period variations in an inertial frame since the inertial frame frequency $\\sigma\\sim0$ for this mode. Although no detection of periodicities whose frequency is approximately equal to $4\\Omega/3$ has so far been reported, if we detect the periodicities produced by the core $r$-modes of $l^\\prime=|m|$ in the X-ray millisecond pulsation, we can use the frequency deviation given by $\\Delta\\bar\\omega\\equiv\\bar\\omega-\\kappa_0\\bar\\Omega\\approx\\kappa_2\\bar\\Omega^3$ to derive information about the equations of state and the thermal stratification in the core. The detectability of the $r$-mode pulsation may depend on the thickness of the crust, the property of the fluid ocean, and the strength of the magnetic field and so on. We have calculated non-axisymmetric low frequency modes of a rotating and magnetized neutron star, where we used a neutron star model composed of a surface fluid ocean, a solid crust, and a fluid core. We have assumed that the star is threaded by a dipole magnetic field but the fluid core can be treated as a non-magnetic region. For this model, we found that for a magnetic field of strength $B_0\\sim10^7$G, Alfv\\'en waves in the surface ocean come in as low frequency modes, which largely modify the gravito-inertial modes in the ocean. However, the oscillation frequencies of the modes in the ocean are dependent on $j_{\\rm max}$, and they do not reach any good convergence even if $j_{\\rm max}$ is increased to $\\sim20$. At present it is not clear that these ocean modes will converge to discrete modes with real frequencies in the limit of $j_{\\rm max}\\rightarrow \\infty$. We also found that no $r$-modes, which are confined to the surface ocean, can be found in the presence of the weak magnetic field. If this is also the case for the accreted envelopes expected in mass accreting neutron stars in binary systems, we need to reconsider the Rossby wave model for the burst oscillation in LMXBs (Heyl 2004, Lee 2004). On the other hand, the toroidal crust modes and the interfacial modes at the core/crust interface, which show good convergence for $j_{\\rm max}\\gtsim10$, are found to be insensitive to magnetic fields of strength $B_0\\ltsim10^{12}$G. We also find that the core $r$-modes and inertial modes are not affected by the magnetic field even if their eigenfunctions extend to the surface through the magnetized crustal and envelope regions. However, this will not be the case for neutron stars having magnetic fields as strong as $B_0\\sim10^{15}$G (e.g., Lee 2008). In the present paper, we employed for modal analysis a low mass neutron star model having a thick solid crust and a cold and thin surface ocean, for which the toroidal crust modes of low radial order and low spherical harmonic degree are well separated from the $f$- and $p$-modes. If we use more massive neutron stars with a hot accreted fluid envelope and a thin solid crust, the frequencies of the crust modes of low radial order and those of the $f$- and $p$-modes may overlap, leading to more complicated frequency spectra. We think it necessary to conduct similar modal analyses for such neutron star models to clarify the properties of the ocean modes in the presence of a magnetic field and to examine the possibility that the small amplitude millisecond X-ray pulsation can be used as a probe into the core $r$-modes. \\begin{appendix}" }, "0911/0911.3410.txt": { "abstract": "We assess the validity of a single step Godunov scheme for the solution of the magneto-hydrodynamics equations in more than one dimension. The scheme is second-order accurate and the temporal discretization is based on the dimensionally unsplit Corner Transport Upwind (CTU) method of Colella. The proposed scheme employs a cell-centered representation of the primary fluid variables (including magnetic field) and conserves mass, momentum, magnetic induction and energy. A variant of the scheme, which breaks momentum and energy conservation, is also considered. Divergence errors are transported out of the domain and damped using the mixed hyperbolic/parabolic divergence cleaning technique by Dedner et al. (J. Comput. Phys., 175, 2002). The strength and accuracy of the scheme are verified by a direct comparison with the eight-wave formulation (also employing a cell-centered representation) and with the popular constrained transport method, where magnetic field components retain a staggered collocation inside the computational cell. Results obtained from two- and three-dimensional test problems indicate that the newly proposed scheme is robust, accurate and competitive with recent implementations of the constrained transport method while being considerably easier to implement in existing hydro codes. %This paper aims to compare two different numerical schemes %for computational MHD in multidimensions. %In the constrained transport method, the magnetic field %has a staggered representation whereby field components live %on the face they are normal to while the remaining variables %retain their usual collocation at the cell center. %This provides a framework by which the induction equation %is more naturally updated using Stoke\u0092s theorem and the divergence-free %condition is fulfilled to machine accuracy. ", "introduction": "\\label{sec:intro} % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% A primary aspect in building stable and robust Godunov type schemes for the numerical solution of the compressible magnetohydrodynamics (MHD) equations relies on an accurate way to control the solenoidal property of the magnetic field while preserving the conservation properties of the underlying physical laws. Failure to fulfill either requisite has been reported as a potential hassle leading to unphysical effects such as plasma acceleration in the direction of the field, incorrect jump conditions, wrong propagation speed of discontinuities and odd-even decoupling, see \\cite{Toth00,BK04}. A comprehensive body of literature has been dedicated to this subject and several strategies to enforce the $\\nabla\\cdot\\vec{B}=0$ condition in Godunov-type codes have been proposed, see for example \\cite{ZMC94, RJF95, RMJF98, BS99, Toth00} and, more recently, \\cite{Balsara04, LdZ04, GS05, Rossmanith06, MBMR08}. The robustness of one method over another can be established on a practical base by extensive numerical testing, see \\cite{Toth00, BK04}. In a first class of schemes, the magnetic field is discretized as a cell-centered quantity and the usual formalism already developed for the Euler equation can be extended in a natural way. Cell-centered methods are appealing since the extensions to adaptive and/or unstructured grids are of straightforward implementation. Moreover, the same interpolation scheme and stencil used for the other hydrodynamic variables can be easily adapted since all quantities are discretized at the same spatial location, thus facilitating the extension to schemes possessing higher than second order accuracy. Unfortunately, numerical methods based on a cell-centered discretization do not naturally preserve Gauss's law of electromagnetism, even if $\\nabla\\cdot\\vec{B}=0$ initially. In the approach suggested by Powell \\cite{Powell94, Powell99}, Gauss's law for magnetism is discarded in the derivation of the MHD equations and the resulting system of hyperbolic laws is no longer conservative by the appearance of a source term proportional to $\\nabla\\cdot\\vec{B}$. Although the source term should be physically zero at the continuous level, Powell showed that its inclusion changes the character of the equations by introducing an additional eighth wave corresponding to the propagation of jumps in the component of magnetic field normal to a given interface. A different approach is followed in the projection scheme \\cite{BB80, ZMC94, RJF95, Crockett05}, where a Helmholtz-Hodge decomposition is applied to resolve $\\vec{B}$ as the sum of an irrotational and a solenoidal part, associated with scalar and vector potentials. A cleaning step allows to recover the divergence-free magnetic field by subtracting the unphysical contribution coming from the irrotational component at the extra cost of solving a Poisson equation. In the approach of Dedner et al. \\cite{Dedner02}, the divergence-free constraint is enforced by solving a modified system of conservation laws where the induction equation is coupled to a generalized Lagrange multiplier. Dedner et al. showed that the choice of mixed hyperbolic/parabolic correction offers both propagation and dissipation of divergence errors with the maximal admissible characteristic speed, independently of the fluid velocity. This approach preserves the full conservation form of the original MHD system at the minimal cost of introducing one additional variable in the system and will be our scheme of choice. Finally, Torrilhon \\cite{Torrilhon05} (see also \\cite{AT08}) showed a general procedure to modify the inter-cell fluxes in the framework of a flux distribution scheme that preserves the value of a certain discrete divergence operator in each control volume. %In cell-centered methods the divergence-free constraint is preserved %either at the truncation error of the scheme or it is exactly %zero in some particular $\\nabla\\cdot\\vec{B} = 0$ at the %truncation level. Alternative approaches suggested by Toth (Field CD). A different strategy is followed in the constrained transport (CT) methods, originally devised by \\cite{EH88} and later built into the framework of shock-capturing Godunov methods by a number of investigators, e.g., \\cite{BS99, Balsara04, LdZ04, GS05, GS08}. In this class of schemes, the magnetic field has a staggered representation whereby the different components live on the face they are normal to. Hydrodynamic variables (density, velocity and pressure) retains their usual collocation at the cell center. CT schemes preserve the divergence-free condition to machine accuracy in an integral sense since the magnetic field is treated as a surface averaged quantity and thus more naturally updated using Stokes' theorem. This evolutionary step involves the construction of a line-averaged electric field along the face edges, thereby requiring some sort of reconstruction or averaging of the electromotive force from the face center (where different components are usually available as face centered upwind Godunov fluxes) to the edges. A variety of different strategies have been suggested, including simple arithmetic averaging \\cite{BS99, RMJF98}, solution of 2-D Riemann problems \\cite{LdZ04, FHT06} or other somewhat more empirical approaches \\cite{GS05,GS08,LD09}. The staggered collocation of magnetic and electric field variables in CT schemes makes their extension to adaptive grids rather arduous and costly. Besides, significant efforts have to be spent in order to develop schemes with spatial accuracy of order higher than second. An alternative constrained transport method, based on the direct solution of the magnetic potential equation (thus avoiding staggered grids), has been presented by \\cite{Rossmanith06}. In the present work we propose a new fully unsplit Godunov scheme for multidimensional MHD, based on a combination of the Corner Transport Upwind of \\cite{Colella90} and the mixed hyperbolic/parabolic divergence cleaning technique of \\cite{Dedner02} (CTU-GLM). The proposed scheme has second order accuracy in both space and time and adopts a cell-centered spatial collocation (no staggered mesh) of all flow variables, including the magnetic field. The scheme is fully conservative in mass, momentum, magnetic induction and energy and the divergence-free constraint is enforced via a mixed hyperbolic/parabolic correction which avoids the computational cost associated with an elliptic cleaning deriving from a Hodge projection. A variant of the scheme, which introduces divergence source terms breaking the conservative properties of some equations, is also presented. We assess the accuracy and robustness of the scheme by a direct quantitative comparison with the 8-wave formulation of \\cite{Powell99} and the recently developed constrained transport method of \\cite{GS05, GS08}. Other similar implementations may be found in \\cite{FHT06,LD09}. The comparison is conveniently handled using the PLUTO code for computational astrophysics \\cite{Mignone07} where both cell-centered and staggered-mesh implementations are available. Our motivating efforts are driven by issues of simplicity, efficiency and flexibility. In this sense, the benefits offered by a method where all of the primary flow variables are discretized at the same spatial location considerably ease the extension to adaptive grids, to more complex physics and to schemes with higher than second order accuracy. The latter possibility will be explored in a companion paper. %Besides the ease of implementation, our motivations reside mainly in the convenience %offered by a scheme where all of the primary flow variables are discretized %at the same location, that is, at the cell midpoint. This endeavor finds its %justification in the extension to adaptive grids and to scheme of higher %than second order accuracy which will be presented in a companion paper. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% ", "conclusions": "\\label{sec:conclusions} % % % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% A second-order, cell-centered numerical scheme for the solution of the MHD equations in two and three dimensions has been proposed. Fully unsplit integration resorts to the Corner Transport Method of Colella \\cite{Colella90} and the divergence-free condition is controlled by using a constrained formulation of the MHD equations where the induction equation is coupled to a generalized Lagrange multiplier (GLM, \\cite{Dedner02}). The system is hyperbolic, easy to implement and does not require expensive cleaning projection steps associated with the solution of elliptic problems. The GLM scheme is fully conservative in mass, momentum, energy and magnetic induction, although we have also considered a slightly modified variant (EGLM) which infringes momentum and energy conservation. In order to assess the reliability and accuracy of the schemes we have performed a number of code benchmarks on standard two- and three-dimensional MHD test problems. Results have been compared with two different numerical schemes: a non-conservative cell-centered method based on the 8-wave formulation (8W, \\cite{Powell99}) and the constrained transport (CT) method where the magnetic field has a staggered collocation. Both the GLM and EGLM schemes give excellent results in terms of accuracy and robustness and do not show, in the tests presented here, any evidence for incorrect jump conditions or wrong wave propagation, as found for the eight wave formulation (in agreement with T\\'oth \\cite{Toth00}). This has been verified on problems involving discontinuous waves and holds true for both the conservative GLM formulation \\emph{and} the EGLM variant which breaks momentum and energy conservation. In this perspective, our results seem to indicate that the presence of source terms in the equations does not necessarily lead to erroneous jumps. Instead, we have found the non-conservative formulation to be more robust for problems involving the propagation of oblique strongly magnetized shocks. Although, this behavior may be attributed to discretization, such a study is beyond the scope of the present paper. The comparison has also revealed an excellent quantitative agreement with the CTU-CT scheme (in the version of \\cite{GS05,GS08}) showing errors with comparable magnitude and similar order of convergence while retaining the desired robustness and stability. For these reasons, we believe that the proposed CTU-GLM and CTU-EGLM schemes provide excellent competitive alternatives to modern staggered-mesh algorithms while being considerably easier and more flexible in their implementations. Owing to the cell-centered collocation of all of the flow fields, the CTU-GLM scheme can be easily generalized to resistive MHD, adaptive and/or unstructured grids and to higher than second-order spatially-accurate numerical schemes. Some of these issues will be presented in forthcoming papers. %Local divergence errors are accurately monitored and controlled %at a very small computational cost. %% The Appendices part is started with the command \\appendix; %% appendix sections are then done as normal sections %% \\appendix %%" }, "0911/0911.1364_arXiv.txt": { "abstract": "{In a previous paper, we found that globular clusters in our Galaxy lie close to a line in the ($\\log{R_e}$, ${\\mathit{SB}}_e$, $\\log{\\sigma}$) parameter space, with a moderate degree of scatter and remarkable axi-symmetry. This implies that a purely photometric scaling law exists, that can be obtained by projecting this line onto the ($\\log{R_e}$, ${\\mathit{SB}}_e$) plane. These photometric quantities are readily available for large samples of clusters, as opposed to stellar velocity dispersion data.}{We study a sample of $129$ Galactic and extragalactic clusters on this photometric plane in the V-band. We search for a linear relation between ${\\mathit{SB}}_e$ and $\\log{R_e}$ and study how the scatter around the best-fit relation is influenced by both age and dynamical environment. We interpret our results in terms of testing the evolutionary versus primordial origin of the fundamental line.}{We perform a detailed analysis of surface brightness profiles, which allows us to present a catalogue of structural properties without relying on a given dynamical model.}{We find a linear relation between ${\\mathit{SB}}_e$ and $\\log{R_e}$, in the form ${\\mathit{SB}}_e = (5.25 \\pm 0.44) \\log{R_e} + (15.58 \\pm 0.28)$, where ${\\mathit{SB}}_e$ is measured in mag/arcsec$^2$ and $R_e$ in parsec. Both young and old clusters follow the scaling law, which has a scatter of approximately $1$ mag in ${\\mathit{SB}}_e$. However, young clusters display more of a scatter and a clear trend in this with age, which old clusters do not. This trend becomes tighter if cluster age is measured in units of the cluster half-light relaxation time. Two-body relaxation therefore plays a major role, together with passive stellar population evolution, in shaping the relation between ${\\mathit{SB}}_e$, $\\log{R_e}$, and cluster age. We argue that the $\\log{R_e}$-${\\mathit{SB}}_e$ relation and hence the fundamental line scaling law does not have a primordial origin at cluster formation, but is rather the result of a combination of stellar evolution and collisional dynamical evolution.}{} ", "introduction": "\\label{Sect:Introduction} In a previous paper \\citep[][hereafter Paper I]{MyFirstPaper}, we demonstrated that Galactic globular clusters (Galactic GCs) occupy a narrow region around a line in the ($\\log{R_e}$, ${\\mathit{SB}}_e$, $\\log{\\sigma}$) parameter space. This fundamental line is located within the fundamental plane of Galactic GCs \\citep[see][]{Djorgovski1995} and its existence was noted by \\cite{Bellazzini} based on principal component analysis techniques. \\cite{Bellazzini} divided the observed Galactic GCs into two sets of inner GCs and outer GCs with respect to the Sun's orbit, and found that outer GCs have a dimensionality of $1$ in the ($\\log{R_c}$, ${\\mathit{SB}}_0$, $\\log{\\sigma}$) space\\footnote{The photometric quantities considered by \\cite{Bellazzini} are core/central quantities (i.e., core radius $R_c$ and central surface brightness ${\\mathit{SB}}_0$), as opposed to the half-light ones adopted in Paper I and in this paper (i.e., half-light radius $R_e$ and average surface brightness within it ${\\mathit{SB}}_e$).}, while the dimensionality rises to $2$ for inner GCs. He argued that inner GCs, which are closer to the Galactic bulge and interact more frequently with the disk, are more dynamically disturbed than outer GCs. Therefore, he suggested that dynamical interactions with the environment are more likely to disrupt than to preserve a fundamental line, which was then speculated to be primordial in origin. By projecting the fundamental line onto the ($\\log{R_e}$, ${\\mathit{SB}}_e$) plane, in Paper I we predicted the existence of a purely photometric scaling law in the form of a linear relation between ${\\mathit{SB}}_e$ and $\\log{R_e}$. This relation can be practically studied for a much larger sample than the fundamental plane, because velocity dispersion data is not required. In particular, bright clusters (both young and old) in the LMC, SMC, and Fornax galaxies may be included in the study, together with Galactic GCs. This allows a comparison to be made between the behaviors in the ($\\log{R_e}$, ${\\mathit{SB}}_e$) plane, of clusters that have significantly different ages and are associated with different environments. The choice of using only photometric variables allows us to consider a relatively large sample of clusters with a substantial age spread. We are then able to attack the problem of the origin and evolution of the ${\\mathit{SB}}_e$-$\\log{R_e}$ scaling law by studying its dependence on cluster age and to look for evidence for or against the picture in which this scaling law originates primordially and diminishes with time because of dynamical evolution. In the present paper, we follow Paper I in deriving cluster structural parameters in a model-independent way, in particular by avoiding fitting a given dynamical model to observational data. Some minor changes in the model-independent algorithm used are described in Sect.~\\ref{Sect:TheData}, where a detailed account of the data that we collect from the literature is also given. We compile a catalogue of fraction-of-light radii, average surface brightness, and absolute magnitudes, as a machine-readable table. Central cusps in the surface brightness profile have been observed in both old, Galactic GCs \\citep[][]{NoyolaGalacticSlopes} and extragalactic clusters \\citep[][]{NoyolaExtragalacticSlopes} and are predicted to form after core-collapse by N-body simulations \\citep[][]{Tr2009}. Extended cluster envelopes have also been observed in young clusters in external galaxies \\citep[e.g.,][]{LarsenEnvelopes, SchweitzerEnvelopes}. The simple, single-mass, spherical isotropic \\cite{KingModels} models that are usually taken to fit cluster surface brightness profiles have a flat core and sharp tidal cutoff and therefore have difficulties in modeling clusters with these features. The problem of extended envelopes prompted \\cite{CatMcLaughlin} to refit a sample of Milky Way, LMC, SMC, and Fornax clusters with King models, power-law models, and \\cite{WilsonModels} models, finding that, for a sizeable fraction of the old GC population, King models do not perform more effectively than either \\cite{WilsonModels} or power-law models in fitting surface brightness profiles. More generally, the construction of physically-motivated self-consistent models continues to advance, by including features such as tidal flattening \\citep[e.g., see][]{GBVarri}, but model-independent approaches remain a valid complement to model-based analyses \\citep[see][]{KronMayall}. While a certain dependence on specific assumptions, such as the details of the extrapolation of the external part of surface brightness profiles, cannot be avoided, the catalogue of model-independent GC structural parameters presented in this paper naturally emerges as a tool useful for avoiding the potentially limiting assumptions required by dynamical models. The fundamental line and its scatter are presented in Sect.~\\ref{Sect:Results}. Discussion and conclusions are given in Sect.~\\ref{Sect:DiscussionAndConclusions}. ", "conclusions": "\\label{Sect:DiscussionAndConclusions} By considering a sample of $129$ Galactic and extragalactic GCs, we have found a linear relation between ${\\mathit{SB}}_e$ and $\\log R_e$, given by Eq.~(\\ref{Eq:SBelogReRelat}). The relevant coefficient $5.25 \\pm 0.44$ is compatible with $5$, i.e., with cluster absolute magnitude not showing a systematic linear trend with $\\log R_e$. Indeed, when examined in the ($\\log R_e$, $M_V$) plane, the cluster absolute magnitude shows no trend with size \\citep[in contrast to the trend exhibited by dwarf spheroidal galaxies; see][]{vandenBerghNearbyDSPH}. The typical scatter about the relation is $1$ mag in ${\\mathit{SB}}_e$. Young clusters, of age below $4$ Gyr, exhibit a larger scatter of about $1.4$ mag. GC age seems to be an important factor in driving this scatter, because young clusters are found to display a clear trend of residuals to the relation with age, with a correlation coefficient of $r = 0.79$, while old GCs do not exhibit signs of such a correlation. If we assume that all the massive young clusters included in our sample are genuine GC progenitors, our finding suggests that the scaling law being considered shows signs of evolution with age, i.e., that GCs of different ages differ with respect to this scaling law. Moreover, the larger scatter displayed by young clusters points to an evolutionary origin of the relation between ${\\mathit{SB}}_e$ and $\\log R_e$, as opposed to a primordial one. This result is in contrast to the suggestion of \\cite{Bellazzini} that ``at earlier times, globular clusters populated a line in the three-dimensional S-space, i.e., their original dynamical structure was fully determined by a single physical parameter''. In principle, modeling the processes that generate the relation between ${\\mathit{SB}}_e$ and $\\log R_e$ would require taking into account passive stellar evolution, internal dynamical phenomena (evaporation and core-collapse driven by two-body relaxation), and dynamical interaction with the environment. We find that the trend with age becomes stronger for young clusters if age is measured in units of the half-light relaxation time of each GC, i.e., if \\emph{dynamical age} with respect to two-body relaxation is considered, the correlation coefficient rises to $0.90$. Given this result, we argue that two-body relaxation is the key ingredient in shaping the evolution of the ${\\mathit{SB}}_e$-$\\log R_e$ scaling law, at least for young clusters. This may not be the case for older clusters, which do not show such a trend. The evolution of these clusters in the ($\\log R_e$, ${\\mathit{SB}}_e$) plane is likely to be influenced by their host environment, but the available data is of limited use in resolving this issue. The distance to the host galaxy is only a simple proxy for the degree of dynamical disturbance that the GC suffers from it, while more detailed orbital data is available only for a sample of $48$ old, Galactic GCs \\citep[][]{AMPOrbits48GC}. In any case, we find no trend between the residuals to the ${\\mathit{SB}}_e$-$\\log R_e$ relation and either distance from the host galaxy or pericenter distance and orbital eccentricity, when these data are available. When GC age is measured in units of the bulge-interaction and disk-interaction destruction times based on the orbital data provided by \\cite{AMPOrbits48GC}, weak ($r \\approx 0.5$) trends are found with ${\\mathit{SB}}_e$-$\\log R_e$ relation residuals. While this suggests that dynamical interaction with the galactic environment may exert an influence on GCs with observable effects in the ($\\log R_e$, ${\\mathit{SB}}_e$) plane, to draw conclusive evidence, the correlation with age alone should be removed from the trend and a larger sample of clusters with accurately measured orbits is necessary. In contrast to what we noted for fundamental plane residuals in Paper I, we find here that central surface brightness profile slopes measured by \\cite{NoyolaGalacticSlopes} and \\cite{NoyolaExtragalacticSlopes} do not correlate with residuals from the ${\\mathit{SB}}_e$-$\\log R_e$ relation. This result suggests that velocity dispersion information is necessary to understand the mechanism producing cuspy GC surface brightness profiles. \\emph{Acknowledgments.} We wish to thank S. Degl'Innocenti, W. Harris, M. Lombardi, M. Trenti, E. Vesperini, and D. Kruijssen for their comments and helpful suggestions. This research has made use of the SIMBAD data-base, operated at CDS, Strasbourg, France. Part of this work was carried out at the Kavli Institute for Theoretical Physics, Santa Barbara (CA), while we participated in the Program \\emph{Formation and Evolution of Globular Clusters}." }, "0911/0911.3361_arXiv.txt": { "abstract": "{ The mass-loss rate is a key parameter of hot, massive stars. Small-scale inhomogeneities (clumping) in the winds of these stars are conventionally included in spectral analyses by assuming optically thin clumps, a void inter-clump medium, and a smooth velocity field. To reconcile investigations of different diagnostics (in particular, unsaturated UV resonance lines vs. $\\rm H_{\\alpha}$/radio emission) within such models, a highly clumped wind with very low mass-loss rates needs to be invoked, where the resonance lines seem to indicate rates an order of magnitude (or even more) lower than previously accepted values. If found to be realistic, this would challenge the radiative line-driven wind theory and have dramatic consequences for the evolution of massive stars. } { We investigate basic properties of the formation of resonance lines in small-scale inhomogeneous hot star winds with non-monotonic velocity fields.} { We study inhomogeneous wind structures by means of 2D stochastic and pseudo-2D radiation-hydrodynamic wind models, constructed by assembling 1D snapshots in radially independent slices. A Monte-Carlo radiative transfer code, which treats the resonance line formation in an axially symmetric spherical wind (without resorting to the Sobolev approximation), is presented and used to produce synthetic line spectra.} { The optically thin clumping limit is only valid for very weak lines. The detailed density structure, the inter-clump medium, and the non-monotonic velocity field are all important for the line formation. We confirm previous findings that radiation-hydrodynamic wind models reproduce observed characteristics of strong lines (e.g., the black troughs) without applying the highly supersonic `microturbulence' needed in smooth models. For intermediate strong lines, the velocity spans of the clumps are of central importance. Current radiation-hydrodynamic models predict spans that are too large to reproduce observed profiles unless a very low mass-loss rate is invoked. By simulating lower spans in 2D stochastic models, the profile strengths become drastically reduced, and are consistent with higher mass-loss rates. To simultaneously meet the constraints from strong lines, the inter-clump medium must be non-void. A first comparison to the observed Phosphorus V doublet in the O6 supergiant $\\lambda$~Cep confirms that line profiles calculated from a stochastic 2D model reproduce observations with a mass-loss rate approximately ten times higher than that derived from the same lines but assuming optically thin clumping. Tentatively this may resolve discrepancies between theoretical predictions, evolutionary constraints, and recent derived mass-loss rates, and suggests a re-investigation of the clump structure predicted by current radiation-hydrodynamic models. } {} ", "introduction": "\\label{Introduction} Mass loss through supersonic stellar winds is pivotal for the physical understanding of hot, massive stars and their surroundings. A change of only a factor of two in the mass-loss rate has a dramatic effect on massive star evolution \\citep{Meynet94}. Winds from these stars are described by the line-driven wind theory \\citep{Castor75,Pauldrach86}, which traditionally assumes the wind to be stationary, spherically symmetric, and homogeneous. Despite this theory's apparent success \\citep[e.g.,][]{Vink00}, evidence for an inhomogeneous and time-dependent wind has over the past years accumulated, recently summarized in the proceedings from the workshop `Clumping in hot star winds' \\citep{Hamann08} and in a general review of mass loss from hot, massive stars \\citep{Puls08}. That line-driven winds should be intrinsically unstable was already pointed out by \\citet{Lucy70}, and was later confirmed first by linear stability analyses and then by direct, radiation-hydrodynamic modeling of the time-dependent wind \\citep[e.g.,][]{Owocki84,Owocki88,Feldmeier95,Dessart05}, where the line-driven (or line-deshadowing) instability causes a small-scale, inhomogeneous wind in both density and velocity. \\textit{Direct observational} evidence of a small-scale, clumped stellar wind has, for O-stars, so far only been given for two objects, $\\zeta$ Pup and HD\\,93129A \\citep{Eversberg98,Lepine08}. Much \\textit{indirect} evidence, however, has arisen from quantitative spectroscopy, where the standard way of deriving mass-loss rates from observations nowadays is via line-blanketed, non-LTE (LTE: local thermodynamic equilibrium) model atmospheres that include a treatment of both the photosphere and the wind. Wind clumping has been included in such codes (e.g., CMFGEN \\citep{Hillier98}, PoWR \\citep{Grafener02}, FASTWIND \\citep{Puls05}) by assuming statistically distributed \\textit{optically thin} density clumps and a void inter-clump medium, while keeping the smooth velocity law. The major result from this methodology is that any mass-loss rate derived from smooth models and density-squared diagnostics ($\\rm H_{\\alpha}$, infra-red and radio emission) needs to be scaled down by the square root of the clumping factor (which describes the over density of the clumps as compared to the mean density, see Sect.~\\ref{wind_stoch}). For example, \\citet{Crowther02}, \\citet{Bouret03}, and \\citet{Bouret05} have concluded that a reduction of `smooth' mass-loss rates by factors $3 \\dots 7$ might be necessary. Furthermore, from a combined optical/IR/radio analysis of a sample of Galactic O-giants/supergiants, \\citet{Puls06} derived upper limits on observed rates that were factors of $2 \\dots 3$ lower than previous $\\rm H_{\\alpha}$ estimates based on a smooth wind. On the other hand, the strength of UV resonance lines (`P Cygni lines') in hot star winds depends linearly on the density and is therefore not believed to be directly affected by optically thin clumping. By using the Sobolev with exact integration technique (SEI; cf.~\\citeauthor{Lamers87} 1987) on the unsaturated Phosphorus V (PV) lines, \\citet{Fullerton06} for a large number of Galactic O-stars derived rates that were factors of $10 \\dots 100$ lower than corresponding smooth $\\rm H_{\\alpha}$/radio values (provided PV is the dominant ion in spectral classes O4 to O7). Such large revisions would conflict with the radiative line-driven wind theory and have dramatic consequences for the evolution of, and the feedback from, massive stars \\citep[cf.][]{Smith06,Hirschi08}. Indeed, a puzzling picture has emerged, and it appears necessary to ask whether the present treatment of wind clumping is sufficient. Particularly the assumptions of optically thin clumps, a void inter-clump medium, and a smooth velocity field may not be adequate to infer proper rates under certain conditions. \\paragraph{Optically thin vs. optically thick clumps.} \\citet{Oskinova07} used a porosity formalism \\citep{Feldmeier03,Owocki04} to scale the opacity from smooth models and investigate impacts from \\textit{optically thick} clumps on the line profiles of $\\zeta$ Pup. Due to a reduction in the effective opacity, the authors were able to reproduce the PV lines without relying on a (very) low mass-loss rate, while simultaneously fitting the optically thin $\\rm H_{\\alpha}$ line. This formalism, however, was criticized by \\citet{Owocki08} who argued that the original porosity concept had been developed for continuum processes, and that line transitions rather should depend on the non-monotonic velocity field seen in hydrodynamic simulations. Proposing a simplified analytic description to account for this velocity-porosity, or `vorosity', he showed how also this effect may reduce the effective opacity. In this first paper we attempt to clarify the most important concepts by conducting a detailed investigation on the synthesis of UV resonance lines from inhomogeneous two-dimensional (2D) winds. We create both pseudo-2D, radiation-hydrodynamic wind models and 2D, stochastic wind models, and produce synthetic line profiles via Monte-Carlo radiative transfer calculations. We account for and analyze the effects from a wind clumped in \\textit{both} density and velocity as well as the effects from a non-void inter-clump medium. Especially we focus on lines with intermediate line strengths, comparing the behavior of these lines with the behavior of both optically thin lines and saturated lines. Follow-up studies will include a treatment of emission lines (e.g., $\\rm H_{\\alpha}$) and an extension to 3D, and the development of simplified approaches to incorporate effects into non-LTE models. In Sect.~\\ref{wind} we describe the wind models and in Sect.~\\ref{rt} the Monte-Carlo radiative transfer code. First results from 2D inhomogeneous winds are presented in Sect.~\\ref{2d}, and an extensive parameter study is carried out in Sect.~\\ref{ps}. We discuss some aspects of the interpretations of these results and perform a first comparison to observations in Sect.~\\ref{Discussion}, and summarize our findings and outline future work in Sect.~\\ref{Conclusions}. ", "conclusions": "\\label{Discussion} \\subsection{The shapes of the intermediate lines} \\label{shapes} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=90]{fig9.ps}} \\caption{Total, absorption part, and re-emission part line profiles for 1D, smooth models with $\\kappa_0=5.0$ (dashed-dotted lines) and $\\kappa_0=5.0/(2\\eta)$ (solid lines, see Sect.~\\ref{shapes}), and for a 2D, stochastic model with density parameters as the default model and a $\\beta=1$ velocity law (dashed lines).} \\label{Fig:1d_fer} \\end{figure} For intermediate lines, the shape of the absorption part of the default model differs significantly from the shape of a smooth model (see Fig.~\\ref{Fig:stochs}, the middle plot in the lower-left panel). We showed in Sect.~\\ref{eta} that the shapes could be qualitatively understood by the behavior of $\\eta$. This is further demonstrated here by scaling the line strength parameter of a 1D, smooth model, using a parameterization $\\kappa_0 \\propto \\eta^{-1}$ outside the radius $r=1.3$ where clumping is assumed to start. Fig.~\\ref{Fig:1d_fer} displays the line profiles of 1D, smooth models with $\\kappa_0=5.0$ and $\\kappa_0=5.0/(2\\eta)$. These profiles are compared to those calculated from a `real' 2D stochastic model with density-clumping parameters as the default model, but with a $\\beta=1$ velocity field. $\\eta$ was calculated from Eq.~\\ref{Eq:heff}, using the parameters of the default model and a $\\beta=1$ velocity law, and the factor of 2 in the denominator of the scaled $\\kappa_0$ was chosen so that the \\textit{integrated} profile strength of the 2D model was roughly reproduced. From Fig.~\\ref{Fig:1d_fer} it is clear that the 1D model with scaled $\\kappa_0$ well reproduces the 2D results, indicating that indeed $\\eta$ governs the shape of the line profile. We notice also that these profiles display a completely black absorption dip in the outermost wind, as opposed to the default model with a non-monotonic velocity field (see Fig.~\\ref{Fig:stochs}, the middle plot in the lower-left panel). This is because the $\\beta$ velocity field does not allow for any clumps to overlap in velocity space (see the discussion in Sect.~\\ref{vel_par}), making the mapping of $\\eta$ almost perfect. Let us also point out that the line shapes can be somewhat altered by using a different velocity law, e.g., $\\beta \\ne 1$. Such a change would affect the distances between clumps as well as the Sobolev length, and thereby the line shapes of both absorption and re-emission profiles. However, in all cases is the shape of the re-emission part similar in the clumped and smooth models. \\subsection{The onset of clumping and the blue edge absorption dip} \\label{outer} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=90]{fig10.ps}} \\caption{\\textit{Upper panel:} Density structures of one slice in the default stochastic model (upper), in the default stochastic model with a modified $\\delta t$ (middle, see Sect.~\\ref{outer}), and in FPP (lower). \\textit{Lower panel:} Line profiles for the absorption part of an intermediate line for the default model (solid line), for the default model with a modified $\\delta t$ (dashed line), and for the default model with an ionization structure decreasing with increasing velocity (dashed-dotted line, see text).} \\label{Fig:holes} \\end{figure} We have used $r=1.3$ as the onset of wind clumping in our stochastic models, which roughly corresponds to the radius where significant structure has developed from the line-driven instability in our RH models. However, \\citet{Bouret03,Bouret05} analyzed O-stars in the Galaxy and the SMC, assuming optically thin clumps, and found that clumping starts deep in the wind, just above the sonic point. Also \\citet{Puls06} used the optically thin clumping approach, on $\\rho^2$-diagnostics, and found similar results, at least for O-stars with dense winds. With respect to our stochastic models, the qualitative results from Sects.~\\ref{2d}~and~\\ref{ps} remain valid when choosing an earlier onset of clumping. Quantitatively, the integrated absorption in intermediate lines becomes somewhat weaker, because the clumping now starts at lower velocities, and of course the line shapes in this region are affected as well. The onset of wind clumping will be important when comparing to observations, as discussed in Sect.~\\ref{cmp_obs}. The stochastic models that de-saturate an intermediate line generally display an absorption dip toward the blue edge (see Figs.~\\ref{Fig:stochs} and \\ref{Fig:1d_fer}), which has been interpreted in terms of low values of $\\eta$ in the outer wind (see Sect.~\\ref{eta}). However, this characteristic feature (not to be confused with the so-called DACs, discrete absorption components) is generally not observed, and one may ask whether it might be an artifact of our modeling technique. In the following we discuss two possibilities that may cause our models to overestimate the absorption in the outer wind; the ionization fraction and too low clump separations. Starting with the former, we have so far assumed a constant ionization factor, $q=1$ (cf. Eq.~\\ref{Eq:kappa0}). This is obviously an over-simplification. For example, an outwards decreasing $q$ would result in less absorption toward the blue edge. Here we merely demonstrate this general effect, parameterizing $q = v_0/v_{\\beta}$ in the stochastic default model (see Table~\\ref{Tab:mod}), with $v_{\\rm 0}=0.1$ the starting point below which $q=1$. Fig.~\\ref{Fig:holes} (lower panel, dashed-dotted lines) shows how the absorption in the outer wind becomes significantly reduced. The temperature structure of the wind is obviously important for the ionization balance. Whereas an isothermal wind is assumed in POF (see Sect.~\\ref{wind_rh}), the FPP model has shocked wind regions with temperatures of several million Kelvin. To roughly map corresponding effects on the line profiles, we re-calculated profiles based on FPP models assuming $q=0$ in all regions with temperatures higher than $T=10^5\\rm\\,K$, and $q=1$ elsewhere. Since the hot gas resides primarily in the low-density regions, however, the emergent profiles were barely affected, and particularly intermediate lines remained unchanged. On the other hand, the X-ray emission from hot stars (believed to originate in clump-clump collisions, see FPP) is known to be crucial for the ionization balance of highly ionized species such as C\\,IV, N\\,V, and O\\,VI \\citep[see, e.g., the discussion in][]{Puls08}. X-rays have not been included here, but could in principle have an impact on our line profiles, by illuminating the over-dense regions and thereby changing the ionization balance. \\citet{Krticka09}, however, find that incorporating X-rays does not influence the PV ionization significantly. Finally, non-LTE analyses including feedback from optically thin clumping have shown that this as well has a significant effect on the derived ionization fractions of, e.g., PV \\citep[]{Bouret05,Puls08b}. To summarize, it is clear that a full analysis of ionization fractions must await a future non-LTE application that includes relevant feedback effects from an inhomogeneous wind on the occupation numbers. In RH models, the average distance between clumps increases in the outer wind, due to clump-clump collisions and velocity stretching \\citep{Feldmeier97,Runacres02}. Neglecting the former effect, our stochastic models have clumps much more closely spaced in the outer wind\\footnote{The effect is minor in POF, since these RH models only extend to $r \\sim 5$ (see Sect.~\\ref{wind_rh}).}. We have therefore modified the default model by setting $\\delta t = 3$ outside a radius corresponding to $v_{\\beta}=0.7$. This is illustrated in the upper panel of Fig.~\\ref{Fig:holes}. The mass loss in the new stochastic model is preserved (because the clumps are more extended, see the figure), and this model now better resembles FPP. Recall that differences in the widths of the clumps are expected, since in the default model $f_{\\rm cl} \\approx f_{\\rm v}^{-1}=4$, whereas in FPP $f_{\\rm cl} \\approx 10$. The corresponding line profile shows how the absorption outside $x \\approx 0.7$ has been reduced, as expected from the higher $\\delta t$. \\subsection{The velocity spans of the clumps} \\label{vgrad} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=90]{fig11.ps}} \\caption{\\textit{Upper:} Velocity spans of density enhancements in the FPP model (squares) and corresponding $\\beta$ intervals (diamonds). \\textit{Lower:} Three density enhancements and corresponding velocity spans in the FPP model, highlighted as in Fig.~\\ref{Fig:contours}.} \\label{Fig:dvdvb} \\end{figure} In Sect.~\\ref{cmp} it was found that $|\\delta v| > \\delta v_{\\beta}$ in the RH models. Fig.~\\ref{Fig:dvdvb}, upper panel, shows the velocity spans of density enhancements (identified as having a density higher than the corresponding smooth value) in the FPP model, and demonstrates that, after structure has developed, $|\\delta v|$ is much higher than $\\delta v_{\\beta}$ throughout the whole wind. These high values essentially stem from the location of the starting points of the density enhancements, which generally lie \\textit{before} the velocities have reached their post shock values (see Fig.~\\ref{Fig:dvdvb}, middle and lower panels). By using a $\\beta$ velocity law (which in principle corresponds to a stochastic velocity law with $v_{\\rm j}=0$ and $\\delta v = \\delta v_{\\beta}$, see Fig.~\\ref{Fig:vj}) together with the density structure from FPP, we simulated a RH wind with low velocity spans. Indeed, for the corresponding intermediate line the equivalent width of the absorption part was $\\sim 35 \\%$ lower than that of the original FPP model. The strong line, on the other hand, remained saturated, because the ICM in FPP is not void. So, again, the RH models would in parallel display de-saturated intermediate lines and saturated strong lines, were it not for the large velocity spans inside the clumps. We suggest that the large velocity span inside a shell (clump) is primarily of kinematic origin, and reflects the formation history of the shell. The shell propagates outwards through the wind, essentially with a $\\beta=1$ velocity law \\citep{Owocki88}. Fast gas is decelerated in a strong reverse shock at the inner rim of the shell. The shell collects ever faster material on its way out through the wind. This new material collected at higher speeds resides on the star-facing side, i.e.~at smaller radii, of the slower material collected before. Thus, a negative velocity gradient develops inside the shell. The fact that $|\\delta v| \\gg \\delta v_{\\beta}$ in FPP seems to reflect that the shell is formed at small radii, and then advects outwards maintaining its steep interior velocity gradient\\footnote{Actually, the velocity gradient may further steepen during advection, due to faster gas trying to overtake slower gas ahead of it; however, this effect is balanced by pressure forces in the subsonic postshock domain.}. From this formation in the inner, steeply accelerating wind, velocity spans within the shells up to (a few) hundred $\\rm km\\,s^{-1}$, as seen in Fig.~\\ref{Fig:dvdvb}, appear reasonable. However, the dynamics of shell formation in hot star winds is very complex due to the creation and subsequent merging of subshells, as caused by nonlinear perturbation growth and the related excitation of harmonic overtones of the perturbation period at the wind base \\citep[see][]{Feldmeier95}. Future work is certainly needed to clarify to which extent the large velocity spans inside the shells in RH models are a stable feature (see also Sect.~\\ref{future}). \\subsection{3D effects} \\label{3d} A shortcoming of our analysis is the assumed symmetry in $\\Phi$. The 2D rather than 3D treatment has in part been motivated by computational reasons (see Appendix~\\ref{rt_code}). More importantly though, we do not expect our \\textit{qualitative} results to be strongly affected by an extension to 3D. Within the broken-shell wind model, all wind slices are treated independently, and distances between clumps increase only in the radial direction. Therefore the expected outcome from extending to 3D is a smoothing effect rather than a reduction or increase in integrated profile strength (similar to the smoothing introduced by $N_{\\Theta}$, see Sect.~\\ref{ang_dep}). Also, we have shown that the main effect from the inhomogeneous winds is on the absorption part of the line profiles (see, e.g., Sect.~\\ref{shapes}). The formation of this part is dominated by radial photons, especially in the outer wind, because of the dependence only on photons released directly from the photosphere. This implies that most photons stay within their wind slice, restricting the influence from any additional `holes' introduced by a broken symmetry in $\\Phi$ to the inner wind. Of course, these expectations hold only within the broken shell model, because in a real 3D wind the clumps will, for example, have velocity components also in the tangential directions. \\subsection{Comparison to other studies} \\label{oskow} To scale the smooth opacity in the formal integral of the non-LTE atmospheric code PoWR, \\citet{Oskinova07} used a porosity formalism in which both $f_{\\rm v}$ and the average distance between clumps enter. Other assumptions were a void ICM, a smooth $\\beta$ velocity field, and a microturbulent velocity $v_{\\rm t} \\approx 50 \\rm\\,km\\,s^{-1}$, the last identified as the velocity dispersion within a clump. However, a direct comparison between their study and ours is hampered by the different formalisms used for the spacing of the clumps. Here we have used the `broken-shell' wind model as a base (see Sect.~\\ref{wind_stoch}), in which each wind slice is treated independently and the distance between clumps increases only in the radial direction (clumps preserve their lateral angles). This gives a radial number density of clumps, $n_{\\rm cl} \\propto v^{-1}$, the same as used by, e.g., \\citet{Oskinova06}, when synthesizing X-ray emission from hot stars. In \\citet{Oskinova07}, on the other hand, the distance between clumps increases in \\textit{all} spatial directions. In a spherical expansion, this gives a radial number density of clumps $n_{\\rm cl} \\propto v^{-1}r^{-2}$, i.e., clumps are distributed much more sparsely within this model, especially in the outer wind. Therefore their choice of $L_{\\rm 0}=0.2$ is not directly comparable with $\\delta t =0.2$ in our models. The shapes of the clumps differ between the two models as well; in \\citeauthor{Oskinova07} clumps are assumed to be `cubes', whereas here the exact shapes of the clumps are determined by the values of the clumping parameters. Despite these differences, our findings confirm the qualitative results of \\citeauthor{Oskinova07} that the line profiles become weaker with an increasing distance between clumps as well as with a decreasing $v_{\\rm t}$. These results may be interpreted on the basis of the effective escape ratio, $\\eta$ (see Eq.~\\ref{Eq:heff}). Both a decrease in $v_{\\rm t}$ and an increase in the distance between clumps mean that the velocity span covered by a resonance zone becomes smaller when compared to the velocity gap between two clumps (see Fig.~\\ref{Fig:fer}, left panel), leading to higher probabilities for line photons to escape their resonance zones without interacting with the wind material. An important result of this paper is that models that de-saturate intermediate lines require a non-void ICM to saturate strong lines. This is confirmed by the \\citeauthor{Oskinova07} model, in which the ICM is void and strong lines indeed do not saturate \\citep{Hamann09}. \\citet{Owocki08} proposed a simplified description of the non-monotonic velocity field to account for vorosity, i.e., the velocity gaps between the clumps. Here, the vorosity effect has been discussed using the quantity $\\eta$ (see Sect.~\\ref{eta}), and we have introduced two new parameters to characterize a non-monotonic velocity field, $\\delta v$ and $v_{\\rm j}$. The reason for introducing a new parameterization is that when using a single velocity parameter, we have not been able to simultaneously meet the constraints from strong, intermediate, and weak lines as listed in Sect.~\\ref{2d}. Tests using a `velocity clumping factor' $f_{\\rm vel}=\\delta v / \\Delta v$ as proposed by \\citet{Owocki08}, together with a smooth density structure, have shown that this treatment indeed can reduce the line strengths of intermediate lines, but that the observational constraints from strong lines may not be met. Still, the basic concept of vorosity holds within our analysis. For example, one may phrase the high values of $\\delta v$ in the RH models in terms of insufficient vorosity. \\subsection{Comparison to observations} We finalize our discussion by performing a first comparison to observations. The two components of the Phosphorus V $\\lambda\\lambda$1118-1128 doublet are rather well separated, and the singlet treatment used here suffices to model the major part of the line complex. Nevertheless, the two components overlap within a certain region (indicated in Fig.~\\ref{Fig:cmp_obs}), so when interpreting the results of this subsection, one should bear in mind that the overlap is not properly accounted for, but treated as a simple multiplication of the two profiles. We used observed FUSE spectra (kindly provided by A. Fullerton) from HD\\,210839 ($\\lambda$ Cep), a supergiant of spectral type O6\\,I(n)fp. When computing synthetic spectra, we first assumed optically thin clumping with a constant clumping factor $f_{\\rm cl}=9$ and a smooth $\\beta=1$ velocity field. $f_{\\rm cl}=9$ agrees fairly well with the analysis of \\citet{Puls06}, who derived clumping factors $f_{\\rm cl}=6.5$ for $r \\approx 1.2 \\dots 4.0$ and $f_{\\rm cl}=10$ for $r \\approx 4.0 \\dots 15$, assuming an un-clumped outermost wind.\\footnote{This stratification has been found to be prototypical for O-supergiants and was, together with its well developed PV P Cygni profiles, the major reason for choosing $\\lambda$~Cep as comparison object instead of, e.g., $\\zeta$~Pup, which displays a somewhat unusual run of $f_{\\rm cl}$.} We took the ionization fraction $q=q(r)$ of PV from \\citet{Puls08b}, calculated with the unified non-LTE atmosphere code FASTWIND for an O6 supergiant, using the Phosphorus model atom from \\citet{Pauldrach01}. The feedback from optically {\\it thin} clumping was accounted for and X-rays were neglected. This ionization fraction was then used as input in our MC-1D code when computing the synthetic spectra. We assigned a thermal plus a highly supersonic `microturbulent' velocity $v_{\\rm t}=0.05$ (corresponding to 110 km\\,s$^{-1}$), as is conventional in this approach. The mass-loss rate was derived using the well known relation between $\\kappa_0$ and $\\dot{M}$ \\citep[e.g.,][]{Puls08}. For atomic and stellar parameters, we adopted the same values as in \\citet{Fullerton06}. The dashed line in Fig.~\\ref{Fig:cmp_obs} represents our fit to the observed spectrum, assuming optically thin clumping, resulting in a mass-loss rate $\\dot{M}=0.24$, in units of $10^{-6}\\rm\\,M_{\\odot}\\,yr^{-1}$. \\citet{Fullerton06} derived $\\langle q \\rangle \\dot{M} = 0.23$ for this star. Because our clumped FASTWIND model predicts an averaged ionization fraction $\\langle q \\rangle \\approx 0.9$ in the velocity regions utilized by \\citeauthor{Fullerton06}, the two rates are in excellent agreement. On the other hand, \\citet{Repolust04} for HD\\,210839 derived $\\dot{M}=6.9$ from $\\rm H_{\\alpha}$ assuming an unclumped wind, yielding $\\dot{M}_{\\rm H_{\\alpha}}=2.3$ when accounting for the reduction implied by our assumed $f_{\\rm cl}=9$ ($\\dot{M}_{\\rm H \\alpha}=\\dot{M}_{\\rm H \\alpha,sm}f_{\\rm cl}^{-1/2}$). This rate is almost ten times higher than that inferred from PV, and thus results in PV line profiles that are much too strong (see Fig.~\\ref{Fig:cmp_obs}, dashed-dotted line). That is, to reconcile the $\\rm H_{\\alpha}$ and PV rates for HD\\,210839 with models that assume optically {\\it thin} clumps also in PV, we would have to raise the clumping factor to $f_{\\rm cl} > 100$. In addition to this very high clumping factor, the low rate inferred from the PV lines conflicts with the theoretical value $\\dot{M}=3.2$ provided by the mass-loss recipe in \\citet{Vink00} \\citep[using the stellar parameters of][]{Repolust04}, and is also strongly disfavored by current massive star evolutionary models \\citep{Hirschi08}. Next we modeled the PV lines using our MC-2D code together with a stochastic 2D wind model. The same clumping factor ($f_{\\rm cl}=9$) and ionization fraction (calculated from FASTWIND, see above) were used. This time, we assigned $v_{\\rm t}=0.005$, i.e., applied no microturbulence. In previous sections, e.g. \\ref{st} and \\ref{shapes}, we showed that stochastic models generally display a line shape different from smooth models, with a characteristic absorption dip at the blue edge as well as a dip close to the line center. Such shapes are not seen in the PV lines in $\\lambda$ Cep. Thus, to better resemble the observed line shapes, we used different values for $\\delta t$ and $\\fic$ in the inner and outer wind (the former modification already discussed in Sect.~\\ref{outer}) and let clumping start close to the wind base. Clumping parameters are given in Table~\\ref{Tab:mod}, model Obs1. As illustrated in Fig.~\\ref{Fig:cmp_obs}, the synthetic line profiles using $\\dot{M}=2.3$, as inferred from $\\rm H_{\\alpha}$, are now at the observed levels. Because of our insufficient treatment of line overlap, we gave higher weight to the $\\lambda$1118 component when performing the fitting, but the profile-strength ratio between the blue and red component was nevertheless reasonably well reproduced (see also discussion in Sect.~\\ref{dens_par}). However, though the fit appears quite good, we did not aim for a perfect one, and must remember the deficits of our modeling technique. For example, while the early onset of clumping definitely improved the fit (using our default value, there was a dip close to line center) and might be considered as additional evidence that clumping starts close to the wind base, the same effect could in principle be produced by non-LTE effects close to the photosphere or by varying the underlying $\\beta$ velocity law. Such effects will be thoroughly investigated in a follow-up paper, which will also include a comparison to observations from many more objects. Clearly, a consistent modeling of resonance lines (at least of intermediate strengths) requires the consideration of a much larger parameter set than if modeling via the standard diagnostics assuming optically thin clumping, and a reasonable fit to a single observed line complex can be obtained using a variety of different parameter combinations. The analysis of PV lines as done here can therefore, at present, only be considered as a consistency check for mass-loss rates derived from other, independent diagnostics, and not as a tool for directly estimating mass-loss rates. Additional insight might be gained by exploiting more resonance doublets, due to the different reactions of profile strengths and shapes on $\\kappa_0$. The different slopes of the equivalent width as a function of $\\kappa_0$ in smooth and clumped models, especially at intermediate line strengths (Sect.~\\ref{dens_par}), may turn out to be decisive. However, because of, e.g., the additional impact from the ICM density, also this diagnostics requires additional information from saturated lines. Taken together, only a consistent analysis using different diagnostics and wavelength bands, and embedded in a suitable non-LTE environment, will (hopefully) provide a unique view. \\label{cmp_obs} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics[angle=90]{fig12.ps}} \\caption{Observed FUSE spectra of the PV doublet $\\lambda\\lambda$1118-1128 for the O6 supergiant $\\lambda$ Cep \\citep{Fullerton06}. The synthetic spectra are calculated for two 1D models assuming optically thin clumping (see Sect.~\\ref{cmp_obs}) and for one 2D stochastic model with parameters as in Table~\\ref{Tab:mod}, model Obs1. The models have mass-loss rates $\\dot{M}\\,[\\rm M_{\\odot}\\,yr^{-1}]$ as given in the figure. The zero point frequency is shifted to the line center of the $\\lambda$1118 component, and the two arrows at the bottom of the figure indicate in which region the two components overlap.} \\label{Fig:cmp_obs} \\end{figure}" }, "0911/0911.1014_arXiv.txt": { "abstract": "We present our optical observations of {\\em Swift} GRB 070518 afterglow obtained at the 0.8-m Tsinghua University-National Astronomical Observatory of China telescope (TNT) at Xinglong Observatory. Our follow-up observations were performed from 512 sec after the burst trigger. With the upper limit of redshift$\\sim$0.7, GRB 070518 is found to be an optically dim burst. The spectra indices $\\beta_{ox}$ of optical to X-ray are slightly larger than 0.5, which implies the burst might be a dark burst. The extinction $A_{V}$ of the host galaxy is 3.2 mag inferred from the X-ray hydrogen column density with Galactic extinction law, and 0.3 mag with SMC extinction law. Also, it is similar to three other low-redshift optically dim bursts,which belong to XRR or XRF, and mid-term duration($T_{90}<10$, except for GRB 070419A, $T_{90}$=116s). Moreover, its $R$ band afterglow flux is well fitted by a single power-law with an index of 0.87. The optical afterglow and the X-ray afterglow in the normal segment might have the same mechanism, as they are consistent with the prediction of the classical external shock model. Besides, GRB 070518 agrees with Amati relation under reasonable assumptions. The Ghirlanda relation is also tested with the burst. ", "introduction": "Since the first optical counterpart of gamma-ray burst (GRB) was identified on 1997 February 28, more and more GRBs have been optically detected, especially after {\\em Swift} was launched successfully, which allows early follow-up observations with ground-based optical telescopes. However, about 40 per cent of {\\em Swift} GRBs still failed to be identified with optical or infrared counterparts (Burrows et al. 2008). For the nature of the dark burst, several models have been proposed: extinction from dusty opaque star-forming regions ( e.g. Groot et al. 1998; Reichart \\& Price 2002); high redshift, which diminishes the GRB optical luminosity (Lamb \\& Reichart 2000), as the ly$\\alpha$ forest is shifted into the observational energy range; intrinsically dim bursts (Fynbo et al. 2001; Rol et al. 2005 ); the observational effect, As a result of the late response to the bursts. Jakobsson et al. (2004) have proposed an operational definition of dark burst (i.e. optical-to-X-ray spectral index $\\beta_{\\mathrm{ox}}<0.5$), according to the fireball model. However, GRBs with an optical afterglow have been studied widely and in depth(e.g. GRB 060218, GRB 080319B). This has resulted in an increasing number of GRBs with known redshift, which has allowed statistical studies to be carried out. The optical luminosity distribution of GRBs varies widely up to several magnitudes (Kann et al. 2006; 2008). After being transformed into a common redshift (e.g., $z=1$), the optical luminosity shows a bi-model phenomenon (Nardini et al. 2006; Liang et al. 2006; Kann et al. 2006; 2008): optically luminous and optically dim. The bi-model distribution is also found to exist in other energy bands (Gendre et al. 2008a,b). Liang \\& Zhang (2006) found that the apparent bimodality cannot be interpreted as a manifestation of the extinction effect, and they suggested that there might be two types of progenitors or two types of explosion mechanisms in operation. Considering the dark burst is usually a type of GRBs with a faint optical afterglow, optically dim bursts might be related to dark bursts (Nardini et al. 2006). However, the number of well-studied optically dim bursts is very limited. Thus, increasing the number of known optically dim bursts may help us to understand the origin of the bi-model distribution, the nature of optically dim bursts and dark bursts. One such interesting case is GRB 070518. It was detected by Gamma-ray Burst Alert Telescope (BAT) onboard {\\em Swift} at 14:26:21 (UT) on 2007 May 18. The X-ray telescope (XRT) and ultraviolent/optical telescope (UVOT) on board {\\em Swift} observed the counterpart beginning at 70 and 100 sec after the trigger, respectively. T$_{90}$ (15-350 KeV) is 5.5$\\pm$0.2 s ,T$_{50}$ is 2.9s (Sakamoto et al. 2008a). The values of duration mean the burst belongs to the short tail of the long group (Kouveliotou et al. 1993). Thus, it might be a classical long-duration GRB or an intermediate-long GRB (Horvath et al. 2008). The ratio of fluence between the 25-50 KeV and 50-100 KeV band (Sakamoto et al. 2008a) is about 1.06, which makes GRB 070518 belong to an X-ray rich burst (XRR), according to the criterion \\footnote{S(25-50KeV)/S(50-100KeV)$=<$0.72 C-GRB; \\\\ 0.72$<$S(25-50KeV)/S(50-100KeV)$=<$1.31 XRR;\\\\ S(25-50KeV)/S(50-100KeV)$>$1.32 XRF.} given by Sakamoto et al. (2008b). The first given magnitude and coordinates of the counterpart are about 18 mag in $white$ band reported by UVOT and RA(J2000) = 16h 56m 47.7s, Dec(J2000) = 55d 17m 42.3s (radius, 90 pen cent containment), respectively (Guidorzi et al. 2007). The afterglow was also observed by other ground-based telescopes later. Besides, the redshift of GRB 070518 was reported to be lower than 0.7 (Cucchiara et al. 2007) based on the photometry of the {\\em Swift} UVOT. No other report about redshift of GRB 070518 has been presented in the literature up to now. In our analysis, the redshift of 0.7 are used to investigate the properties of GRB 070518. In this paper, we report on optical photometric follow-up observations of GRB 070518 at Xinglong Observatory of National Astronomical Observatories, Chinese Academy of Sciences. Afterglow observations and results are described in $\\S$2. An anaysis and discussion are presented in $\\S$3, and a summary and conclusion are given in $\\S$4. In our calculations, a flat universe is assumed with matter density $\\Omega_M = 0.3$, cosmological constant $\\Omega_\\Lambda=0.7$, and Hubble constant $H_0=70$ km s$^{-1}$ Mpc$^{-1}$. The formula $f\\propto$$t^{-\\alpha}\\nu^{-\\beta}$ is used for the afterglow decay analysis. ", "conclusions": "We have presented optical afterglow observations of GRB 070518 from the TNT and an analysis of the optical and X-ray light curves. The optical afterglow shows a constant power law decay $\\sim$0.87, while the X-ray shows a \"steep-shallow-normal\" decay behavior. Based on the upper limit of redshift of 0.7, the optical luminosity at 12 hours in the rest frame was estimated to be $2\\times10^{28}$ erg s$^{-1}$, which means that the burst is an optically dim burst. The conclusion is also supported by comparing with three other optically dim bursts: GRB 050416A, GRB 060512 and GRB 070419A. The spectral indices of $\\beta_{\\mathrm{ox}}$ at 1000 s, 12 hours,10$^{5}s$ are calculated. All $\\beta_{\\mathrm{ox}}$ were slightly larger than 0.5, indicating that GRB 070518 might be a dark burst or a gray burst, according to the definition of Jakobsson et al. (2004). The optical extinction in its host galaxy inferred from X-ray hydrogen column density is 3.2 with Galactic extinction law, and 0.3 with SMC extinction law. Afterglow model has been applied to the light curves. Energy injection is found to occurr for the X-ray afterglow before the start of the normal decay segment. The behavior of the normal segment of the X-ray afterglow and the optical afterglow is consistent with the prediction of the classical external shock model (Urata et al. 2007), indicating that they come from the same origin. As an XRR burst, GRB 070518 agrees with Amati relation. Moreover, it fills the gape of high-luminosity and low-luminosity bursts in the Amati relation, similar to GRB 050416A (Sakamoto et al. 2005). It is also found that GRB 070518 agrees with the Ghalanda relation within reasonable assumptions. Moreover, a comparison with three optically dim bursts (GRB 050416A, GRB 060512 and GRB 070419A and GRB 070518) reveals that several similar properties were shared: burst duration T$_{90}$$<$10s (expect for GRB 070419A, $T_{90}=116s$), soft spectrum at BAT band $\\Gamma$$>2$, low redshift $z<$1, etc." }, "0911/0911.5401_arXiv.txt": { "abstract": "A general equation of state is used to model unified dark matter and dark energy (dark fluid), and it has been proved that this model is equivalent to a single fluid with time-dependent bulk viscosity. In this paper, we investigate scalar perturbation of this viscosity dark fluid model. For particular parameter selection, we find that perturbation quantity can be obtained exactly in the future universe. We numerically solve the perturbation evolution equations, and compare the results with those of $\\Lambda$CDM model. Gravitational potential and the density perturbation of the model studied here have the similar behavior with the standard model, though there exists significant value differences in the late universe. ", "introduction": "Astrophysics and cosmology observations in recent years delineate the cosmological picture on its constituents more and more accurately, i.e, the precision cosmology era comes. Except for the long standing puzzling dark matter component, an unknown cosmic matter-energy constituent refereed to the so called dark energy may also exist that accelerates our universe expansion now, which contradicts with our traditional attitude on the behavior of conventional matter but it is eventually confirmed by recent observations like SNe Ia \\cite{sn1} \\cite{sn2} and CMB observations \\cite{CMB}. The cosmological dark sector, often divided as the mysterious DM and DE sectors respectively, takes around $95\\%$ of total energy budget of our universe. The concord $\\Lambda$CDM model could be consistent with most of global astrophysics observational results. But the introduction of the cosmological constant simultaneously results in the yet to answer problems related directly to how to understand the fundamental physics theory, like the fine-tuning and the coincidence problems respectively. At the same time, ``most'' of course does not equal to ``all'', some astrophysics problems still need be clarified and solved in the framework of the $\\Lambda$CDM model \\cite{six}, such as the the core singularities of the cold matter halo profiles. With the aim to understand the cosmic acceleration or dark energy phenomena, many theoretical models have been proposed, like the scalar field models and the modified Einstein gravity models \\cite{r1} \\cite{r2} \\cite{r3}. Due to the limited scope of our experimental and observational tools, we have not yet been able to understand the ``dark'' nature and to detect the origin of DM and DE. Purely gravitational probes can not provide enough information to differentiate these two kinds of mysterious constitutions either. Therefore, from the phenomenological and practical point of view, a single (unified DM with DE) fluid description may be more plausible at least in the cosmic evolution description, which utilizes a single equation of state to model the dark matter and dark energy contributions together \\cite{df1} \\cite{df2} \\cite{df3} \\cite{df4} \\cite{df5} \\cite{Liddle}. Generally, such models has a non-constant equation of state (EoS), which reflects both its dynamical and thermodynamics characters. The density dependent equation of state is widely investigated, such as the famous Chaplygin gas and generalized Chaplygin gas models, which assume an EoS form like the $p=-A/\\rho^{\\alpha}$ where the $\\alpha$ is a model parameter. Another practical method to modify the EoS is by the introduction of cosmic viscosity media contribution which replaces the simplest perfect fluid EoS on a more physical and realistic basis. In the homogeneous and isotropic Friedmann-Robertson-Walker frame, only a bulk viscosity term which behaves as an additive pressure contribution can mimic both the two dark components and their coupling effects by playing a main role to influence the cosmic evolution. Different forms of the viscosity coefficients have been proposed like the density $\\rho$ dependent \\cite{d} or the redshift $z$ dependent \\cite{md}. In this letter, we will proceed our effort on the investigation of the viscosity dark fluid model to study the scalar perturbation of this model, for perturbation analysis can provide us a powerful tool to differentiate and constrain cosmology models finely as the calculation of perturbation quantities links the theoretical models with more plentiful and precise observations, like the cosmic microwave background (CMB) and large scale structure observation. There have been some researches on the perturbation evolution of the viscosity models \\cite{FGR} \\cite{lb} \\cite{z}, by which we know that after corresponding model parameters chosen properly, the Chaplygin gas formulation can be viewed as a special case of the density dependent viscosity model. In the non-perturbative (zero order) level, the Chaplygin gas model can be exactly solved and fit the observational data well. But it has been found that in the perturbation level, there exist some unacceptable behaviors, like the blow up of density perturbation evolution and other peculiar behaviors \\cite{z} \\cite{ps}. One motivation to build other kinds of the viscosity models is to overcome these difficulties the Chaplygin gas models possess. Here, we will consider a time-dependent viscosity coefficient model, which is equivalent to the introduction of a general Equation of state(EoS) \\cite{RM}. The general EoS is \\begin{equation}\\label{1} p=(\\gamma-1)\\rho+p_{0}+w_{H}H+w_{H2}H^{2}+w_{dH}\\dot H. \\end{equation} In the background, this model can fit the current astrophysics observational datasets consistently. We derive its perturbation equations that govern the evolution of gravitational potential and density perturbation below. We numerically solve the perturbation equation, and compare it with that of conventional $\\Lambda$CDM model and the Chaplygin gas model finding that the dark fluid model behaves well in different scales. Though there exists some value difference between the $\\Lambda$CDM model and the dark fluid model in the late time evolution, their gravitational potential and density contrast shape and evolution behavior are similar by plotting respectively. This paper is organized as follows: In Sec. \\textbf{II}, we summarize the calculations of scalar perturbation, and give the general evolution equation of the gravitational potential. In Sec. \\textbf{III}, we briefly review the background evolution of the dark fluid model. In Sec. \\textbf{IV}, we discuss the perturbation evolution of the dark fluid model. In Sec. \\textbf{V}, we numerically solve the perturbation equation and compare it with other models. Finally, we present the conclusions in the last section. ", "conclusions": "In this paper, we investigate extensively the dark fluid model proposed in \\cite{RM}, equivalently this model can be viewed as a single fluid with time-dependent bulk viscosity. Scale factor and density evolution can be exactly solved in this model. In the background, the dark fluid model can fit the supernova data acceptable. Our main task in this paper is to analysis the behavior of this model in the perturbation level. We derive equations govern the perturbation quantities. For the condition that $T_{1}$ is smaller than $0$, the universe will enter de Sitter phase in the $t\\rightarrow\\infty$ future. We solve exactly the gravitational equation in this condition and obtain the solution for both long and short wave case. Generally, the perturbation evolution equations are solved numerically. When compare the results with those of $\\Lambda$CDM model, we find that \\begin{itemize} \\item In the early time and the large scale, both the gravitational potentials of two models behave as a constant. \\item Though the gravitational potentials of two models have similar behavior and shape, as can be seen form FIG. ~3, there exists about $5\\%-10\\%$ significant value difference in the late time. \\end{itemize} Perturbation analysis also provides constraint on model parameter. For different selection of parameter $\\tilde\\gamma$, both $\\tilde\\gamma<0$ and $\\tilde\\gamma>0$ can give consistent prediction curve of distance modulus, but numerical results indicate in $\\tilde\\gamma<0$ case, the gravitational potential deviates from $\\Lambda$CDM significantly from the early time. This result strongly constrain the selection region of $\\tilde\\gamma$. We suggest $\\tilde\\gamma>0$, which can also produce positive sound speed naturally." }, "0911/0911.2367_arXiv.txt": { "abstract": "The properties of galactic cosmic rays are investigated with the KASCADE-Grande experiment in the energy range between $10^{14}$ and $10^{18}$~eV. Recent results are discussed. They concern mainly the all-particle energy spectrum and the elemental composition of cosmic rays. ", "introduction": "To reveal the origin of galactic cosmic rays is the main objective of the KASCADE and KASCADE-Grande experiments. The results of KASCADE contributed significantly to the understanding of the origin of the knee in the all-particle energy spectrum of cosmic rays at energies around $4\\cdot10^{15}$~eV. It could be shown that the knee is caused by a fall-off in the flux of light nuclei \\cite{ulrichapp,Apel:2008cd}. Energy spectra for five dominant elemental groups could be reconstructed (p, He, CNO, Si, and Fe), they exhibit a fall-off behavior approximately proportional to the nuclear charge of the elemental groups. In the energy region between $10^{17}$ and $10^{18}$~eV a transition from a galactic to an extra-galactic origin of cosmic rays is expected \\cite{behreview}. Main focus of KASCADE-Grande is to understand the end of the galactic cosmic-ray component through detailed investigations of the energy spectrum and the mass composition in this energy region. KASCADE-Grande comprises 37 detector stations equipped with plastic scintillators, covering an area of 0.5~km$^2$ to measure the electromagnetic shower component \\cite{grande}. It also includes the original KASCADE experiment, consisting of several detector systems \\cite{kascadenim}. A $200 \\times 200$~m$^2$ array of 252 detector stations, equipped with scintillation counters, measures the electromagnetic and, below a lead/iron shielding, the muonic parts of air showers. A 130~m$^2$ streamer tube detector, shielded by a soil-iron absorber serves to reconstruct the tracks of high-energy ($E_\\mu>0.8$~GeV) muons \\cite{mtdnim}. An iron sampling calorimeter of $16 \\times 20$~m$^2$ area detects hadronic particles \\cite{kalonim}. It has been calibrated with a test beam at the SPS at CERN up to 350~GeV particle energy \\cite{kalocern}. \\begin{figure}\\centering \\epsfig{file=dipierro-lat.eps, width=\\columnwidth} \\caption{Measured lateral distribution of a single event for charged particles and muons \\cite{icrc09-dipierro}.} \\label{lat} \\end{figure} As an example for a measured air shower, the lateral distributions for charged particles and for muons are shown in \\fref{lat} \\cite{icrc09-dipierro}. The charged-particle density is measured with the Grande detectors and the muonic component by the shielded array detectors of KASCADE. ", "conclusions": "The recent results from the KASCADE-Grande experiment indicate that we make substantial progress in measuring the energy spectrum and the elemental composition of cosmic rays in the energy region between $10^{16}$ and $10^{18}$~eV. The new data will significantly contribute to the understanding of the end of the galactic component and the transition to an extra-galactic component. The data will be crucial to distinguish between different astrophysical scenarios." }, "0911/0911.3588_arXiv.txt": { "abstract": "We present a photometric study of $I$-band variability in the young association Cepheus OB3b. The study is sensitive to periodic variability on timescales of less than a day, to more than 20 days. After rejection of contaminating objects using $V$, $I$, $R$ and narrowband H$\\alpha$ photometry, we find 475 objects with measured rotation periods, which are very likely pre-main-sequence members of the Cep OB3b star forming region. We revise the distance and age to Cep OB3b, putting it on the self-consistent age and distance ladder of \\cite{2008MNRAS.386..261M}. This yields a distance modulus of 8.8$\\pm$0.2 mags, corresponding to a distance of 580$\\pm$60 pc, and an age of 4-5Myrs. The rotation period distribution confirms the general picture of rotational evolution in young stars, exhibiting both the correlation between accretion (determined in this case through narrowband H$\\alpha$ photometry) and rotation expected from disc locking, and the dependence of rotation upon mass that is seen in other star forming regions. However, this mass dependence is much weaker in our data than found in other studies. Comparison to the similarly aged NGC 2362 shows that the low-mass stars in Cep OB3b are rotating much more slowly. This points to a possible link between star forming environment and rotation properties. Such a link would call into question models of stellar angular momentum evolution, which assume that the rotational period distributions of young clusters and associations can be assembled into an evolutionary sequence, thus ignoring environmental effects. ", "introduction": "\\label{sec:introduction} There are sound theoretical reasons to expect that accretion processes help determine the angular momentum of T-Tauri stars. Historically, the slow rotation rates of T Tauri stars (relative to their break-up velocity) has been explained by the disc-locking theory \\citep{konigl91,shu94}. In this theory, magnetic field lines connect the star to the disc, enforcing synchronous rotation between the star and the material in the disc at some radius, near where the magnetic field disrupts the disc. The simplistic theory has been expanded in a variety of models where angular momentum is removed from the star by a combination of the disc and an accretion-driven wind \\citep[e.g.][]{fendt07,romanova07,matt08}. Whether the star is spun up or down by the star-disc interaction depends upon the balance between accretion spinning up the star and magnetic (or wind) torques slowing it down. Theoretically, this issue is unresolved; some studies find the star spins up \\citep{bessolaz08}, some find it spins down \\citep[e.g.][]{long05}. Observationally, the evidence for the influence of accretion disks on rotation is much stronger than it was a few years ago \\citep{herbst07}. Previously, conflicting results had arisen \\citep[e.g.][] {herbst02,stassun99,littlefair05}. These were most likely due to a combination of small sample sizes, and ambiguous diagnostics of the presence of accretion disks. \\cite{rebull06} resolved these issues with a large sample of rotation periods in the ONC, and accretion disk status defined from Spitzer IRAC data. \\cite{rebull06} found a clear correlation between mid-IR excess and rotation, in the sense that stars with mid-IR excess were much more likely to be slow-rotators. An analysis of the slightly older NGC 2264 found the same result \\citep{cieza07}, and confirmed the \\cite{rebull06} result, via a refined analysis of the same data. There is thus strong observational evidence that the star-disc interaction is responsible for extracting angular momentum from young stars. Interestingly, a small population of rapidly rotating stars with mid-IR excesses exists; it is possible that these stars are being spun-up by accretion, hinting that the braking process of young stars is intermittent. Also, in both the ONC and NGC 2264, there exists a significant population of slow rotators with {\\em no} mid-IR excess. These are often interpreted as being recently released from disc-locking, but there are problems with this interpretation \\citep[see][for example]{bouvier07}. The firm link established between accretion and rotation represents significant progress, but there are still open questions regarding the rotation of young stars. For example, the low-mass stars are rotating more rapidly than the high-mass stars, and appear to spin-up more rapidly as they contract towards the main sequence \\citep{herbst02,irwin07a}. The reason for this is still not known. Also, the angular momentum evolution of young stars is determined by assembling different clusters into an evolutionary sequence, and assuming the period distribution of the older clusters can be modelled using the young clusters as a starting point. In doing so, the possibility of an environmental effect on rotation is ignored. Such an environmental effect may be indicated by the data; the young stars in IC348 rotate much slower than those in the similarly aged NGC 2264 \\citep{littlefair05}. Here we present a photometric study of the young association Cepheus OB3b (hereafter Cep OB3b). Cep OB3 is a young association covering a region of the sky from approximately 22$^h$46$^m$ to 23$^h$10$^m$ in right ascension and +61$^{\\circ}$ to +64$^{\\circ}$ in declination. The subgroup Cep OB3b lies closest to the molecular cloud, and has a rich pre-main-sequence (PMS) population, confirmed by both spectroscopy \\citep{pozzo03} and X- ray data \\citep{getman06}. In section~\\ref{sec:obs} we present the observations and data reduction techniques applied. Section~\\ref{sec:detect} describes the techniques used to identify periodic variables. Section~\\ref{sec:distance} provides a revision of the age of, and distance to, Cep OB3b. In Sections~\\ref{sec:results} and \\ref{sec:discussion} we present our results and analysis of the data, whilst in section~\\ref{sec:concl} we draw our conclusions. ", "conclusions": "\\label{sec:concl} We present a photometric study of $I$ band variability towards the young association Cepheus OB3b. The study is sensitive to periodic variability on timescales of less than a day, to more than 20 days. The result is a database of 704 periodic variables in the field of Cep OB3b. A random inspection of 200 of these objects suggest that around 97 per cent of these periods are genuine. Colour cuts using $V$, $I$, $R$ and narrowband H$\\alpha$ photometry are used to reject contaminating objects, leaving 475 objects with measured rotation periods, which are very likely pre-main-sequence members of the Cep OB3b star forming region. We revise the distance and age to Cep OB3b, putting it on the consistent age and distance ladder of \\cite{2008MNRAS.386..261M}. This yields a distance modulus of 8.8$\\pm$0.2 mags, corresponding to a distance of 580$\\pm$60 pc, and an age of 4-5Myrs. For the purposes of this paper we therefore adopt an age of 4.5Myr. The rotation period distribution confirms the general picture of rotational evolution in young stars, exhibiting both the correlation between accretion and rotation expected from disc locking, and the dependence of rotation upon mass that is seen in other star forming regions. However, this mass dependence is much weaker than seen in other regions. Comparison to the similarly aged NGC 2362 shows that the low-mass stars in Cep OB3b are rotating much more slowly. This points to a possible link between star forming environment and rotation properties. Such a link would call into question models of stellar angular momentum evolution, which assume that associations can be assembled into an evolutionary sequence, thus ignoring environmental effects." }, "0911/0911.1258_arXiv.txt": { "abstract": "We present spectra of high-redshift supernovae (SNe) that were taken with the Subaru low resolution optical spectrograph, FOCAS. These SNe were found in SN surveys with Suprime-Cam on Subaru, the CFH12k camera on the Canada-France-Hawaii Telescope (CFHT), and the Advanced Camera for Surveys (ACS) on the Hubble Space Telescope (HST). These SN surveys specifically targeted $z>1$ Type~Ia supernovae (SNe~Ia). From the spectra of 39 candidates, we obtain redshifts for 32 candidates and spectroscopically identify 7 active candidates as probable SNe~Ia, including one at $z=1.35$, which is the most distant SN~Ia to be spectroscopically confirmed with a ground-based telescope. An additional 4 candidates are identified as likely SNe~Ia from the spectrophotometric properties of their host galaxies. Seven candidates are not SNe~Ia, either being SNe of another type or active galactic nuclei. When SNe~Ia are observed within a week of maximum light, we find that we can spectroscopically identify most of them up to $z=1.1$. Beyond this redshift, very few candidates were spectroscopically identified as SNe~Ia. The current generation of super red-sensitive, fringe-free CCDs will push this redshift limit higher. ", "introduction": "\\label{sec:introduction} Type~Ia supernovae (SNe~Ia) have proven to be very good standard candles for cosmological studies. They are bright enough to detect at cosmological distances, $z\\sim1.5$, if one uses 8--10~m class ground-based optical telescopes or the Hubble Space Telescope (HST), and their luminosities can be standardised using empirical relations between luminosity and light curve shape (\\cite{phillips1993}) and colour \\citep{tripp1998}. SNe~Ia have played a leading role in measuring the expansion history of the universe since two independent teams, the Supernova Cosmology Project (SCP) and the High-$z$ Team, discovered the accelerating expansion of the universe (\\cite{perlmutter1999}; \\cite{riess1998}). Since then, many projects to discover and identify SNe~Ia have been organized. For example, the Carnegie Supernova Project (CSP; \\cite{hamuy2006}), the Nearby Supernova Factory (SNfactory; \\cite{aldering2002}), the Harvard-Smithsonian Center for Astrophysics (CfA) supernova survey (\\cite{jha2006}; \\cite{hicken2009}), the Sloan Digital Sky Survey-II Supernova Survey (SDSS-II SN Survey; \\cite{sako2008}; \\cite{frieman2008}), the Supernova Legacy Survey (SNLS; \\cite{astier2006}), the Equation of State: SupErNovae trace Cosmic Expansion (ESSENCE) survey (\\cite{miknaitis2007,woodvasey2007}), and Higher-Z team (\\cite{riess2007}) have detected around 1000 SNe up to redshift $z\\sim 1.5$. The combination of SN~Ia data with measurements of cosmic microwave background (CMB) fluctuations (\\cite{spergel2003}; \\cite{spergel2007}; \\cite{komatsu2009}), baryon acoustic oscillations (BAO; \\cite{eisenstein2005}) and galaxy cluster number counts (\\cite{vikhlinin2009}) have constrained the dark energy equation of state parameter. However, the limits are consistent with very different dark energy models, so the fundamental nature of dark energy remains unclear. It is currently one of the biggest mysteries in physics, and combined astronomical observations (SNe~Ia, CMB, BAO and weak lensing) seem to be the only way to constrain its properties. Since the discovery of the accelerating expansion of the universe, the SCP has been carrying out imaging surveys for SNe~Ia at $z\\gtrsim 1$, an epoch during which the expansion of the universe is expected to be decelerating. Spectroscopic follow-up observations are an essential part of these surveys, providing spectroscopically determined redshifts and, when necessary and possible, direct confirmation of the SN type. In this paper, we present spectra of SNe and their host galaxies taken with the Faint Object Camera And Spectrograph (FOCAS; \\cite{kashikawa2002}) on the Subaru 8.2-m telescope. Twelve SN candidates shown here were found in ground-based observations targeting blank fields and nearby galaxy clusters with Suprime-Cam (\\cite{miyazaki2002}) on Subaru and the CFH12k camera (\\cite{cuillandre2000}) on the Canada-France-Hawaii telescope (CFHT). The remaining 27 SN candidates were discovered using the Advanced Camera for Surveys (ACS; \\cite{benitez2003}) on HST\\footnote{Based on observations made with the NASA/ESA Hubble Space Telescope and obtained from the data archive at the Space Telescope Institute. STScI is operated by the Association of Universities for Research in Astronomy, Inc. under the NASA contract NAS 5-26555. The observations are associated with program 10496.} targeting high redshift galaxy clusters in a program called the HST Cluster SN Survey (Program number 10496, PI: Perlmutter). These SN searches specifically targeted $z>1$ SNe~Ia, for which there have been relatively few spectroscopically confirmed SNe~Ia (\\cite{aldering1998}; \\cite{coil2000}; \\cite{tonry2003}; \\cite{barris2004}; \\cite{riess2004}; \\cite{lidman2005}; \\cite{riess2007}). In \\S\\ref{sec:observationanddata}, we summarize the SN searches and present data for both imaging and spectroscopy. Spectroscopic data reductions are described in \\S\\ref{sec:observationanddata}. SN and host galaxy classifications are shown in \\S\\ref{sec:sntyping} and \\S\\ref{sec:focashostspectraforhstclustersn}, respectively. In \\S\\ref{sec:typeideff}, we describe factors that influence the classification of SNe. \\S\\ref{sec:summary} is a summary of the paper. We use the standard $\\Lambda$CDM cosmological parameters of $(H_0, \\Omega_M, \\Omega_\\Lambda)=(70, 0.27, 0.73)$ for calculating the age of the universe at a certain redshift. All magnitudes are measured in the AB system. ", "conclusions": "\\label{sec:summary} We presented spectra of SN candidates obtained with FOCAS on the Subaru 8.2-m telescope. Seven active candidates were identified as SNe Ia, including SCP06G4 at $z_{\\rm{SN}}=1.35$, the most distant SN~Ia to be spectroscopically identified with a ground-based telescope. Redshifts were obtained for all but 7 of the remaining 32 candidates based on the host galaxy spectra. An additional 4 candidates are identified as likely SNe~Ia from the spectrophotometric properties of their hosts. The spectral properties of these hosts found in the HST Cluster Supernova Survey were examined by comparing their spectra with BC03 model spectra. All of the host galaxy spectra indicate that they are passively evolving galaxies that have quenched their star forming activities. We also investigated the factors affecting the classification of SNe and found that it is critical to take spectra within a week of maximum light. This requires secure early detection, well-sampled light curves and prompt spectroscopic follow-up. \\bigskip We thank the anonymous referee for providing helpful comments and suggestions. TM and YI are financially supported by the Japan Society for the Promotion of Science (JSPS) through the JSPS Research Fellowship. CL acknowledges the financial support from the Oskar Klein Centre at the University of Stockholm. This work was supported in part with scientific research grants (15204012 and 17104002) from the Ministry of Education, Science, Culture, and Sports of Japan, and a JSPS core-to-core program ``International Research Network for Dark Energy''. Financial support for this work was provided in part by NASA through program GO-10496 from the Space Telescope Science Institute, which is operated by AURA, Inc., under NASA contract NAS 5-26555. This work was also supported in part by the Director, Office of Science, Office of High Energy and Nuclear Physics, of the U.S. Department of Energy under Contract No. AC02-05CH11231. Part of the Suprime-Cam observations were done during the guaranteed time observation of Suprime-Cam and we thank for the Suprime-Cam instrument team. We also appreciate much help by the SDF and SXDS project team members. We also thank for Youichi Ohyama, who helped our observations as a support scientist of FOCAS. Data analysis were in part carried out on common use data analysis computer system at the Astronomy Data Center, ADC, of the National Astronomical Observatory of Japan." }, "0911/0911.4820_arXiv.txt": { "abstract": "The coupled dark energy models, in which the quintessence scalar field nontrivially couples to the cold dark matter, have been proposed to explain the coincidence problem. In this paper we study the perturbations of coupled dark energy models and the effects of this interaction on the current observations. Here, we pay particular attention to its imprint on the late-time Integrated Sachs-Wolfe (ISW) effect. We perform a global analysis of the constraints on this interaction from the current observational data. Considering the typical exponential form as the interaction form, we obtain that the strength of interaction between dark sectors is constrained as $\\beta<0.085$ at $95\\%$ confidence level. Furthermore, we find that future measurements with smaller error bars could improve the constraint on the strength of coupling by a factor two, when compared to the present constraints. ", "introduction": "Current cosmological observations, such as the cosmic microwave background (CMB) measurements of temperature anisotropies and polarization \\cite{Komatsu:2008hk} and the redshift-distance measurements of Type Ia Supernovae (SNIa) at $z<2$ \\cite{Kowalski:2008ez}, have demonstrated that the Universe is now undergoing an accelerated phase of expansion and that its total energy budget is dominated by the dark energy component. The nature of dark energy is one of the biggest unsolved problems in modern physics and has been extensively investigated in recent years, both under the theoretical and the observational point of view. The simplest candidate of dark energy is the cosmological constant, where the equation of state $w$ always remains $-1$. In the standard $\\Lambda$-Cold Dark Matter ($\\Lambda$CDM) cosmology, cold dark matter only interacts with other components by gravity, while dark energy is simply a cosmological constant without any evolution and perturbation. Although this concordance model can fit the current observational data very well \\cite{Komatsu:2008hk,Kowalski:2008ez}, the possibility of dynamical dark energy models cannot be ruled out yet. Furthermore, this model suffers of the fine-tuning and coincidence problems, i.e. why the Universe is dominated by dark energy in late times \\cite{CCproblem1,CCproblem2}. In order to lift these severe tensions, many alternative dynamical dark energy models, such as quintessence \\cite{Quint1,Quint2,Quint3,Quint4}, phantom \\cite{Phantom}, K-essence \\cite{kessence1,kessence2} and quintom \\cite{Feng:2004ad}, have been proposed recently (for a review see Ref.\\cite{copeland}). For example, the quintessence model, a scalar field $\\phi$ slowly rolling down a potential energy $V(\\phi)$, has the spatial fluctuations and its equation of state $w_{\\phi}$ can evolve with the cosmic time. More importantly, being a dynamical component, the quintessence is naturally expected to interact with the other components, such as the cold dark matter \\cite{Copeland98,Amendola00} or massive neutrinos \\cite{gu03,fardon04}, in the field theory framework. If these interactions really exist, it would open up the possibility of detecting the dark energy non-gravitationally. In the coupled dark energy models, the quintessence scalar field could nontrivially couple to the cold dark matter component. The presence of the interaction clearly modifies the cosmological background evolutions. The evolution of cold dark matter energy density $\\rho^{}_{c}(a)$ will be different, when compared to that of the minimally coupled models, and dependent on the quintessence field $\\phi\\,$: \\begin{equation} \\rho^{}_{\\rm c}(a)=\\rho^{}_{\\rm c0}a^{-3+\\delta(\\phi)}~,\\label{cdm} \\end{equation} where $a$ is the scale factor, $\\rho^{}_{\\rm c0}$ is the present value of cold dark matter energy density, and $\\delta(\\phi)$ is the modification due to the interactions and could be non-zero during the evolution of Universe. In this case, at early times there will be more (less) cold dark matter energy density, when $\\delta(\\phi)<0$ ($>0$). On the other hand, the interaction between dark sectors will also affect the evolution of cosmological perturbations, which has been widely investigated recently (see e.g. Refs.\\cite{hwangnoh02,amendola04,koivisto05,lee06, olivares06,brookfield08,bean08,mainini08,valiviita08,he09a,chong09}). Due to the different evolution of cold dark matter energy density, the interaction between dark sectors could shift the matter-radiation equality scale factor $a_{\\rm eq}$, and affect the locations and amplitudes of acoustic peaks of CMB temperature anisotropies and the turnover scales of large scale structure (LSS) matter power spectrum consequently. In addition, the interaction will also affect the late ISW effect \\cite{isw} at large scales which is produced by the CMB photons passing through the time-evolving gravitational potential well, when dark energy or curvature becomes important at later times \\cite{brookfield08,he09b}. Therefore, with the accumulation of observational data and the improvements of the data quality, it is of great interest to investigate the non-minimally coupled dark energy models from the current observational data. In this paper we study the perturbations of coupled dark energy models and present the constraints on the interactions from the observational data in detail. The structure of the paper is as follows: in Sec.II and Sec.III we show the basic equations of background evolution and linear perturbations of the coupled dark energy model, respectively. In Sec.IV we present the current and future observational datasets we used. Sec.V contains our main global fitting results from the current and future observations, while Sec.VI is dedicated to the summary. ", "conclusions": "The coupled dark energy models, in which the quintessence scalar field non-minimally couples to the cold dark matter, could affect the CMB temperature anisotropies and LSS matter power spectrum. In this paper we present constraints on this coupled dark energy model using the latest observations and future measurements. The WMAP5 data alone could only give a weak constraint on the strength of coupling $\\beta$. And then we exploit the capabilities of the late-time ISW effect in constraining the non-minimal coupling, using the cross-correlation signal between the quasar sample of SDSS DR6 \\cite{richards09} and the WMAP5 ILC map \\cite{hinshaw09}. We find that the current ISW data could slightly improve the constraint on $\\beta$. If we add the BAO and SNIa data into our calculations, the constraint on $\\beta$ will improves significantly, namely the $95\\%$ upper limit is $\\beta<0.085$. Finally we simulate the future measurements with smaller error bars and find that the future measurements could improve the constraint on the strength of coupling by a factor of two, when compared to the present constraints." }, "0911/0911.2637_arXiv.txt": { "abstract": "{} {We study three Galactic \\HII\\ regions -- RCW~79, RCW~82 and RCW~120 -- where triggered star formation is taking place. Two stellar population are observed: the ionizing stars of each \\HII\\ region and young stellar objects on their borders. Our goal is to show that they represent two distinct populations, as expected from successive star forming events. } {We use near--infrared integral field spectroscopy obtained with SINFONI on the VLT to make a spectral classification. We derive the stellar and wind properties of the ionizing stars using atmosphere models computed with the code CMFGEN. The young stellar objects are classified according to their $K$--band spectra. In combination with published near and mid infrared photometry, we constrain their nature. Linemaps are constructed to study the geometry of their close environment. } {We identify the ionizing stars of each region. RCW~79 is dominated by a cluster of a dozen O stars, identified for the first time by our observations. RCW~82 and RCW~120 are ionized by two and one O star, respectively. All ionizing stars are early to late O stars, close to the main sequence. The cluster ionizing RCW~79 formed 2.3$\\pm$0.5 Myr ago. Similar ages are estimated, albeit with a larger uncertainty, for the ionizing stars of the other two regions. The total mass loss rate and ionizing flux is derived for each regions. In RCW~79, where the richest cluster of ionizing stars is found, the mechanical wind luminosity represents only 0.1\\% of the ionizing luminosity, questioning the influence of stellar winds on the dynamics of these three \\HII\\ regions. The young stellar objects show four main types of spectral features: H$_{2}$ emission, \\brg\\ emission, CO bandheads emission and CO bandheads absorption. These features are typical of young stellar objects surrounded by disks and/or envelopes, confirming that star formation is taking place on the borders of the three \\HII\\ regions. The radial velocities of most YSOs are consistent with that of the ionized gas, firmly establishing that their association with the \\HII\\ regions. Exceptions are found in RCW~120 where differences up to 50 \\kms\\ are observed. Outflows are detected in a few YSOs. All YSOs have moderate to strong near--IR excess. In the [24] versus $K-$[24] diagram, the majority of the sources dominated by H$_{2}$ emission lines stand out as redder and brighter than the rest of the YSOs. The quantitative analysis of their spectra indicates that for most of them the H$_{2}$ emission is essentially thermal and likely produced by shocks. We tentatively propose that they represent an earlier phase of evolution compared to sources dominated by \\brg\\ and CO bandheads. We suggest that they still possess a dense envelope in which jets or winds create shocks. The other YSOs have partly lost their envelopes and show signatures of accretion disks. Overall, the YSOs show distinct spectroscopic signatures compared to the ionizing sources, confirming the presence of two stellar populations. } {} ", "introduction": "\\label{s_intro} Massive stars play a significant role in several fields of astrophysics. They produce the majority of heavy elements and spread them in the interstellar medium, taking an active part in the chemical evolution of galaxies. But they also end their life as supernovae and gamma--ray bursts. Through their strong winds and ionizing fluxes they power \\HII\\ regions and bubbles which are often used to trace metallicity gradients in galaxies. The energy they release in the interstellar medium is thought to trigger second--generation star formation events. Observations of young stellar objects (YSO) in molecular clouds surrounding (clusters of) massive stars lend support to this mechanism \\citep[e.g.][]{walborn02,hatano06}. A particular case concerns star formation on the borders of \\HII\\ regions. According to the collect and collapse model \\citep{el77}, a dense shell of material is trapped between the shock and ionization front of an expanding \\HII\\ regions. When the amount of collected material is large enough, global shell fragmentation occurs and new stars are formed. The observation of molecular condensations on the borders of several \\HII\\ regions and the subsequent identification of YSOs within these clumps \\citep{de03,de05,za06,za07,de08a,de09,po09} confirms that this mechanism is at work at least in some \\HII\\ regions. Other mechanisms leading to triggered star formation exist. Some work qualitatively as the collect and collapse model in the sense that the clumps are formed during the \\HII\\ region expansion. For instance, dynamical instabilities of the ionization front \\citep{vishniac83,garcia96} create molecular condensations separated by zones of lower densities. The newly formed clumps grow until they become Jeans unstable and collapse. Alternatively, second generation star formation can happen in pre-existing clumps. If the neutral gas in which the \\HII\\ region expands is not homogeneous, the outer layers of the molecular overdensities are ionized like the borders of a classical \\HII\\ region. A shock front precedes the ionization front inside these clumps, leading to their collapse \\citep{duvert90,ll94}. A number of questions regarding triggered star formation remain unanswered. The properties of the observed YSOs are poorly known besides a crude classification in class I or class II objects by analogy with low--mass stars. YSOs usually display near--infrared spectra with CO, \\brg\\ and/or H$_{2}$ emission lines \\citep{bik06}. The relation, if any, between objects with different spectroscopic appearance is not clear. Besides, in the regions where the collect and collapse process is at work, the quantitative properties of the ionizing sources of the \\HII\\ regions are not known. In particular, the relative role of ionizing radiation and stellar winds on the dynamics of such regions is debated. The timescales under which material accumulates and fragments, and the properties of the resulting clumps depend on the strength of those two factors \\citep{whit94}. Hence, one might wonder whether the nature of the newly formed objects depends on the properties of the stars powering the \\HII\\ regions. In the present study, we tackle these questions by investigating the properties of the ionizing stars and YSOs of three Galactic \\HII\\ regions: RCW~79, RCW~82 and RCW~120 \\citep{rod60}. Those regions are known to be the sites of triggered star formation \\citep{za06,za07,po09,kang09}. We have used SINFONI on the VLT to obtain near--infrared spectra of both the ionizing stars and a selection of YSOs in each region. Our main goals were: \\begin{itemize} \\item Identify the ionizing stars of each region, and derive their stellar properties using atmosphere models. In particular, we want to determine their ionizing fluxes and mass loss rates in order to better understand the dynamics of the \\HII\\ regions. Equally important is the determination of the age of those stars, since it can be related to the presence of YSOs to quantitatively confirm the existence of triggered star formation. \\item Constrain the nature of the YSOs on the borders of the \\HII\\ regions. In combination with infrared photometry, spectroscopy can reveal the presence of disks or envelopes. The evolutionary status of those objects can thus be better understood. In particular, it can be clearly shown whether they are stars still in their formation process or objects already on the main sequence. \\end{itemize} The paper is organized as follows. In Sect.\\ \\ref{pres_hii} we present the three \\HII\\ regions targeted in this study. Sect.\\ \\ref{s_obs} describes our observations. The analysis of the ionizing stars is presented in Sect.\\ \\ref{s_ex}, while the YSOs are discussed in Sect.\\ \\ref{s_yso}. We discuss our results in Sect.\\ \\ref{s_disc} and summarize our conclusions in Sect.\\ \\ref{s_conc}. ", "conclusions": "\\label{s_conc} We have obtained near--infrared integral field spectroscopy of three \\HII\\ regions (RCW~79, RCW~82, RCW~120) with SINFONI on the VLT. The main results of our study are: \\begin{itemize} \\item[$\\bullet$] We have identified the ionizing sources of all three \\HII\\ regions. RCW~79 is powered by a cluster of O stars unraveled here for the first time. The most massive stars have a spectral type O4--6V/III. A number of later type OB stars are also detected. RCW~82 is ionized by two O9--B2V/III stars and a third star, probably of spectral type Be. Finally, a single O6--8V/III star ionizes the \\HII\\ region in RCW~120. \\item[$\\bullet$] The ionizing stars of each region have been analyzed with atmosphere models computed with the code CMFGEN. The derived stellar properties have been used to build the HR diagram. In the case of RCW~79, the large number of stars allowed a reliable age determination: the ionizing cluster formed 2.3$\\pm$0.5 Myr ago. For the other two \\HII\\ regions, similar ages are suggested but are less well constrained due to the small number of sources. An upper limit of $10^{-7}$ \\myr\\ was derived for the mass loss rate of the ionizing stars of all three regions. \\item[$\\bullet$] The cumulative mass loss rate due to the O stars at the center of RCW~79 is $< 10^{-6}$ \\myr. The resulting wind mechanical luminosity is about $10^{-3}$ the ionizing luminosity. This is ten times weaker than in the numerical simulations of \\citet{freyer03}, raising the question of the quantitative role of winds on the dynamics of the three \\HII\\ regions studied here. The presence of hot dust emission close to the ionizing stars argues against a strong influence of stellar winds. \\item[$\\bullet$] Spectroscopy of the infrared sources on the borders of the \\HII\\ regions revealed typical signatures of young stellar objects. Four main categories of YSOs have been identified: (1) sources dominated by H$_{2}$ emission lines, (2) sources showing mainly \\brg\\ emission, (3) sources with CO bandheads in emission and (4) sources with CO bandheads in absorption. \\item[$\\bullet$] Near and mid--infrared color--color and color--magnitude diagrams indicate that all YSOs have SEDs dominated by non stellar emission. In the Spitzer/IRAC [3.6]$-$[4.5] vs [5.8]$-$[8.0] diagram, essentially all YSOs are grouped at the position between class I and class II objects. Only a couple of H$_{2}$ sources seem to be redder. The trend is more clearly seen in the [24] vs $K$--[24] diagram where the H$_{2}$ sources form clearly a distinct group with larger 24~\\mum\\ emission and redder $K$--[24] colors. \\item[$\\bullet$] The detailed analysis of the H$_{2}$ emitting sources through line ratios and excitation diagrams indicate that thermal emission is the dominant mechanism in the majority of sources. Slow J--shocks or fast C--shocks are the preferred explanation for this emission. Some sources show a possible weak contribution of fluorescence emission, and three are dominated by this mechanism. \\item[$\\bullet$] Measurement of radial velocities and line width was performed in a few sources. In general, the derived radial velocities are consistent with that of the ionized gas in the \\HII\\ region, confirming that the YSOs are well associated with the regions. The exception is RCW~120, for which most sources move 20 to 50 \\kms\\ faster than the ionized gas. In sources showing both \\brg\\ and H$_{2}$ lines, the former are systematically wider (150/250 \\kms vs non--resolved/100 \\kms), possibly due to a different spatial origin. \\item[$\\bullet$] Morphological studies reveal that when both \\brg\\ and \\htwo\\ emission are present, \\brg\\ is centered on the position of the YSO while \\htwo\\ emission is more extended. In some cases, the \\htwo\\ emission peak is offset compared to that of the YSO. RCW120 51 presents a double \\htwo/\\brg\\ cone structure following the ionization front. RCW82 7 (and possibly RCW120 49 and RCW120 C4--67) has a \\htwo\\ velocity structure typical of an outflow. \\item[$\\bullet$] The spectrophotometric and kinematic properties of YSOs paint the following plausible picture: H$_{2}$ dominated sources are stellar objects surrounded by an envelope in which shocks (due to jets or stellar/disk winds) produce H$_{2}$ emission; other YSOs correspond to more evolved sources in which the envelope partially or totally disappeared, revealing a disk structure from which CO and/or \\brg\\ emission is produced. \\end{itemize} Future tailored observations of a few key YSOs will provide more information on the dynamics of the circumstellar material. Additional UV and optical spectroscopy of the ionizing stars of RCW~79 will help to refine the age estimate and to reveal a possible age difference with the stars in the CHII region on the border of RCW~79. This would be a first direct proof of the existence of triggered massive star formation. Better estimates of the mass loss rates will also be possible, helping to quantify the role of stellar winds in the dynamics of \\HII\\ regions." }, "0911/0911.5221_arXiv.txt": { "abstract": "A model is presented for generation of fast solar wind in coronal holes, relying on heating that is dominated by turbulent dissipation of MHD fluctuations transported upwards in the solar atmosphere. Scale-separated transport equations include large-scale fields, transverse Alfv\\'enic fluctuations, and a small compressive dissipation due to parallel shears near the transition region. The model accounts for proton temperature, density, wind speed, and fluctuation amplitude as observed in remote sensing and \\emph{in situ} satellite data. ", "introduction": "An open question in solar and heliospheric physics is to identify the physical processes responsible for heating the corona and accelerating the fast solar wind streams emanating from coronal holes. This requires that a fraction of the energy available in photospheric motions be transported through the chromospheric transition region, and dissipated in the corona. The measured speeds of fast solar wind streams require spatially extended heating \\citep{WithbroeNoyes77,HolzerLeer80,Withbroe88}. The physical mechanisms for this transport and dissipation have remained elusive. Some models have resorted to use of a parametrically defined heat deposition (a ``heat function'') that decays exponentially with height, or anomalous heat conduction that redistributes energy along field-lines \\citep{HabbalEA95,McKenzieEA95,BanaszkiewiczEA98}. One-dimensional (1D) models of this kind, extending from the chromosphere to 1\\,AU \\citep{HansteenEA94,HansteenEA97}, have helped in understanding the regulation of the solar wind mass flux and can reproduce fast solar wind streams originating in cool electron coronal holes. Here we present a model that demonstrates solar wind acceleration due to heating by a quasi-incompressible turbulent cascade triggered by coronal stratification \\citep{MatthaeusEA99}, and supplemented by compressive heating near the base. This model accounts for most presently available coronal and interplanetary observations. The idea that broadband plasma fluctuations might heat the extended corona and accelerate the solar wind has long been discussed \\citep{Coleman68,BelcherDavis71,Hollweg86,HollwegJohnson88,Velli93a}. However, the mechanisms of transfer of fluctuation energy to small scales---and, in particular, the role of Alfv\\'enic turbulence (as observed in the solar wind) and cascade processes (e.g., phase mixing, ponderomotive driving, shocks, etc.)---have not been described self-consistently. Two recent papers shed light on these relationships \\citep{SuzukiInutsuka05,CranmerEA07}. \\citet{SuzukiInutsuka05} incorporate 1D compressive nonlinear interactions driven by Alfv\\'en waves and leading to shock heating. This model produces good agreement with solar wind speed profiles. As low-frequency Alfv\\'en waves propagate upward, their wave pressure compresses the plasma. Unable to refract or mode-mode couple into a perpendicular wavenumber cascade, these waves must dissipate in 1D shock fronts. This model provides a valuable demonstration that MHD fluctuations can act as a conduit to transport energy to the requisite altitudes. However the restriction to 1D cascade is at odds with the well-established propensity for an incompressible MHD cascade to proceed mainly through wavevectors perpendicular to a strong mean magnetic field \\citep{RobinsonRusbridge71,ShebalinEA83,% OughtonEA94,BieberEA96,ChoEA02-strongB0}. Furthermore the corona exhibits a clear transverse structuring, and the initial fluctuations must have perpendicular correlation lengths not much larger than a super-granulation scale ($15\\,000$\\,km). Another recent model \\citep{CranmervanBall05,CranmerEA07} incorporates a low-frequency cascade model \\citep{VerdiniEA06-soho}; however, the treatment of propagation and dissipation differs significantly from the present approach. Their scheme treats nonlinear effects as a perturbation, and it is unclear if it converges for strong turbulent heating. Here we employ a strong turbulence closure. We also do not rely on electron heat conduction to boost radial energy transport. Instead we compute an internal energy associated with the protons only. This approach supports comparisons with results employing improved representation of turbulence, such as shell models \\citep{VerdiniEA09-emp, VerdiniEA09-apj} and (potentially) full MHD simulation. We find here that reflection of Alfv\\'enic turbulence alone does not lead to a full corona/solar wind stationary state -- a compressible contribution is required. This is supported by recent observations from \\emph{Hinode} \\citep{LangangenEA08,DePontieuEA09}. % When we include a small component of compressive heating near the coronal base, fast solar wind streams are then accounted for. ", "conclusions": "The above model shows that turbulence near the coronal base, originating through chromospheric transmission of fluctuations, can heat the plasma in an expanding coronal hole flux-tube and produce a fast solar wind that matches a number of observational constraints. The turbulence is mainly of the low-frequency Alfv\\'enic type. A small amount of compressive heating between the transition region and the sonic point appears to be needed to match the observations. This additional heating may be due to type II spicules that supply broadband low-frequency vertical fluctuations at transition region heights, thus launching compressive MHD modes near the coronal base \\citep{DePontieuEA09}. Most of the fluctuation energy is in low-frequency turbulence, and this sustains a strong anisotropic MHD cascade through reflections from local density gradients. This type of anisotropic cascade is favored in MHD turbulence in the presence of a strong DC magnetic field \\citep{RobinsonRusbridge71,ShebalinEA83,OughtonEA94}. Heat conduction does not enter the present model at all, since it mainly affects electron internal energy, which evolves independently in this approximation. Similar assumptions work well in understanding observations of solar wind turbulence \\citep{BreechEA09,CranmerEA09}. In these ways, the present model differs from other recent models that incorporate turbulent heating \\citep{SuzukiInutsuka05,CranmerEA07}. In particular we believe that this model demonstrates, possibly for the first time, that a model (almost) free of ad hoc heat functions, artificial equations of state, and ad hoc assumptions about heat conduction, can indeed heat the corona and accelerate the solar wind. We plan further study of the required small amount of compressible heating, attributed here to spicule-driven magnetosonic activity. Another\tuseful extension would be to include separate electron and proton internal energy budgets, which will enable additional observational constraints, and will permit study of the role of kinetic dissipation processes \\citep{BreechEA09,CranmerEA09}." }, "0911/0911.2771_arXiv.txt": { "abstract": "We analysed the orientation of galaxy groups in the Local Supercluster (LSC). It is strongly correlated with the distribution of neighbouring groups in the scale till about 20 Mpc. The group major axis is in alignment with both the line joining the two brightest galaxies and the direction toward the centre of the LSC, i.e. Virgo cluster. These correlations suggest that two brightest galaxies were formed in filaments of matter directed towards the protosupercluster centre. Afterwards, the hierarchical clustering leads to aggregation of galaxies around these two galaxies. The groups are formed on the same or similarly oriented filaments. This picture is in agreement with the predictions of numerical simulations. ", "introduction": "\\citet{b2} was the first who found that major axes of galaxy clusters tend to point towards their neighbours. Later the existence of this effect was discussed by several authors and usually the significant alignment was reported. The distance between clusters for which the effect was detected changed from $10 h^{-1} Mpc$ till $150 h^{-1} Mpc$ (where $h=H_0/100\\,km\\,s^{-1} Mpc^{-1}$). The strength of the effect decreases with distance \\citep{s2,f3,r1,u1,p2,w4,c1,h3}. These investigations involved both optical and X-ray data, as well as clusters belonging (or not) to superclusters. Nowadays, it is accepted that the effect is not due to selection effects but is real and its distance scale is between $10 - 60 h^{-1} Mpc$. The alignment of galaxy group was studied by \\citet{w3}. He used CfA group catalog \\citep{g1} and a catalog based on SSRS ( Southern Sky Redshit Survey) \\citep{m1}. Each group should have at least $4$ objects and less than $100$. Decontamination of the groups by foreground and background objects was performed by removing objects with redshift difference from the group mean over $1000\\, km\\,s^{-1}$. They were $59$ groups and Binggeli effect was observed among galaxy groups till $15- 30 h^{-1} Mpc$. One should note that the investigation of galaxy group orientation is more difficult than in the case of galaxy clusters. The position angle of a group consisting of a few objects is determined with much greater error than for rich clusters. Moreover, statistics with small number are less reliable. The other investigations of galaxy groups were performed by \\citet{p1}. During study the orientation of $92$ out of $100$ compact groups listed in the Catalogue of the Compact Groups \\citep{h5}, they do not find alignment, but location of groups along long chains is noted. The interpretation of the effect has changed with development of theories, but the main idea that this should reflect conditions during the structure formation is still very popular. Numerical simulations gave a better understanding of physical processes leading to structure formation. These simulations were performed in the framework of the cold dark matter (CDM) model presently regarded as the correct description of the large scale structure formation. Using different approaches and codes, these investigations led to the conclusion that the preferred orientation of galaxy clusters in the CDM model is a natural consequence of processes leading to structure formation due to gravitational interaction. \\citet{o1} using large - scale simulation found that in $\\Lambda$CDM cosmological model effect reaches the distance up to $30 Mpc$, while in $\\tau$CDM model the range of effects is twice smaller. Also in SCDM and OCDM models, where smaller scale simulations were performed, some alignment effect could be noted. N - body simulation for standard $\\Lambda$CDM model \\citep{f1} for $3000$ clusters showed the alignment of neighbouring clusters in the distance range from $10 - 15 Mpc$, while the Binggeli effect till about $100 Mpc$ is observed. The strong alignment, decreasing with the increasing distance between clusters, observed till about 100 Mpc was reported \\citep{h6}. The preferred orientation of clusters belonging to superclusters existed also in SHM + N body simulation . Moreover, from some of the numerical simulations, it follows that the structure formation occurred along filamentary structures rather than the walls \\citep{f2,s1,b2,h1,h2,a1,w1,w2}. In order to confirm, or deny, this scheme of structure origin we carry out an analysis of the Local Supercluster (LSC) galaxy groups alignment as well as the distribution of the acute angle between the position angle of the structure and the direction to all remaining clusters. Groups were taken from Nearby Galaxies (NBG) Catalog \\citep{t3}. We investigated also the alignment of the brightest group galaxy and the parent group. We determined the position of the line joining the two brightest galaxies and checked the orientation of this line in respect to parent group and direction towards Virgo cluster. The distributions of these angles, as well as differences between some of these angles were tested for isotropy. The paper is organised in the following manner. Section 2 describes observational data, section 3 presents statistical method used in the paper, section 4 presents results and discussions. In section 5 we formulate conclusions. ", "conclusions": "We used two samples of data. The first one contains $61$ groups having at least $10$ members, in agreement with West's (1989) wish to have such sample of data. From the second set of data poorer groups were eliminated and $35$ groups having at least $20$ members remained for analysis. All groups are nearby and they are located within the Local Supercluster. Our main results are: \\begin{enumerate} \\item The group major axis is in alignment with the line joining the two brightest galaxies. \\item The group major axis is in alignment with the direction toward the centre of the LSC. \\item The acute angle between the position angle of the group and direction towards each remaining group is not isotropic. \\item The structures have tendency to point each other when the distance between groups is smaller than $20 Mpc$ \\end{enumerate} We performed study of the acute angle between the position angle of the group and direction towards each remaining group denoted as $\\phi$. The sample was divided according to the distance between group centres. Each subsample was analysed independently using two coordinate systems. In the equatorial coordinate system the distributions of the $\\phi$ - angle for all subsamples disregarding subsample $D<10 Mpc$ were isotropic. This is not the case of the coordinate system connected with the Local Supercluster. We found that for closer neighbours ($D<10 Mpc$) strong alignment is observed. It is at almost $5 \\sigma$ level. The Binggeli effect is diminishing with distance increase, vanishing at about $20 Mpc$. We used the K - S test, which is usually applied for alignment investigation. \\citet{c2} criticised it, because it does not point the place, where the departure of isotropy is observed. They preferred to use the Wilcox test on rank-sum, which, in their opinion, gave a higher confidence signal for alignment. However, \\citet{o1} applied both tests finding a little difference in the obtained statistics. Furthermore, we used the K - S test only to check the isotropy of the distribution and not to find the anisotropy location. Our results obtained with the help of K-S test are confirmed by linear regression analysis. The fact that detection of anisotropy is connected with LSC coordinate system supports the point of view that formation of galaxies occurred within protostructures. The analysis of the differences between position angles shows that it is possible that there exists the alignment of the line joining two brightest galaxies with both position angle of the parent group and direction towards Virgo cluster centre. The fact that detection of anisotropy is connected with LSC coordinate system supports the point of view that formation of galaxies occurred within protostructures. The analysis of the differences between position angles shows that it is possible that there exists the alignment of the line joining two brightest galaxies with both position angle of the parent group and direction towards Virgo cluster centre. From the presented analysis of the orientation of galaxy groups in the Local Supercluster the following picture of the structure formation appears. The two brightest galaxies were formed first. They originated in the filamentary structure directed towards the centre of the protocluster. This is the place where the Virgo cluster centre is located now. Due to gravitational clustering, the groups are formed in such a manner that galaxies follow the line determined by the two brightest objects. Therefore, the alignment of structure position angle and line joining two brightest galaxies is observed. The other groups are forming on the same or nearby filament. The flatness of the LSC additionally contributes to the observed alignment of galaxy groups. The majority of the groups lie close to us. Due to completeness of the Catalog, the lack of groups further than the Virgo Cluster centre is observed, but nearby groups are very well selected and they contain only more massive galaxies. This picture is in agreement with predictions of several CDM models, in which structure formation is due to hierarchical clustering. Moreover, the formation is occurring on the filamentary structure. The further investigation considering groups clearly inside and outside superclusters on the greater data set will be very useful to support or reject this picture." }, "0911/0911.5198_arXiv.txt": { "abstract": "Uninhibited radiative cooling in clusters of galaxies would lead to excessive mass accretion rates contrary to observations. One of the key proposals to offset radiative energy losses is thermal conduction from outer, hotter layers of cool core clusters to their centers. However, thermal conduction is sensitive to magnetic field topology. In cool-core clusters where temperature decreases inwards, the heat buoyancy instability (HBI) leads to magnetic fields ordered preferentially in the direction perpendicular to that of gravity, which significantly reduces the level of conduction below the classical Spitzer-Braginskii value. However, the cluster cool cores are rarely in perfect hydrostatic equilibrium. Sloshing motions due to minor mergers and stirring motions induced by cluster galaxies or active galactic nuclei (AGN) can significantly perturb the gas. The turbulent cascade can then affect the topology of the magnetic field and the effective level of thermal conduction. We perform three-dimensional adaptive mesh refinement magnetohydrodynamical (MHD) simulations of the effect of turbulence on the properties of the anisotropic thermal conduction in cool core clusters. We show that very weak subsonic motions, well within observational constraints, can randomize the magnetic field and significantly boost effective thermal conduction beyond the saturated values expected in the pure unperturbed HBI case. We find that the turbulent motions can essentially restore the conductive heat flow to the cool core to level comparable to the theoretical maximum of $\\sim 1/3$ Spitzer for a highly tangled field. Runs with radiative cooling show that the cooling catastrophe can be averted and the cluster core stabilized; however, this conclusion may depend on the central gas density. Above a critical Froude number, these same turbulent motions also eliminate the tangential bias in the velocity and magnetic field that is otherwise induced by the trapped $g$-modes. Our results can be tested with future radio polarization measurements, and have implications for efficient metal dispersal in clusters. \\\\ ", "introduction": "\\label{section:intro} X-ray clusters show a strong central surface brightness peak, and many of them have short central cooling times. However, {\\it Chandra} and {\\it XMM-Newton} observations have shown that actual gas cooling rates fall significantly below classical values from a standard cooling flow model; emission lines such as Fe XVII expected from low temperature gas are not observed \\citep{peterson06}. This suggests that besides quasi-hydrostatic equilibrium, clusters might also be in quasi-thermal equilibrium, with some form of heating balancing cooling. Many forms of heating have been postulated. Hybrid models can account for the observations; leading contenders generally involve some combination of AGN heating and thermal conduction. The remarkable discovery by {\\it Chandra} of AGN blown bubbles in clusters has triggered a widespread renaissance in the study of AGN feedback (see \\citet{mcnamara07} for a recent review). It has been suggested that the AGN feedback {\\it alone}, generally fails to distribute heat in the spatially distributed fashion required to offset cooling (e.g., Vernaleo \\& Reynolds 2006), both in its radial dependence and solid angle coverage. However, recent simulations by Br{\\\"u}ggen \\& Scannapieco (2009) involving subgrid model for Rayleigh-Taylor-driven turbulence show that AGN heating can self-regulate cool cores. In these non-MHD simulations the subgrid turbulence model is crucial to simulate the evolution of the cool core. In a realistic situation, the ability to self-regulate undoubtely depends on a range of the mechanisms used to transmit the energy from the AGN blown bubbles to the ICM (e.g., dissipative sound waves, $pdV$ work, cosmic rays, suppression of both the Rayleigh-Taylor instability and the mixing of the bubble material by (even weak) magnetic fields). These issues are a matter of intense debate. At the same time, it is a remarkable fact that if one simply uses observed temperature profiles in clusters to construct the Spitzer conductive flux from the cluster outskirts, it is very nearly equal to that required to balance the radiative cooling rate as indicated by the observed X-ray surface brightness profile, for some reasonable fraction $f\\sim 0.3$ of the Spitzer value (e.g., Fig. 17 of Peterson \\& Fabian 2006). There is no reason in principle why such close agreement should exist, and it is a tantalizing hint that nature somehow ``knows'' about Spitzer conductivity. Models which invoke both mechanisms are able to reproduce observed temperature and density profiles without any fine-tuning of feedback or conduction parameters \\citep{ruszkowski02,guo08a}, unlike conduction-only models \\citep{bertschinger86,zakamska03}. Nonetheless, even if no fine-tuning of conductivity is required, it is unsatisfying to invoke an ad-hoc suppression factor $f$; the anisotropic transport of heat along field lines can be calculated from first principles. If the field lines are tangled and chaotic, then $f\\sim 0.2$ might be appropriate \\citep{narayan01}. However, recent work has uncovered instabilities arising in stratified plasmas, which rearrange the field lines on large scales, with strong implications for heat transport. Unlike convective instability, which only arises if entropy declines with radius, these instabilities depend on temperature gradients, and arise irrespective of the sign of the gradient. If $dT/dr < 0$, as in the outer regions of clusters, the resulting magnetothermal instability (MTI; \\citet{balbus00,parrish05,parrish07,parrish08}) drives field lines to become preferentially radial, substantially boosting conductivity to close to the Spitzer value, $f \\sim 1$. On the other hand, if $dT/dr > 0$, as in the cores of cool-core clusters, then the resulting heat buoyancy instability (HBI; \\citet{quataert08, parrish09}, Bogdanovi{\\'c} et al. 2009) will drive field lines to become preferentially tangential, essentially shutting off radial conduction ($ f \\ll 1$), and leading to catastrophic cooling in the cluster core. Thus, it would seem that a more careful calculation vitiates the ad-hoc assumptions of models which invoke conduction. However, this presumes that there are no competing mechanisms which also affect field line topology. One possibility which has been speculated upon in the literature \\citep{ruszkowski07,guo08b,bogdanovic09} is AGN bubbles, which upon rising would leave radially orientated field lines in their wake. \\citet{guo09} showed that time-variable conduction, perhaps triggered by AGN outbursts, could allow clusters to cycle between the cool core and non-cool core states. Another, perhaps more general mechanism is simply turbulent gas motions, which can be induced by mergers \\citep{ascasibar06}, the motion of galaxies \\citep{kim07,conroy08}, AGN outbursts \\citep{mcnamara07}, or cosmic-ray driven convection \\citep{chandran07,sharma09}\\footnote{Note, however, that since the HBI is a buoyancy instability rather than a convective instability, turbulence cannot be self-consistently generated by the HBI itself, since the induced velocities are small \\citep{bogdanovic09,parrish09}.}. Indeed, \\citet{sharma09a} suggested that the amount of turbulence needed to avoid the stable end-state induced by the HBI could be estimated from the Richardson number \\citep{turner73}, which is simply the ratio of the restoring buoyant force to the inertial $(\\rho {\\bf u}\\cdot \\nabla {\\bf u})$ term. If one defines $Ri \\equiv g r ( d \\, {\\rm ln}T/d \\, {\\rm ln} r)/\\sigma^{2}$ for typical cluster conditions, the critical turbulent velocity is \\citep{sharma09a}: \\begin{eqnarray} \\sigma \\approx 135 \\, {\\rm km \\, s^{-1}} {g_{-8}}^{1/2} r_{10}^{1/2}\\left( \\frac{d \\, {\\rm ln}T/d \\, {\\rm ln} r}{0.15} \\right)^{1/2} \\left( \\frac{Ri_{c}}{0.25} \\right)^{-1/2}\\\\ \\nonumber \\end{eqnarray} \\noindent where $g_{-8}$ is the gravitational acceleration in units of $10^{-8} \\, {\\rm cm^{2} s^{-1}}$, $r_{10}$ is a characteristic scale height in units of $10$ kpc, and $Ri_{c}$ is the critical Richardson number. $Ri_{c} \\sim 1/4$ is typical for hydrodynamic flow; order unity differences might exist for the MHD case. Such levels of turbulence are easily seen in cosmological simulations, even in very relaxed clusters \\citep{evrard90,norman99,nagai03,vazza09a, vazza09b}, presumably due to continuing gas accretion and the supersonic motions of galaxies through the ICM. Gas ``sloshing'' in clusters has been invoked to explain observed cold fronts \\citep{ascasibar06}, and biases in X-ray mass profiles, particularly in the cluster core \\citep{lau09}. In the future, direct measurement of turbulent velocities could best be performed by a high-spectral resolution imaging X-ray spectrometer \\citep{rebusco08}, although early upper limits exist from {\\it XMM-Newton} (\\citet{sanders09}; for A1835, non-thermal broadening is less than $\\sim 275 \\, {\\rm km \\, s^{-1}}$). Besides the strong impact on conductive heat transport, the interaction of turbulence with conduction has testable implications for metal dispersal (which is strongly enhanced; \\citep{sharma09}). With an instrument as as the Square Kilometer Array, the field topology itself can potentially be mapped out via the rotation measure of background radio sources. Our goal in this paper is to quantitatively examine for the first time if indeed turbulent motions can overwhelm the tangential field topology driven by the HBI (we defer the MTI dominated regime to future work), and allow thermal conduction at the level required to avert a cooling catastrophe. We begin with a MHD {\\it FLASH} simulation of a cluster which in its unperturbed state suffers the HBI, consistent with previous work. We then simulate turbulence via a spectral forcing scheme that enables statistically stationary velocity fields (Eswaran \\& Pope 1998). An important feature of our study is that even if the medium is stirred isotropically, the turbulent velocity field itself can be anisotropic, and interact in non-trivial ways with thermal conduction. This is because when the turbulent driving frequency falls below the \\brunt frequency, {\\it g}-modes can be excited; the trapped modes induce tangentially biased vortices, due to the restoring forces which act on vertical motion \\citep{riley00}. Thus, even in the absence of the HBI, the excitation of $g$-modes can tangentially bias the magnetic field. In realistic clusters, $g$-modes also greatly enhance the efficacy of stirring, since trapped modes allow the turbulence to be volume-filling, although our present simulations already have volume-filling turbulence by design. The \\brunt frequency, which determines where $g$-modes can be produced, depends both on the gravitational potential and the gas temperature and density profile. The latter can in turn be affected by heating and thermal conduction. Thus, the cluster potential, gas properties, and thermal conduction interact in often subtle ways. Overall, we find that turbulence can indeed restore conductive heat flow back to the values $f\\sim 0.3$ expected for a highly tangled field, and required to stem a cooling flow. The organization of this paper is as follows. In \\S\\ref{section:turbulence}, we describe simple scaling relations to guide our intuition; in \\S\\ref{section:methods}, we describe our computational methods; in \\S\\ref{section:results}, our results; we then conclude in \\S\\ref{section:conclusions}. ", "conclusions": "\\label{section:conclusions} We have shown that a very low level of turbulent perturbations, $\\sim 50-150 \\, {\\rm km/s}$ to the ICM --- entirely consistent with the expectations for cosmological infall, galaxy motions, mergers, or AGN activity --- can entirely alter the magnetic field distribution resulting from the HBI instability. Instead of preferentially tangential magnetic fields, the final field configuration can be easily randomized. As expected from simple theoretical considerations, this happens when the typical Froude number (equation \\ref{eq:Fr_MHD}) $Fr^{\\rm MHD} \\gsim \\mathcal{O}(1)$. Note that weak, low frequency perturbing motions lead to trapped $g$-modes and result in preferentially tangential gas motions, due to strong restoring buoyancy forces in the vertical direction; thus, magnetic fields could in principle become tangential even in the absence of thermal conduction and the HBI. However, this tangential velocity bias is also lifted for $Fr_{\\rm MHD} \\gsim \\mathcal{O}(1)$. This has several immediate consequences: \\begin{itemize} \\item{The boost in thermal conduction restores the heat flow from the hot outer layers to the central cool core. This is crucial to averting massive overcooling and heating the cluster cores in a stable fashion. Indeed, our runs with radiative cooling showed that a cooling catastrophe was averted even in the absence of AGN heating. Thus, naive models which assume thermal conductivity with Spitzer fraction $f\\sim 1/3$ may perhaps be reasonably applied to clusters after all. Since conduction is regulated by turbulence, it will in general be time-dependent. A sudden boost in thermal conduction can modulate a transition between a cool core (CC) to a non cool core (NCC) cluster (\\citet{guo09}; we also see some hints of this in Fig \\ref{plot4}). Thus, mergers, while generally insufficient in the pure hydrodynamic case to effect a CC to NCC metamorphosis \\citep{poole08}, may be effective once the boost in thermal conduction due to additional turbulence is taken into account. If AGN are the main source of turbulence, this implies feedback between accretion onto the AGN and thermal conduction. Thus, AGN could exert another level of self-regulation and thermostatic control beyond straight kinetic heating. } \\item{Our predictions might be testable in the future: magnetic topology can be measured by radio polarization measurements (e.g., Pfrommer \\& Dursi 2009), while turbulence can be measured by a high spectral-resolution calorimeter. Alternatively, our predictions can be used to provide indirect constraints on turbulence. For instance, the absence of large-scale ordered fields would require a minimal level of turbulence, and argue against a fully relaxed ICM. The resulting turbulent pressure support affects mass estimates which assume hydrostatic equilibrium \\citep{lau09}, and the use of clusters for cosmology.} \\item{The distribution of metals in clusters is known to be broader than that of the stars. As shown by \\citet{sharma09,sharma09a}, metal mixing in a stratified plasma is much more effective once conduction is at play, because of the weaker nature of restoring buoyancy forces. This allows broad metallicity profiles to be obtained without inversion of the central entropy profile (which is not observed). The restoration of thermal conduction by stirring motions greatly aids this process. In the limit where conduction is so efficient that the cluster becomes isothermal (so that $\\omega_{\\rm BV}^{\\rm HBI} \\rightarrow 0$) or conduction becomes isotropic, fluid elements have the same temperature (and hence density) as their surrounding, and fluid elements become neutrally buoyant. This makes metal mixing maximally efficient. We note that any mechanism which disperses metals over wide distances (e.g., turbulent diffusion, Rebusco et al. 2008) generally excites turbulent velocities sufficient to isotropize the magnetic field. Thus, the relatively broad metallicity profiles in clusters is consistent with non-negligible stirring.} \\end{itemize} Our stirring methods ensure by construction that the ensuing turbulence is volume-filling. However, this is by no means guaranteed in nature. For instance, galactic wakes will be relatively narrow, with a cross-section of order the galaxy size; similarly, rising AGN bubbles will not induce velocity fluctuations over a large solid angle. A straightforward way to ensure volume-filling turbulence, as is required to stem the HBI, is to excite trapped $g$-modes which are repeatedly reflected and focused within a resonance region where $\\omega < \\omega_{\\rm BV}^{\\rm MHD}$, as discussed in \\S\\ref{section:turbulence}. At the same time, $\\omega > \\omega_{\\rm BV}^{\\rm MHD}$ is required to overwhelm buoyancy forces, an apparently contradictory requirement. Of course, the assumption of a single frequency is too simplistic: stirring motions contain many harmonics and will be scale dependent, generally increasing in frequency as turbulence cascades toward smaller scales. Thus, large scale $g$-modes could fall below the \\brunt frequency, ensuring trapping and amplification, but the resulting small scale turbulence could potentially still have turnover frequencies exceeding $\\omega_{\\rm BV}^{\\rm MHD}$. We will examine this in future work. Although we have focused on the HBI in this paper, the same Froude number considerations apply with equal force to the MTI unstable outer regions of the cluster, where temperature falls with radius and the MTI causes field lines to become preferentially radial. Here, if $Fr^{\\rm MHD} > \\mathcal{O}(1)$, turbulent motions can also overwhelm buoyant forces and randomize the field. Recently, there has been tantalizing evidence from the polarized emission surrounding the magnetic drapes of galaxies sweeping up magnetic fields in the Virgo cluster, that outside a central region, the magnetic field might be preferentially oriented radially \\citep{pfrommer09}, as one might expect from the MTI. If this result continues to hold up, this would imply the Froude number is below the critical value at these radii (thus providing an indirect constraint on turbulent velocities), or additional physical processes, not modeled here, are at play. \\\\" }, "0911/0911.3820.txt": { "abstract": "The present operation of the ground-based network of gravitational-wave laser interferometers in ``enhanced'' configuration and the beginning of the construction of second-generation (or advanced) interferometers with planned observation runs beginning by 2015 bring the search for gravitational waves into a regime where detection is highly plausible. The development of techniques that allow us to discriminate a signal of astrophysical origin from instrumental artefacts in the interferometer data and to extract the full range of information are therefore some of the primary goals of the current work. Here we report the details of a Bayesian approach to the problem of inference for gravitational wave observations using a network (containing an arbitrary number) of instruments, for the computation of the Bayes factor between two hypotheses and the evaluation of the marginalised posterior density functions of the unknown model parameters. The numerical algorithm to tackle the notoriously difficult problem of the evaluation of large multi-dimensional integrals is based on a technique known as Nested Sampling, which provides an attractive (and possibly superior) alternative to more traditional Markov-chain Monte Carlo (MCMC) methods. We discuss the details of the implementation of this algorithm and its performance against a Gaussian model of the background noise, considering the specific case of the signal produced by the in-spiral of binary systems of black holes and/or neutron stars, although the method is completely general and can be applied to other classes of sources. We {also} demonstrate the utility of this approach by introducing a new coherence test to distinguish between the presence of a coherent signal of astrophysical origin in the data of multiple instruments and the presence of incoherent accidental artefacts, and the effects on the estimation of the source parameters as a function of the number of instruments in the network. ", "introduction": "\\label{s:intro} Searches for gravitational waves are entering a crucial stage with the network of ground-based laser interferometers -- LIGO~\\cite{BarishWeiss:1999, lsc-s5-instr}, Virgo~\\cite{virgo} and GEO\\, 600~\\cite{geo} -- now fully operational and engaged in a new year-long data taking period~\\cite{Smith:2009,eligo} at ``enhanced'' sensitivity, which may allow the first direct detection of gravitational radiation. Construction has already begun for the upgrade of the instruments to advanced configuration (second generation interferometers) with installation at the sites that will start at the end of 2011~\\cite{Smith:2009,advligo,advvirgo}. {When} science observations resume at much improved sensitivity {by 2015}, several gravitational-wave events are expected to be observed, opening a new means to explore a variety of astrophysical phenomena (see \\emph{e.g.} Ref.~\\cite{Cutler:2002me,Kokkotas:2008} and references therein). Coalescing binary systems of compact objects -- black holes and neutron stars -- will be the workhorse source for gravitational wave observations. Ground-based laser interferometers will monitor the last seconds to minutes of the coalescence of these systems. The theoretical modelling of the (in-spiral) waveform is well in hand (see \\emph{e.g.}~\\cite{Blanchet:2006} and references therein), and the search algorithms are well understood~\\cite{lsc-cbc-s5-12-to-18, lsc-cbc-s5y1, lsc-cbc-spins, lsc-cbc-s3s4, lsc-cbc-s2-bbh, lsc-cbc-primordial-bh, lsc-cbc-s2-bns, lsc-cbc-s1}. The detection rate for on-going searches and observations with second generation instruments is estimated to lie in the range $9\\times 10^{-5}\\,\\mathrm{yr}^{-1} - 0.7\\,\\mathrm{yr}^{-1}$ and $0.2\\,\\mathrm{yr}^{-1} - 1000\\,\\mathrm{yr}^{-1}$, respectively, see~\\cite{cbc-rates} for a review. It is likely that in a few years time ground-based laser interferometers will allow us to extract a wealth of new information ranging from the formation and evolution of binary stars, the nature of precursors of (short) gamma-ray bursts, dynamical processes in star clusters, and could yield a new set of standard candles for precise cosmography. As instruments are beginning to operate at a meaningful sensitivity from an astrophysical, cosmological and fundamental physics point of view, much emphasis is now being placed on the development of methods that offer the maximum discriminating power to separate {disturbances} of instrumental origin from a true astrophysical signal, and to extract the full range of information from the detected signals. Bayesian inference provides a powerful approach to both model selection (or hypothesis testing) and parameter estimation. Despite the conceptual simplicity of the Bayesian framework, there has been only limited use of these methods for ground-based gravitational-wave data analysis due to their computational burden, in this case related to the need to compute large multi-dimensional integrals. Additionally, the Gaussian likelihood functions considered so far do not address the instrumental glitches which are present in data from the current generation of gravitational wave detectors. Here we present an efficient method to compute concurrently the full set of quantities at the heart of Bayesian inference: the Bayes factor between competing hypotheses and the posterior density functions (PDFs) on the relevant model parameters. The method is based on the \\emph{Nested Sampling} algorithm~\\cite{Skilling:AIP,Skilling:web,Sivia} to perform multi-dimensional integrals, that present the practical and computationally intensive challenge for the implementation of Bayesian methods. We demonstrate the algorithm by considering multi-detector observations of gravitational waves generated during the in-spiral phase of the coalescence of a binary system, modelled using the restricted post$^{2.0}$-Newtonian stationary phase approximation, which is the waveform used so far for searches of non-spinning binary objects~\\footnote{We note that searches for low-mass (that is with total mass $\\le 35\\,M_\\odot$) binary systems in the data now collected by LIGO and Virgo are based on templates computed at the post$^{3.5}$-Newtonian order. All the results presented here apply directly to that case at no additional computational costs.}. Initial results based on this method and applied to simplified gravitational waveforms were reported in~\\cite{VeitchVecchio:2008a,VeitchVecchio:2008b}. An application to the study of different waveform approximants to detect and estimate the parameters of signals generated through the numerical integration of the Einstein's equations for the two body problem in the context of the Numerical INJection Analysis (NINJA) Project was reported in~\\cite{CadonatiElAl:2009,AylottEtAl:2009,AVV:2009}. In this paper we: \\begin{itemize} \\item Provide for the first time full details about the theoretical and technical issues on which the computation of the evidence is based; \\item Discuss the errors associated with the computation of the integrals and the associated computational costs; \\item Show how from the nested samples one can construct at negligible computational cost the {marginalised} posterior PDFs on the source parameters; \\item Demonstrate the performance of this technique {in detecting a binary in-spiral signal against a Gaussian model of background noise in coherent observations using a network of detectors, and} by introducing a new {\\emph{coherence test}} to distinguish between the presence of a coherent signal of astrophysical origin in the data of multiple instruments and the presence of incoherent accidental artefacts, and \\item {Show the effects of the number of instruments in the network on the estimation of the source parameters. } \\end{itemize} The method in its present implementation can be extended in a straightforward way to the full coalescence waveform of binary systems -- in fact the first example applied on the NINJA data set is described in~\\cite{AylottEtAl:2009,AVV:2009} -- and the software is available as part of the LSC Analysis Library Applications (LALapps)~\\cite{lal,lalapps}. Work is already on-going in this direction. Furthermore, the approach can also be extended to other gravitational-wave signals. The paper is organised as follows: in Section~\\ref{s:modelselection} we review the key concepts of Bayesian inference and the signal generated by in-spiralling gravitational wave signals; in Section~\\ref{s:NSalgo} we describe our implementation of the numerical computation of multi-dimensional integrals for the computation of the marginal likelihood and marginalised posterior density functions based on {nested-sampling} the limited likelihood function; Section~\\ref{s:NSimplementation} contains the implementation details of the algorithm, including a quantification of the errors that affect the results as a function of the choice of the main tuning parameters of the algorithm and the effects on the scaling of the computational costs. Section~\\ref{s:results} contains the results from the applications of this method to several test cases, including a Bayesian coherence test that we introduce here for the first time. Throughout the paper we use geometric units, in which $G = c = 1$. ", "conclusions": "By taking a Bayesian approach to the analysis of data for the detection and characterisation of in-spiral signals, we have been able to implement a conceptually simple yet flexible framework for drawing inference from observations. Although the Bayesian formalism calls for the evaluation and integration of high-dimensional likelihood functions, we have shown that the nested sampling {technique} provides us with a means to both search for and estimate the parameters of a signal. Further work remains to be done in improving the efficiency and reliability of detection at low masses, but the particular implementation that we describe here provides a solid basis for these future improvements. We have used our implementation to demonstrate the power of Bayesian model selection in classifying putative gravitational wave signals, through the use of the coherence test described in Section \\ref{sec:coherent}. The coherence test goes some way to implementing a robust and Bayesian defence against glitches present in gravitational wave data which do not resemble coherent gravitational waves by providing an internal consistency check within the detector network. Tests of this kind may provide a useful new additional discriminator when analysing candidate gravitational waves, and are only achievable through the treatment of the detector network in a coherent way. Notably, this test is only possible through the use of the Occam factor and the comparison of probability distributions which differ in their dimensionality, and could not be possible with point estimates of maximum likelihood. Indeed, the maximum likelihood of the coherent and incoherent models can be trivially shown to be identical. In addition to such tests, it would also be desirable to model the detector data in a way which is non-stationary or Gaussian, but this is beyond the scope of this work. Furthermore, the use of the Bayesian parameter estimation framework is invaluable in inferring the signal parameters, and produces the full posterior probability distribution, not just maximum likelihood points and estimates of variance. The covariance and interdependence of the parameters does not prevent us from calculating consistent joint PDFs on the full parameter space even when point estimates have little meaning, and this allows us to see the benefit of using a network of detectors in a coherent fashion. {One of} the main benefits in the use of the nested sampling and Bayesian evidence approach is the ease with which it can be extended or adapted to different signal models with minimal changes needed to the implementation. In other work past and ongoing, we have shown how this method may be applied to the discrimination between candidate waveforms when comparing to a numerical relativity simulation; how one may place bounds on the Compton wavelength of the graviton given a single or multiple observations of an in-spiral signal, and testing the effects of including spins in the detection and estimation of in-spiral signals \\cite{AylottEtAl:2009,WDP:lambdaG}\\,. Future work on the implementation of the core algorithm will focus on achieving a better reliability and efficiency in the sampling of the parameter space, which should lead to further improvements in the performance in a real world situation, and on testing on the full set of waveform approximants used in present searches for coalescing binary systems. The work presented here forms the basis of the \\texttt{ lalapps\\_inspnest } program, which is can be found in the LALApps software distribution, released under the terms of the GNU General Public Licence \\cite{lalapps}." }, "0911/0911.3176_arXiv.txt": { "abstract": "We report on the properties of pre-main-sequence objects in the Taurus molecular clouds as observed in 7 mid- and far-infrared bands with the Spitzer Space Telescope. There are 215 previously-identified members of the Taurus star-forming region in our $\\sim$44 square degree map; these members exhibit a range of Spitzer colors that we take to define young stars still surrounded by circumstellar dust (noting that $\\sim$20\\% of the bonafide Taurus members exhibit no detectable dust excesses). We looked for new objects in the survey field with similar Spitzer properties, aided by extensive optical, X-ray, and ultraviolet imaging, and found 148 candidate new members of Taurus. We have obtained follow-up spectroscopy for about half the candidate sample, thus far confirming 34 new members, 3 probable new members, and 10 possible new members, an increase of 15-20\\% in Taurus members. Of the objects for which we have spectroscopy, 7 are now confirmed extragalactic objects, and one is a background Be star. The remaining 93 candidate objects await additional analysis and/or data to be confirmed or rejected as Taurus members. Most of the new members are Class II M stars and are located along the same cloud filaments as the previously-identified Taurus members. Among non-members with Spitzer colors similar to young, dusty stars are evolved Be stars, planetary nebulae, carbon stars, galaxies, and AGN. ", "introduction": "\\label{sec:intro} A complete inventory of all the coeval stars in a young stellar association, cluster, or group (hereafter ``association\") enables studies of the initial mass function (IMF), disk fraction, and stellar rotational properties, among other pursuits. Information from associations with a range of ages enables understanding of the overall formation and evolution of young stars, including the change with time of disk fraction and stellar rotation rate. However, identifying all of the member stars of a given young association can be quite difficult. (By ``member,'' we mean objects that are clearly young, close to the same age, and often still associated with their natal cloud.) Finding all such members requires that one employ multiple observational techniques. These methods include but are not limited to X-ray surveys (e.g., Alcal\\'a \\etal\\ 1996, Wolk \\etal\\ 2006), H$\\alpha$ surveys (e.g., Ogura \\etal\\ 2002), variability surveys (e.g., Carpenter \\etal\\ 2001, Rebull 2001), ultraviolet (UV) surveys (e.g., Rebull \\etal\\ 2000), and infrared (IR) surveys (e.g., J{\\o}rgensen \\etal\\ 2006, Rebull \\etal\\ 2007). At each wavelength, we can use the fact that stellar youth implies more flux at a given radiometric band (X-rays, H$\\alpha$, UV, IR), or more flux variability, than older stars of comparable mass, allowing separation of association members from field contaminants. When we combine the information from surveys in multiple wavelengths, we must remember that the influence of extinction due to circumstellar matter and/or the molecular cloud is vastly different at different wavelengths; extinction affects UV and optical wavelengths much more strongly than IR. For young associations, many, but not all, legitimate members are identifiable using just one or a few survey methods. There are advantages and disadvantages to studying nearby young associations. Identification of young members in an association is made easier if the objects are located at smaller heliocentric distances and therefore brighter on the whole. This is especially the case for low-mass members with very low luminosities; a complete census of these stars and brown dwarfs can be made only for very nearby star-forming regions. On the other hand, nearby star-forming molecular clouds cover large areas of sky and therefore require large investments of observing time. At just 137 pc (with a depth of $\\sim$ 20 pc; Torres \\etal\\ 2007, 2009), the Taurus star-forming region is one of the closest large cloud complexes with hundreds of low-mass stars, ongoing star formation, and objects ranging in age up to $\\sim$5 Myr. Studies of Taurus objects have significantly influenced our basic understanding of the star-formation process for decades (e.g., Herbig \\& Rao 1972, Kenyon \\etal\\ 2008). However, the Taurus Molecular Cloud is close enough that it subtends more than 100 square degrees of sky; surveying all or even most of it is difficult within typical telescope time allocations. Infrared surveys led to the discovery that some stars have infrared excesses, interpreted as circumstellar matter (e.g., Aumann \\etal\\ 1984, Beichman \\etal\\ 1986). Most if not all low-mass stars form with circumstellar accretion disks, resulting in IR excesses for as long as the dusty circumstellar material survives (e.g., Hernandez \\etal\\ 2007). By using IR to survey a star-forming region, the stars with IR excesses are relatively easily distinguished from stars without such excesses. The Spitzer Space Telescope (Werner \\etal\\ 2004) provides an excellent platform for surveying star-forming regions in the mid-IR and far-IR, enabling stars with IR excesses to be identified; the member stars which do not have IR excesses must be recovered using different techniques such as the ones listed above. Because Spitzer relatively efficiently maps large regions of sky, it is a particularly useful tool for surveying large star forming regions. We have conducted a large multi-wavelength imaging and spectroscopic survey of the Taurus Molecular Cloud (TMC) in order to test if our inventory of Taurus members with infrared excesses is, in fact, complete. The Spitzer imaging component is referred to as the Taurus Spitzer Survey, and is described by Padgett \\etal\\ (2008; hereafter P08) and Padgett \\etal\\ (2009; hereafter P09). It covers $\\sim$44 square degrees from 3.6 to 160 \\mum\\ and is a Spitzer Legacy Project, so enhanced data products have been delivered back to the Spitzer Science Center (SSC), including the catalogs on which this present paper is based. In addition to the Spitzer component, there are four other major components to our Taurus survey. XMM-Newton was used by the XMM-Newton Extended Survey of the Taurus Molecular Cloud (XEST) program (e.g., G\\\"udel \\etal\\ 2007 and references therein), which mapped $\\sim$5 square degrees, most of which was also mapped by the Spitzer observations; the XEST data include X-ray imaging but also include ultraviolet data from the XMM-Newton Optical Monitor (Audard \\etal\\ 2007). XEST was deliberately pointed towards aggregates of previously identified Taurus members. In the optical, the Canada-France-Hawaii Telescope (CFHT) survey (Monin \\etal\\ in preparation; G\\\"udel, Padgett, \\& Dougados 2007) mapped $\\sim$28 square degrees (all of which are encompassed by the Spitzer area), and the Sloan Digital Sky Survey (SDSS) (Finkbeiner \\etal\\ 2004; Padmanabhan \\etal\\ 2008) mapped $\\sim$48 square degrees in two perpendicular strips, about half of which overlaps the Spitzer area. Finally, the Five College Radio Astronomy Observatory (FCRAO) millimeter wavelength survey (Goldsmith \\etal\\ 2008) mapped $\\sim$100 square degrees in the CO(1-0) line, covering the Spitzer survey area entirely. The relative coverages of these surveys is shown in P09. Our extended collaboration has already begun to use this rich dataset to search for new members of Taurus; Scelsi \\etal\\ (2007,2008) identified new candidate members using the XEST data, and Guieu \\etal\\ (2006, 2007) identified new brown dwarf members using the CFHT data to study their disk properties using the Spitzer data. Ongoing investigations include searches for members via emission line spectra (Knapp \\etal, in prep), Herbig-Haro (HH) objects (Stapelfeldt \\etal\\ in prep), and transition disks (McCabe \\etal\\ in prep). In this paper, we select new candidate Taurus members with infrared excesses using Spitzer and Two-Micron All-Sky Survey (2MASS; Skrutskie \\etal\\ 2006) data. We construct color-magnitude and color-color diagrams for point sources, then use the locations of previously-identified young stars in these diagrams to select new candidate members with infrared excesses. To discard obvious extragalactic sources, we examine the source morphology in all available bands of the multiwavelength Taurus survey. We construct spectral energy distributions (SEDs) from the photometry over all available bands, again discarding objects we believe to be galaxies. Follow-up spectroscopy has been obtained to assess whether or not our new candidate Taurus objects are likely Taurus members. We also present Spitzer flux densities for the 215 previously-identified members found in the region covered by our Spitzer survey. Note that (a) the 215 previously-identified objects include those members without infrared excesses identified via other mechanisms; (b) the 215 previously-identified objects are those covered by our Spitzer map -- there are other legitimate Taurus members outside the region we observed, such as in L1551; (c) our new candidate member list is necessarily just those with IR excesses and exclusively within the regions covered by our Spitzer observations. Some objects are resolved in one or more of the Spitzer images, and extended source photometry may be a better representation of the complete flux from the object; many of the extended sources are discussed individually in other papers (e.g., Tobin \\etal\\ 2008, Stapelfeldt \\etal\\ in prep) The observations, data reduction, and ancillary data are described in \\S\\ref{sec:obs}. Section~\\ref{sec:pickysos} describes our young stellar object (YSO) selection process which is based on the colors of previously-identified Taurus members, also presented here. We describe 34 objects that we have, thus far, identified as new members of Taurus (plus 3 probable new members and 10 possible new members) in \\S\\ref{sec:discussion} and discuss the properties of the new objects in conjunction with (and comparison to) the previously-identified members of Taurus. Finally, we summarize our main points in \\S\\ref{sec:concl}. The Appendix contains spectral energy distributions and discussion of some specific objects. ", "conclusions": "\\label{sec:concl} We have presented here Spitzer flux densities for 215 previously-identified members of the Taurus Molecular Cloud young stellar object population. We constructed Spitzer color-color and color-magnitude diagrams, investigated where the previously identified Taurus members were located, and then used those diagrams to select additional candidate Taurus objects out of a catalog of $\\sim$700,000 objects observed with Spitzer over 44 square degrees. We used a wealth of supporting data (including ground-based optical imaging) to winnow that list down to 148 candidate new Taurus members. We obtained follow-up optical spectroscopy for about half the sample, thus far finding 34 new members, 3 probable new members, and 10 possible new members, a potential increase of 15-20\\% of young Taurus members. Most of the new members are located in close (projected) proximity to the previously-identified Taurus members; most of them are Class II M stars. In addition to the new members in our sample, there are 60 stars needing additional follow-up observations, and 33 objects pending any follow-up observations at all, so more Taurus members may yet be confirmed out of our list of candidate members. We also found a background Be star, a new planetary nebula, a new carbon star, many background giants, and 100s of galaxies, just 7 of which made it into our final list of 148 YSO candidates. As part of this project's classification as a Spitzer Legacy Project, enhanced data products have been delivered back to the Spitzer Science Center (SSC), including the catalogs on which this present paper is based. This study has demonstrated the unique power of Spitzer to efficiently survey large areas of the sky and provide new information on membership and YSO properties even in nearby star forming regions such as Taurus, which has been extensively studied for decades. Even after many decades of study, our knowledge of membership in Taurus is still incomplete, but definitely improving, thanks to Spitzer. \\appendix" }, "0911/0911.4560_arXiv.txt": { "abstract": "Both analytical and numerical works show that magnetic reconnection must occur in hot accretion flows. This process will effectively heat and accelerate electrons. In this paper we use the numerical hybrid simulation of magnetic reconnection plus test-electron method to investigate the electron acceleration and heating due to magnetic reconnection in hot accretion flows. We consider fiducial values of density, temperature, and magnetic parameter $\\beta_e$ (defined as the ratio of the electron pressure to the magnetic pressure) of the accretion flow as $n_{0} \\sim 10^{6} {\\rm cm^{-3}}$, $T_{e}^0\\sim 2\\times 10^9 {\\rm K}$, and $\\beta_e=1$. We find that electrons are heated to a higher temperature $T_{e}=5\\times 10^9$K, and a fraction $\\eta\\sim 8\\%$ of electrons are accelerated into a broken power-law distribution, $dN(\\gamma)\\propto \\gamma^{-p}$, with $p\\approx 1.5$ and $4$ below and above $\\sim 1$ MeV, respectively. We also investigate the effect of varying $\\beta$ and $n_0$. We find that when $\\beta_e$ is smaller or $n_0$ is larger, i.e, the magnetic field is stronger, $T_e$, $\\eta$, and $p$ all become larger. ", "introduction": "Magnetic reconnection serves as a highly efficient engine which converts magnetic energy into plasma thermal and kinetic energy. The first theoretical model for magnetic reconnection was proposed by Sweet and Parker \\citep{Parker 1958}. Ever since then, theoretical studies along this line have been reported in various papers including analytical theories and numerical simulations \\citep{Parker 1963, Drake 2005, Litvinenko 1996}. Magnetic reconnection has been widely used to explain explosive phenomena in space and laboratory plasmas such as solar flares \\citep{Somov 1997}, the heating of solar corona \\citep{Cargill 1997}, and substorms in the Earth's magnetosphere \\citep{Bieber 1982, Baker 2002}. Magnetic reconnection also plays an important role in high-energy astrophysical environments such as the magnetized loop of the Galactic center \\citep{Heyvaerts 1988}, jets in active galactic nuclei \\citep{Schopper 1998, Larrabee 2003, Lyutikov 2003}, and $\\gamma$ ray bursts \\citep{Michel 1994}. For example, by assuming that particle acceleration is due to direct current (DC) electric field and is balanced by synchrotron radiative losses, Lyutikov (2003) found that the maximum energy of electrons is $\\sim 100$ TeV. Larrabee calculated the self-consistent current density distribution in the regime of collisionless reconnection at an X-type magnetic neutral point in relativistic electron-positron plasma and studied the acceleration of non-relativistic and relativistic electron populations in three-dimensional tearing configurations. But so far there is very few work on the magnetic reconnection in hot accretion flows. In accretion flows, the gravitational energy is converted into turbulent and magnetic energy due to the magnetorotational instability \\citep[MRI;][]{Balbus 1991}. Both analytical works \\citep[e.g.,][]{Quataert 1999, Goodman 2008, Uzdensky 2008} and three-dimensional magnetohydrodynamic simulations of black hole accretion flow \\citep{Hawley 2002, Machida 2003, Hirose 2004} have shown that magnetic reconnection is unavoidable and plays an important role in converting the magnetic energy into the thermal energy of particles. In fact, turbulent dissipation and magnetic reconnection are believed to be the two main mechanisms of heating and accelerating particles in hot accretion flows. In the case of large $\\beta$ ($\\beta$ is defined as the ratio of the gas pressure to the magnetic pressure), the magnetic reconnection is even likely to the dominant mechanism of electrons heating and acceleration \\citep[][]{Quataert 1999, Bisnovatyi-Kogan 1998}. On the observational side, flares are often observed in black hole systems. One example is the supermassive black hole in our Galactic center, Sgr A*, where infrared and X-ray flares occur almost every day \\citep[e.g.,][]{Baganoff 2001, Hornstein 2007, Dodds-Eden 2009}. To explain the origin of these flares, we often require that some electrons are transiently accelerated into a relativistic power-law distribution by magnetic reconnection within the accretion flow \\citep[e.g.,][]{Baganoff 2001, Yuan 2003, Yuan 2004, Dodds-Eden 2009, Eckart 2009} or in the corona above the accretion flow (e.g., Yuan et al. 2009), and their synchrotron or synchrotron-inverse-Compton emission will produce the flares. Regarding the study of the electron acceleration by reconnection, recently many works have been done on the relativistic magnetic reconnection with the approach of Particle-in-Cell (PIC) simulation. \\cite{Zenitani 2007} presented the development of relativistic magnetic reconnection, whose outflow speed is on the order of the light speed. It was demonstrated that particles were strongly accelerated in and around the reconnection region and that most of the magnetic energy was converted into a ``non-thermal'' part of plasma kinetic energy. Magnetic reconnection during collisionless, stressed, X-point collapse was studied using kinetic, 2.5D, fully relativistic PIC numerical code in \\cite{Tsiklauri 2007} and they also found high energy electrons. PIC electromagnetic relativistic code was also used in \\cite{Karlicky 2008} to study the acceleration of electrons and positrons. They considered a model with two current sheets and periodic boundary conditions. The electrons and positrons are very effectively accelerated during the tearing and coalescence processes of the reconnection. All these work investigate the circumstance of the Sun. In this paper we use the numerical hybrid simulation of magnetic reconnection plus test-electron method to investigate the electron heating and acceleration by magnetic reconnection in hot accretion flows. We would like to note that the results should also be applied to corona of accretion flow and jet, if the properties of the plasma in those cases are similar to what we will investigate in the present paper. Self-consistent electromagnetic fields are obtained from the hybrid code, in which ions are treated as discrete particles and electrons are treated as massless fluid. Thereafter, test electrons are placed into the fields to study their acceleration. The input parameters include density, temperature and magnetic field of the accretion flow. The values of density and temperature of astrophysical accretion flows can differ by several orders of magnitude among various objects. Here we take the accretion flow in Sgr A* as reference \\citep[see][for details]{Yuan 2003}, but we also investigate the effect of varying parameters. For the strength of the magnetic field, we consider mainly $\\beta_{e}=1$. Here $\\beta_{e}$ is defined as the ratio of the electron pressure to magnetic pressure, $\\beta_{e}\\equiv P_{e}/P_{\\rm mag}$. Note that what we usually use in hot accretion flows is the ratio of the gas pressure and the magnetic pressure, $\\beta\\equiv P_{gas}/P_{\\rm mag}$. The hot accretion flow is believed to be two-temperature, with $T_{i} \\gg T_{e}$. For a reasonable value of $T_{i}=10T_{e}$ \\citep[ref.][]{Yuan 2003}, the above $\\beta_{e}$ corresponds to the usual $\\beta =10$. This is the typical value in the hot accretion flow as shown by MHD numerical simulations \\citep[e.g., ][]{Hirose 2004}. Given that the value of $\\beta$ is very inhomogenous in the accretion flow \\citep{Hirose 2004}, we also consider $\\beta_{e}=10, 0.1$, and $0.01$. In Section 2, we use the numerical hybrid simulation of magnetic reconnection to get the structure of self-consistent electric and magnetic fields. In Section 3 we use the obtained electromagnetic fields as the background of test electrons to investigate the electron acceleration. Finally we summarize our results and discuss the application in interpreting the flares in Sgr A* in Section 4. ", "conclusions": "In this paper we use hybrid simulation code of magnetic reconnection plus test-electron method to study the electrons heating and acceleration by magnetic reconnection process in hot accretion flows. The self-consistent electromagnetic fields are obtained from the hybrid simulation and they are then used in the test-electron calculation. The fiducial parameters of the background accretion flows we adopt are density $n_{e}=10^6 {\\rm cm}^{-3}$ and electron temperature $T_{e}^0=2\\times 10^{9.3}$K (ions temperature $T_{i}=10 T_{e}^0$). These values are taken from the accretion flow in Sgr A*, the supermassive black hole in our Galactic center \\citep[ref.][]{Yuan 2003}. For the strength of the magnetic field in the accretion flow, we set $\\beta_{e} (\\equiv P_{e}/P_{\\rm mag})=1.0$. This is the typical value according to the MHD numerical simulation of hot accretion flow. We find that the electrons are heated and accelerated by the reconnection. The new distribution can be fitted by a Maxwellian distribution plus a broken power-law. The temperature is increased to $10^{9.65}$K and the power-law is described by $dN(E)\\propto E^{-1.47}$ between 1 and 10 MeV and $dN(E)\\propto E^{-4}$ between 10 MeV and 40 MeV. The fraction of electrons with energy above 1 MeV is $\\sim 8\\%$. Given that the magnetic field and density in accretion flow are inhomogeneous, we also consider $\\beta_{e}=10, 0.1$, and $0.01$ and $n_{e}=10^7 {\\rm cm}^{-3}$. We find that the results are qualitatively similar, namely the distribution of electrons can be fitted by a Maxwellian one with a higher ``heated'' temperature plus a broken power-law. When the magnetic field is stronger ($\\beta_{e}$ is smaller or $n_{e}$ is larger), the heating and acceleration become more efficient. The ``heated'' temperature is higher and more electrons are accelerated. In addition, the ``low-energy'' part of the broken power-law becomes softer (refer to Table 1). In our work we assume ideal MHD for the hybrid simulation, thus the Joule heating to the accretion flow is neglected. Obviously the electrons will be heated to a higher temperature when this effect is considered." }, "0911/0911.1459_arXiv.txt": { "abstract": "Almost 80 years have passed since Trumpler's analysis of the Galactic open cluster system laid one of the main foundations for understanding the nature and structure of the Milky Way. Since then, the open cluster system has been recognised as a key source of information for addressing a wide range of questions about the structure and evolution of our Galaxy. Over the last decade, surveys and individual observations from the ground and space have led to an explosion of astrometric, kinematic and multiwavelength photometric and spectroscopic open cluster data. In addition, a growing fraction of these data is often time-resolved. Together with increasing computing power and developments in classification techniques, the open cluster system reveals an increasingly clearer and more complete picture of our Galaxy. In this contribution, I review the observational properties of the Milky Way's open cluster system. I discuss what they can and cannot teach us now and in the near future about several topics such as the Galaxy's spiral structure and dynamics, chemical evolution, large-scale star formation, stellar populations and more. ", "introduction": "In 1930, Robert Trumpler \\nocite{Trumpler1930} published his seminal paper in which he proves the existence of the interstellar medium (ISM). In that paper, Trumpler showed how distances to open clusters (OCs) derived from the apparent magnitudes of cluster stars of known spectral types are systematically greater than those derived from their apparent sizes, and how the effect increases with distance. He correctly attributed this phenomenon to the presence of an intervening medium. This was the first paper using OCs as tracers of Galactic structure. ", "conclusions": "" }, "0911/0911.2493_arXiv.txt": { "abstract": "We describe an exact, flexible, and computationally efficient algorithm for a joint estimation of the large-scale structure and its power-spectrum, building on a Gibbs sampling framework and present its implementation \\textsc{ARES} (Algorithm for REconstruction and Sampling). \\textsc{ARES} is designed to reconstruct the 3D power-spectrum together with the underlying dark matter density field in a Bayesian framework, under the reasonable assumption that the long wavelength Fourier components are Gaussian distributed. As a result \\textsc{ARES} does not only provide a single estimate but samples from the joint posterior of the power-spectrum and density field conditional on a set of observations. This enables us to calculate any desired statistical summary, in particular we are able to provide joint uncertainty estimates. We apply our method to mock catalogs, with highly structured observational masks and selection functions, in order to demonstrate its ability to reconstruct the power-spectrum from real data sets, while fully accounting for any mask induced mode coupling. ", "introduction": "\\nextext{Throughout cosmic history a wealth of information on the origin and evolution of our Universe has been imprinted to the large scale structure via the gravitational amplification of primordial density perturbations. Harvesting this information from probes of the large scale structure, such as large galaxy surveys, therefore is an important scientific task to further our knowledge and to establish a conclusive cosmological picture. In recent years great advances have been made, both in retrieving huge amounts of data and increasing sensitivity in galaxy redshift surveys. Especially the recent galaxy surveys, the 2dF Galaxy Redshift Survey \\citep[][]{COLLESS2001} and the Sloan Digital Sky Survey \\citep[][]{SDSS7} provide sufficient redshifts to probe the 3D galaxy distribution on large scales. In particular, the two point statistics of the matter distribution contains valuable information to test standard inflation and cosmological models, which describe the origin and evolution of all observed structure in the Universe. Measuring the power-spectrum from galaxy observations therefore has always attracted great interest. Precise determination of the overall shape of the power-spectrum can for instance place important constraints on neutrino masses, help to identify the primordial power-spectrum, and break degeneracies for cosmological parameter estimation from CMB data \\citep[e.g.][]{hu-98,wmap-spergel,HANNESTAD2003,Efstathiou_2002,PERCIVAL2002,wmap-spergel,VERDE2003}. In addition, several characteristic length scales have been imprinted to the matter distribution throughout cosmic history, which can serve as new standard rulers to measure the Universe. A prominent example of these length scales is the sound horizon, which yields oscillatory features in the power-spectrum, the so called baryon \\nextext{acoustic} oscillations (BAO) \\citep[e.g.][]{SILK1968,PEEBLES1970,SUNYAEV1970}. Since the physics governing these oscillatory features is well understood, precise measurements of the BAO will allow us to establish a new, precise standard ruler to measure the Universe through the distance redshift relation \\citep[][]{BLAKE2003,SEO2003}. Precision analysis of large scale structure data therefore is a crucial step in developing a conclusive cosmological theory.} \\nextext{Unfortunately, contact between theory and observations cannot be made directly, since observational data is subject to a variety of systematic effects and statistical uncertainties. Such systematics and uncertainties arise either from the observational strategy or are due to intrinsic clustering behavior of the galaxy sample itself \\citep[][]{SANCHEZ2008}. Some of the most prominent uncertainties and systematics are:} \\begin{itemize} \\item survey geometry and selection effects \\item close pair incompleteness due to fiber collisions \\item galaxy biases \\item redshift space distortions \\end{itemize} \\nextext{The details of galaxy clustering, and how galaxies trace the underlying density field are in general very complicated. The bias between galaxies and mass density is most likely non-linear and stochastical, so that the estimated galaxy spectrum is expected to differ from that of the mass \\citep[][]{DEKEL1999}. Even in the limit where a linear bias could be employed, it still differs for different classes of galaxies \\citep[see e.g.][]{COLE2005}. In addition, the apparent density field, obtained from redshift surveys, will generally be distorted along the line-of-sight due to the existence of peculiar velocities.} \\nextext{ However, the main cause for the systematic uncertainties in large scale power-spectrum estimations is the treatment of the survey geometry \\citep[][]{TEGMARK1995,BALLINGER1995}. Due to the survey geometry the raw power-spectrum yields an expectation value for the power-spectrum, which is the true cosmic power-spectrum convolved with the survey mask \\citep[][]{COLE2005}. This convolution causes an overall distortion of the power-spectrum shape, and drastically decreases the visibility of the baryonic features.} \\nextext{ The problems, mentioned above, have been discussed extensively in literature, and many different approaches to power-spectrum analysis have been proposed. Some of the main techniques to recover the power-spectrum from galaxy surveys are Fourier transform based, such as the optimal weighting scheme, which assigns a weight to the galaxy fluctuation field, in order to reduce the error in the estimated power \\citep[see e.g.][]{FELDMAN1994,TEGMARK1995,HAMILTON1997A,YAMAMOTO2003,PERCIVAL2004}. Alternative methods rely on Karhunen-Lo\\`{e}ve decompositions \\citep[][]{TEGMARK1997,TEGMARK_2004,POPE2004} or decompositions into spherical harmonics, which is especially suited to address the redshift space distortions problematic \\citep[][]{FISHER1994,HEAVENS1995,TADROS1999,PERCIVAL2004,PERCIVAL2005}. In addition, to these deconvolution methods there exists a variety of likelihood methods to estimate the real space power-spectrum \\citep[][]{BALLINGER1995,HAMILTON1997A,HAMILTON1997B,TADROS1999,PERCIVAL2005}. In order to not just provide the maximum likelihood estimate but also conditional errors, \\cite{PERCIVAL2005} proposed a Markov Chain Monte Carlo method to map out the likelihood surface. } \\nextext{ Nevertheless, as the precision of large scale structure experiments has improved, the requirement on the control and characterization of systematic effects, as discussed above, also steadily increases. It is of critical importance to propagate properly the uncertainties caused by these effects through to the matter power-spectrum and cosmological parameters estimates, in order to not underestimate the final uncertainties and thereby draw incorrect conclusions on the cosmological model. } \\nextext{We therefore felt inspired to develope a new Bayesian approach to extract information on the two point statistics from a given large scale structure dataset. We prefer Bayesian methods to conventional likelihood methods, as the yield more general and profound statements about measurements. In example, conventional likelihood methods can only answer questions of the inner form like :\" Given the true value \\(s\\) of a signal, what is the probability distribution of the measured values \\(d\\)?\" A Bayesian method, on the other hand, answers questions of the type :\"Given the observations \\(d\\), what is the probability distribution of the true underlying signal \\(s\\)?\" For this reason, Bayesian statistics answers the underlying question to every measurement problem, of how to estimate the true value of the signal from observations, while conventional likelihood methods do not \\citep[][]{MICHEL1999}.} \\nextext{Since the result of any Bayesian method is a complete probability distribution they permit fully global analyses, taking into account all systematic effects and statistical uncertainties. In particular, here, we aim at evaluating the power-spectrum posterior distribution \\(\\mathcal{P}\\left(\\{P(k_i)\\}|\\{d_i\\}\\right)\\), with \\(P(k_i)\\) being the power-spectrum coefficients of the \\(k_i\\)th mode and \\(d_i=d(\\vec{x}_i)\\) is an observation at position \\(\\vec{x}_i\\) in three dimensional configuration space.} This probability distribution would then contain all information on the two point statistics supported by the data. In order to explore this posterior distribution we employ a Gibbs sampling method, previously applied to CMB data analysis \\citep[see e.g.][]{WANDELT2004,2004ApJS..155..227E,JEWELL2004}. \\nextext{ Since direct sampling from \\(\\mathcal{P}\\left(\\{P(k_i)\\}|\\{d_i\\}\\right)\\) is impossible or at least difficult, they propose instead to draw samples from the full joint posterior distribution \\(\\mathcal{P}\\left(\\{P(k_i)\\},\\{s_i\\}|\\{d_i\\}\\right)\\) of the power-spectrum coefficients \\(P(k_i)\\) and the 3D matter density contrast amplitudes \\(s_i\\) conditional on a given set of data points \\(\\{d_i\\}\\). This is achieved by iteratively drawing density samples from a Wiener-posterior distribution and power-spectrum samples via an efficient Gibbs sampling scheme (see figure \\ref{fig:flowchart} for an illustration). Here, artificial mode coupling, as introduced by survey geometry and selection function, is resolved by solving the Wiener-filtering equation, which naturally regularizes inversions of the observational response operator in unobserved regions.} In this fashion\\nextext{,} we obtain a set of Monte Carlo samples from the joint posterior, which allows us to compute any desired property of the joint posterior density, with the accuracy only limited by the sample size. In particular\\nextext{,} we obtain \\nextext{the power spectrum posterior} \\(\\mathcal{P}\\left(\\{P(k_i)\\}|\\{d_i\\}\\right)\\) by simply marginalizing \\nextext{the joint posterior} \\(\\mathcal{P}\\left(\\{P(k_i)\\},\\{s_i\\}|\\{d_i\\}\\right)\\) over the auxiliary density amplitudes \\(s_i\\), which is trivially achieved by ignoring the \\(s_i\\) samples. The Gibbs sampler also offers unique capabilities for propagating systematic uncertainties end-to-end. Any effect, for which we can define a sampling algorithm, either jointly with or conditionally on other quantities, can be propagated seamlessly through to the final posterior. It is worth noting, that our method differs from traditional methods of analyzing galaxy surveys in a fundamental aspect. Traditional methods consider the analysis task as a set of steps, each of which arrives at intermediate outputs which are then fed as inputs to the next step in the pipeline. Our approach is a truly global analysis, in the sense that the statistics of all science products are computed jointly, respecting and exploiting the full statistical dependence structure between various components. In this paper we present \\textsc{ARES} (Algorithm for REconstruction and Sampling), a computer algorithm to perform a full Bayesian data analysis of 3D redshift surveys. In section \\ref{LSS_SAMPLER} we give an introduction to the general idea of the large scale structure Gibbs sampling approach, followed by section \\ref{MAP_SAMPLER} and \\ref{PS_SAMPLER}, where we describe and derive in detail the necessary ingredients to sample the 3D density distribution and the power-spectrum respectively. The choice of the prior and the relevance for the cosmic variance are discussed in section \\ref{Prior_and_Variance}. Details concerning the numerical implementation are discussed in section \\ref{NUMERICAL_IMPLEMENTATION}. We then test \\textsc{ARES} thouroughly in section \\ref{Gaussian_Test_Cases}, particularly focussing on the treatment of survey masks and selection functions. In section \\ref{OPERATIONS_GIBBS_SAMPLES} we demonstrate the running median filter, and use it as an example to demonstrate how uncertainties can be propagated to all inferences based on the set of Gibbs samples. Finally we conclude in section \\ref{Conclusion}, by discussing the results of the method and giving an outlook for future extensions and application of our method. ", "conclusions": "\\label{Conclusion} In this work, we presented \\textsc{ARES}, a new Bayesian computer algorithm, designed to extract the full information on the two point statistics from any given probe of the three dimensional large scale structure. The scientific aim of this algorithm is to provide an estimate of the power-spectrum posterior \\(\\mathcal{P}\\left(\\{P(k_i)\\}|\\{d_i\\}\\right)\\), conditional on a data set, while accounting and correcting for all possible sources of uncertainties, such as survey geometry, selection effects and biases. This is achieved by exploring the power-spectrum posterior \\(\\mathcal{P}\\left(\\{P(k_i)\\}|\\{d_i\\}\\right)\\) via a Gibbs sampling approach. While direct sampling from the power-spectrum posterior is not possible, it is possible to draw samples from the full joint posterior distribution \\(\\mathcal{P}\\left(\\{P(k_i)\\},\\{s_i\\}|\\{d_i\\}\\right)\\) of the power-spectrum coefficients \\(P(k_i)\\) and the three dimensional matter density contrast amplitudes \\(s_i\\) conditional on a given set of data points \\(\\{d_i\\}\\). The entire Gibbs sampling algorithm therefore consists of two basic sampling steps, in which first a full three dimensional Wiener reconstruction algorithm is applied to the data and then a power-spectrum is drawn from the inverse gamma distribution. In this fashion we obtain a set of power-spectrum and signal amplitude samples, which provide a full representation of the full joint posterior distribution \\(\\mathcal{P}\\left(\\{P(k_i)\\},\\{s_i\\}|\\{d_i\\}\\right)\\). The scientific output of this Bayesian method therefore is not a single estimate but a complete probability distribution, enabling us to calculate any desired statistical summary such as the mean, mode or variance. \\begin{figure*} \\centering{\\resizebox{1.\\hsize}{!}{\\includegraphics{./figures/picture13}}} \\caption{Volume rendering of the ensemble variance (upper panels) and the ensemble mean (lower panels) obtained from the mock galaxy catalog analysis.} \\label{fig:DELUCIA_VARIANCE} \\end{figure*} \\begin{figure*} \\centering{\\resizebox{1.\\hsize}{!}{\\includegraphics{./figures/picture14}}} \\caption{Running median estimates of the power-spectra (upper panels) and the according wiggle functions (lower panels) for the set of Gibbs samples with the Jeffrey's prior (left panels), and the inverse gamma prior (right panels). The black lines represent the ensemble mean of the sample set and the light gray and dark gray shaded regions denote the one and two sigma confidence regions respectively. Additionally we show the according input power-spectra. The blue line shows the cosmological power-spectrum from which the matter field realization was drawn, and the red line is the power-spectrum of this specific matter field realization.} \\label{fig:RUNNING_MEDIAN} \\end{figure*} We also demonstrated, that given a set of Gibbs samples, it is possible to provide an analytic approximation to the power-spectrum posterior \\(\\mathcal{P}\\left(\\{P(k_i)\\}|\\{d_i\\}\\right)\\). This Blackwell-Rao estimator has an analytically appealing form enabling us to calculate any desired moment of the probability distribution in a simple analytic way. In addition, since the full joint probability distribution is available, it is easy to carefully propagate all uncertainties through to the result of further post-processing analysis steps, such as parameter estimation. In this work, we focused on thoroughly testing the performance and behavior of our method by applying it to simulated data with controlled properties. These tests were designed to highlight the problematic of survey geometry and selection effects, for the two cases of Gaussian random fields and a mock galaxy catalog based on the Millennium run. One of the main goals of these tests was to build up intuition on the phenomenological behavior of the Gibbs sampling algorithm, estimating particularly issues, such as the correlation length of the Gibbs chain, burn-in and convergence times. The result of these tests is of special relevance, as it shows how long the Gibbs sampling chain has to run in order to produce a sufficient amount of independent samples. Through these experiments we found that the longest correlation lengths are dominated by the poorly constrained Nyquist modes of the box, which can be easily alleviated by imposing some prior knowledge on these modes. In doing so we found that the maximal correlation length for the Gibbs chain was on the order of hundred Gibbs samples. Thus, creating a large number of independent samples in a full scale data analysis is numerically very well feasible. However, the most important result of these tests is, that our method is able to correct for artificial mode coupling due to the survey geometry and selection effects. This was tested by examining the correlation structure of the Gibbs samples, which showed that the maximal residual correlation can be reduced to less than \\(10 \\%\\), demonstrating that this method correctly accounts for geometry effects. The application of \\textsc{ARES} to a galaxy mock catalog, based on the Millennium run, demonstrated the ability of our method to capture the characteristics of the fully nonlinearly evolved matter field. This is owed to the fact, that the Wiener filter is a linear operation on the data, and as such does not destroy the intrinsic statistical characteristics of the data set. Nevertheless, a full real data analysis of existing redshift surveys requires the treatment of additional systematic effects such as scale or luminosity dependent biases or redshift space distortion corrections, which we defer to future works. However, the Bayesian framework, as presented here, can take all these effects naturally into account and treats them statistically fully consistent. Beside the possibility to include various kinds of uncertainties, the Gibbs sampling approach also allows for a natural joint analysis of different data sets, taking into account the systematics of each individual data set. In summary, we showed that \\textsc{ARES} is a highly flexible and adaptive machinery for large scale structure analysis, which is able to account for a large variety of systematic effects and uncertainties. For this reason, \\textsc{ARES} has the potential to contribute to the era of precision cosmology." }, "0911/0911.5566_arXiv.txt": { "abstract": "We use high-resolution cosmological hydrodynamical AMR simulations to predict the characteristics of \\lya emission from the cold gas streams that fed galaxies in massive haloes at high redshift. The \\lya luminosity in our simulations is powered by the release of gravitational energy as gas flows from the intergalactic medium into the halo potential wells. The UV background contributes only $<20\\%$ to the gas heating. The \\lya emissivity is due primarily to electron-impact excitation cooling radiation in gas $\\sim 2\\times 10^4$~K. We calculate the \\lya emissivities assuming collisional ionisation equilibrium (CIE) at all gas temperatures. The simulated streams are self-shielded against the UV background, so photoionisation and recombination contribute negligibly to the \\lya line formation. We produce theoretical maps of the \\lya surface brightnesses, assuming that $\\sim 85\\%$ of the \\lya photons are directly observable. We do not consider transfer of the \\lya radiation, nor do we include the possible effects of internal sources of photoionisation such as star-forming regions. Dust absorption is expected to obscure a small fraction of the luminosity in the streams. We find that typical haloes of mass $\\Mv\\sim10^{12-13}\\msun$ at $z\\sim 3$ emit as \\lya blobs (LABs) with luminosities $10^{43-44}\\ergs$. Most of the \\lya comes from the extended ($50-100\\kpc$) narrow, partly clumpy, inflowing, cold streams of $(1-5)\\times 10^4$K that feed the growing galaxies. The predicted LAB morphology is therefore irregular, with dense clumps and elongated extensions. The integrated area contained within surface-brightness isophotes of $2 \\times 10^{-18}\\ergsc$ is $\\sim 2-100\\, {\\rm arcsec}^2$, consistent with observations. The linewidth is expected to range from $10^2$ to more than $10^3 \\kms$ with a large variance. The typical \\lya surface brightness profile is $\\propto r^{-1.2}$ where $r$ is the distance from the halo centre. Our simulated LABs are similar in luminosity, morphology and extent to the observed LABs, with distinct kinematic features. The predicted \\lya luminosity function is consistent with observations, and the predicted areas and linewidths roughly recover the observed scaling relations. This mechanism for producing LABs appears inevitable in many high-$z$ galaxies, though it may work in parallel with other mechanisms. Some of the LABs may thus be regarded as direct detections of the cold streams that drove galaxy evolution at high $z$. ", "introduction": "\\label{sec:intro} Hundreds of Lyman-alpha blobs (LAB) have been detected so far in the redshift range $z = 2 - 6.5$, mostly near $z \\sim 3$ \\citep{steidel, matsuda, saito, ouchi, yang}. Their luminosities range from below $10^{43}$ to above $10^{44}\\ergs$, and they extend on the sky to $30-50\\kpc$ and more. The two main open questions are (a) the origin of the extended cold and relatively dense gas capable of emitting \\lya, and (b) the continuous energy source for exciting the gas to emit \\lya. The emitting hydrogen gas should be at a temperature $T \\gsim 10^4$K, relatively dense and span a much larger area than covered by the stellar component of galaxies. It may arise from outflows or inflows. The energy sources discussed in the literature include photoionisation by obscured AGNs, early starbursts or extended X-ray emission \\citep{haimanb, jimenez, scharf}, as well as compression of ambient gas by superwinds to dense \\lya emitting shells \\citep{mori}, and star formation triggered by relativistic jets from AGNs \\citep{rees}. Many of the bright LABs are found in the vicinity of massive, star-forming galaxies \\citep{matsudab}. Multi-wavelength observations reveal that a fraction of the LABs are associated with sub-millimetre and infrared sources that indicate very high star-formation rates (SFR) in the range $10^2-10^3 \\sy$ \\citep{chapman, geacha, geachb} or with obscured active galactic nuclei (AGN) \\citep{basu, scarlata}. While stellar feedback and AGNs could in principle provide the energy source for the \\lya luminosity, many LABs are not associated with sources of this sort that are powerful enough to explain the observed \\lya luminosities \\citep{kim, sj}. This indicates that star formation and AGNs are not the sole drivers of LABs, and may not even be the dominant source for LABs. Indeed, high-redshift galaxies exhibit a generic mechanism that simultaneously provides both the cold gas and the energy source. It is a direct result of the phenomenon robustly established by simulations and theoretical analysis, where high-z massive galaxies are continuously fed by narrow, cold, intense, partly clumpy, gaseous streams that penetrate through the shock-heated halo gas into the inner galaxy, grow a dense, unstable, turbulent disc with a bulge and trigger rapid star formation \\citep{bd03, keresa, db06, ocvirk, keresb, nature, peter, dsc, cd}. Massive clumpy star-forming disks observed at $z\\sim 2$ \\citep{genzel08, genel, foerster2} may have been formed via the smooth and steady accretion provided by cold flows, as opposed to merger events. The streaming of the gas into the dark-matter halo potential well is associated with transfer of gravitational energy to excitations of the hydrogen atoms followed by cooling emission of \\lya \\citep{haiman, fardal, db06, db08, ko08, furlanetto2, furlanetto}. There were two earlier attempts to compute the \\lya cooling radiation from hydrodynamic SPH cosmological simulations. Based on their analysis, \\citet{fardal} pointed out the potential association of this \\lya cooling emission with the first observed LABs. These early simulations did not allow a proper resolution of the detailed structure of the cold streams as they penetrate through the hot medium. Their shortcomings included intrinsic limitations of the SPH technique, a limited force resolution of $7 \\hkpc$ (comoving), not allowing for radiative cooling below $10^4$K and neglecting the photoionisation by the UV background. \\citet{furlanetto} used SPH simulations with a somewhat higher resolution of $\\sim$1\\,kpc to make more detailed comparisons of the simulated and observed luminosity functions and size distributions of LABs powered by cold accretion. In their pessimistic model, assuming no emissivity from gas that is self-shielded from the ionising background radiation, they end up with low \\lya luminosity. They comment that cooling IGM gas may explain the observations only if one adopts an optimistic scenario, where the self-shielded gas is emitting \\lya at CIE. The latter scenario will also be adopted in this paper. Other sources of ionisation are ignored and radiative transfer is not applied. \\citet[][hereafter DL09]{dijkstra} have recently worked out an analytic toy model for \\lya cooling radiation from cold streams, based on the general properties of the cold streams as reported from cosmological simulations. They conclude that the streams could in principle provide spatially extended \\lya sources with luminosities, line widths and abundances that are similar to those of observed LABs. They point out that the filamentary structure of cold flows may explain the wide range of observed LAB morphologies. They also highlight the fact that the most luminous cold flows are associated with massive haloes, which preferentially reside in dense large-scale surroundings, in agreement with the observed presence of bright LABs in dense environments. This model presents a successful feasibility test for the role of cold streams in powering the LAB emission, and it provides physical intuition into the way by which this process could be manifested. However, a comparison to our simulations indicates that this simplified model does not capture the detailed hydrodynamic properties of the cold streams. In particular, it significantly over-predicts the cold-gas density, under-predicts the gas temperature, and does not account for the partly clumpy nature of the streams and their characteristic radial distribution in the halo. As a result, the DL09 model underestimates the total \\lya luminosity by a factor of a few, and it therefore has to appeal to the excessive clustering of the LABs in order to boost up the predicted luminosity function for a match with the observations. In the current paper, we calculate the \\lya emission from the cold gas in high-redshift galactic haloes using state-of-the-art hydrodynamical AMR cosmological simulations. With a maximum resolution better than 70\\,pc in our simulations we can quite accurately map the extended \\lya sources. This allows us to study their individual shapes, morphologies and kinematics. We measure quantities such as the distribution of surface brightness within each halo, the area covered with surface brightness above a given isophotal threshold, the predicted observed linewidth, the typical total \\lya luminosity per halo of a given mass, and the overall \\lya luminosity function of LABs. These predicted properties are compared with the observed LABs. This paper is organised as follows. In \\se{feasibility} we work out a simple feasibility test where we use a simple toy model to estimate the expected gravitational heating power as a proxy for the \\lya luminosity. In \\se{sim} we introduce the two sets of cosmological simulations used. In \\se{lal} we explain our methodology of computing \\lya emissivity and luminosity as a function of halo mass. In \\se{lyagas} we identify the gas that contributes to the \\lya emission. In \\se{images} we apply our methodology to the simulations and provide predicted images of LABs. In \\se{L(M,z)} we determine the scaling of \\lya luminosity as a function of halo mass and redshift. In \\se{lumfun} we compare our predicted \\lya luminosity function with observational results. In \\se{obs} we measure the predicted surface-density profile, isophotal area and linewidth and compare to observations. In \\se{grav} we show that gravitational heating is the main source of energy driving the \\lya luminosity in our simulations, while the role of photoionisation is minor. In \\se{con} we discuss our analysis and results and draw our conclusions. ", "conclusions": "\\label{sec:con} Hydrodynamical cosmological simulations robustly demonstrate that massive galaxies at high redshifts were fed by cold gas streams, inflowing into dark-matter haloes at high rates along the cosmic web \\citep{keresa, db06, nature}. In this paper we have shown that these streams should be observable as luminous \\lya sources, with elongated irregular structures stretching for distances of over 100 kpc. The release of gravitational potential energy by the instreaming gas as it falls into the halo potential well is the origin of the \\lya luminosity, which ranges from $10^{42}$ to $10^{44}\\ergs$ for haloes in the mass range $10^{11}-10^{13}\\msun$ at $z \\sim 3$. The predicted \\lya emission morphologies and luminosities make such streams likely candidates for the sources of observed high redshift \\lya blobs. Most of the gas in the cold streams is at temperatures in the range $(1-5)\\times 10^4$K, for which the \\lya emissivity is maximised. The hydrogen gas densities in the streams are in the range $0.01-0.1 \\cmpc$, and are higher in clumps that flow with the streams, leading to the high surface brightnesses. Using state-of-the-art cosmological AMR simulations with 70 pc resolution and cooling to below $10^4$K, and applying a straightforward analysis of \\lya emissivity due to electron impact excitation, we produced maps of \\lya emission from simulated massive galaxies at $z \\sim 3$. We computed the average \\lya luminosity per given halo mass and the predicted luminosity function of these extended \\lya sources, with an uncertainty of $\\pm 0.4$ dex. The properties of the individual images resemble those of the observed LABs in terms of morphology and kinematics. The predicted luminosity function is close to the observed LAB luminosity function. The simulated LABs qualitatively reproduce the observed correlations between the global LAB properties of \\lya luminosity, isophotal area and linewidth. The LAB properties can be understood using a very simple toy model that accounts for gravitational heating of the inflowing gas. In comparison, the more elaborate toy model of DL09 predicts a steeper power-law relation between luminosity and halo mass (their Fig.~3, with $f_{\\rm g}=0.2$). In the mass range $10^{12} - 10^{13}\\msun$, DL09 predict luminosities in the range $1.5 \\times 10^{42} - 2.5\\times 10^{44} \\ergs$. A comparison to our \\fig{lumiz246} indicates that they underestimate the luminosity at $\\Mv \\sim 10^{12}\\msun$ by a factor of a few. The DL09 toy model matches the observed luminosity function only after the predictions are boosted, trying to account for an overdense region of the universe. The association of the predicted LABs with massive haloes places them preferentially in overdense environments, as observed \\citep{steidel, matsuda, matsudab}, but this is already accounted for by the number density of haloes in a fair sample. The association of observed LABs with star-forming LBGs \\citep{hayashino}, sub-millimetre galaxies \\citep{chapman, geacha} or active galactic nuclei \\citep{geachc, basu} is not surprising since star formation and AGN activity are triggered by the same process of streaming of cold gas into massive haloes. \\lya emission can be driven by any of these sources of energy. For example, in photoionised gas in star-forming regions the \\lya luminosity is $L_{\\rm L\\alpha} = 1.5 \\times 10^{42}\\, {\\rm SFR}\\, {\\rm erg}\\, {\\rm s}^{-1}$ where SFR is the star-formation rate (in $M_\\odot$ yr$^{-1}$) for continuous star formation for a Kroupa initial mass function \\citep{shp}. Thus the typical LAB luminosity of 10$^{43}$ erg s$^{-1}$ would require SFR $\\sim 10$ M$_\\odot$ yr$^{-1}$. However, most of this radiation is likely absorbed by dust, and confined to the central galaxy. A limitation of our AMR simulations arises from the artificial pressure floor imposed in order to properly resolve the Jeans mass. This may have an effect on the temperature and density of the cold gas in the streams, with potential implications on the computed \\lya emission. Still, the AMR code is the best available tool for recovering the stream properties. With 70 pc resolution and proper cooling below $10^4$K, the {\\CD} simulations provide the most reliable description of the cold streams so far. The rather small correction that we had to apply to the luminosities extracted from the {\\MN} simulation indicates that the \\MN galaxies can be used to recover the mass and redshift dependence of the global stream properties and the scaling relations between them. Another source of uncertainty has to do with the simplified way the ionisation by the UV background is handled in the simulations, and with the post-processing calculation of the ionisation state of the cold gas. In the \\CD simulations, the centres of the streams, where the gas density is higher than $0.1 \\cmmc$, were assumed to be self-shielded against the UV background, in agreement with our analytic estimates in \\se{lal}. The ionisation state of the gas, which is a key ingredient in evaluating the \\lya emissivity, was computed in post-processing assuming CIE. Cooling rates were computed for the given gas density, temperature, metallicity, and UV background based on \\textsc{Cloudy} (\\CD) or assuming ionisation equilibrium for H and He, including both collisional- and photo-ionisation (\\MN). Photoionisation is neglected since the streams are sufficiently thick to be self-shielded. An alternative calculation using \\textsc{Cloudy} yielded lower fractions of neutral hydrogen at $n \\leq 0.01 \\cmmc$, and total luminosities that are typically three times smaller than obtained using CIE. This calculation assumed that each cell, with a given hydrogen density $n$, is in the middle of a uniform slab of thickness 1 kpc. Based on our estimates of self-shielding, \\equ{shielding}, we adopt the CIE results as the more reliable approximation. In our computations so far we did not consider the radiative transfer of \\lya photons through the streams, or through the intervening intergalactic medium. We assume that most of the radiation emitted at $z\\sim 3$ will reach the observer at $z=0$ without undergoing significant attenuation. Resonant line scattering in the optically thick streams will tend to spread out the \\lya emission region, while the repeated Doppler shifts broaden the line profiles, and the increased path length of the random walk amplifies the absorption by dust. Such effects are expected to modify the images and line profiles, but to have only a small effect on the total \\lya luminosity. An analysis of \\lya emission from our simulated galaxies including the effects of radiative transfer and dust absorption and a more accurate treatment of photoionisation from stars will be presented in a forthcoming paper (Kasen et al, in preparation). Other hydrogen emission lines, such as L$\\beta$ or H$\\alpha$, are expected to be two orders of magnitude less luminous than \\lya in our model \\citep{baldwin, miller2}. The column densities of $N_{\\rm HI}\\sim 10^{20}\\cmms$ in the cold streams should also be detectable as \\lya absorption in quasar spectra \\citep{prochaska99, wolfe}. Considering all the galaxies that are intersected by a line of sight to a background quasar with an impact parameter $<\\Rv$, we crudely estimate from the simulations that absorption by $N_{20}>1$ hydrogen should be detected in $\\sim 20\\%$ of the galaxies. Alternatively, \\lya photons that are emitted form the central galaxy should be absorbed in the radial streams feeding the same galaxy at $N_{20}>10$, but only in $\\sim 5\\%$ of the galaxies. The streams could also be detectable as low-ionisation metal absorbers \\citep[e.g.][]{gnat} as long as the metallicity in the streams is greater than $\\sim 0.01$ solar (paper in preparation). Our results support the idea that the observed LABs are direct detections of the cold steams that drive the evolution of massive galaxies at high redshifts. Even though the observed LABs are sometimes associated with central sources that are energetic enough to power the observed \\lya emission, such as starbursts and AGNs, these central sources are very different from each other in the different galaxies, and in many LABs they are absent altogether \\citep{yang, geachb}. The gravitational heating associated with the inflowing cold streams is a natural mechanism for driving the extended \\lya cooling radiation observed as LABs, and this extended \\lya emission is inevitable in most high-redshift galaxies." }, "0911/0911.0669_arXiv.txt": { "abstract": "Future galaxy surveys will map the galaxy distribution in the redshift interval $0.5 -15.4$; a flux limit of at least $\\logfha = -16$ is required for an \\ha\\ sample to become competitive in effective volume. ", "introduction": "A number of approaches have been proposed to uncover the nature of the accelerating expansion of the Universe which involve measuring the large scale distribution of galaxies \\citep[e.g ][]{albrecht06,peacock06}. The ability of galaxy surveys to discriminate between competing models depends on their volume. Once the solid angle of a survey has been set, the useful volume can be maximised by choosing a tracer of the large-scale structure of the Universe which can effectively probe the geometrical volume. This depends on how the abundance of tracers drops with increasing redshift, and how much of this decline is offset by an increase in the clustering amplitude of the objects. Several wide-angle surveys have probed the redshift interval between $00.5$, falls into the near-infrared part of the electromagnetic spectrum \\citep{thompson96,mccarthy99,hopkins00,shim09}. \\ha\\ emission is powered by UV ionizing photons from massive young stars. The only source of attenuation is dust, which is less important at the wavelength of \\ha\\ than it is for shorter wavelength lines. This makes \\ha\\ a more direct tracer of galaxies which are actively forming stars than other lines such as \\lya, OII, OIII, H$\\beta$ or H$\\gamma$, which suffer from one or more sources of attenuation (i.e. dust, stellar absorption, resonant scattering) and which are more sensitive to the metallicity and ionisation state of the gas. The second option is to use some form of multi-slit spectrograph to carry out a redshift survey of a magnitude limited sample. The use of a slit means that unwanted background is reduced, allowing fainter galaxies to be targetted. Also, it is easier to identify which spectrum belongs to which galaxy with a slit than it is with slitless spectroscopy. Targets could be selected in the H-band at an effective wavelength of just over 1 micron, which is around the centre of the near infrared wavelength part of the spectrum. Space missions designed to carry out redshift surveys like the ones outlined above are currently being planned and assessed on both sides of the Atlantic. At the time of writing, the European Space Agency is conducting a Phase A study of a mission proposal called Euclid \\footnote[1]{http://sci.esa.int/science-e/www/object/index.cfm?fobjectid=43226}, one component of which is a galaxy redshift survey. Both of the selection techniques mentioned above are being evaluated as possible spectroscopic solutions. The slit solution for Euclid is based on a novel application of digital micromirror devices (DMDs) to both image the galaxies to build a parent catalogue in the H-band and to measure their redshifts (see Cimatti et~al. 2009 for further details about the Euclid redshift survey). A \\ha\\ mission is also being discussed in the USA \\footnote[2]{http://jdem.gsfc.nasa.gov/}. At this stage, the sensitivity of these missions is uncertain and subject to change. For this reason we consider a range of \\ha\\ flux limits and H-band magnitudes when assessing the performance of the surveys. The specifications and performance currently being discussed for these missions have motivated the range of fluxes that we consider. A simple first impression of the relative merits of different selections methods can be gained by calculating the effective volume of the resulting survey. This requires knowledge of the survey geometry and redshift coverage, along with the redshift evolution of the number density of sources and their clustering strength. In this paper we use published galaxy formation models to predict the abundance and clustering of different samples of galaxies in order to compute the effective volumes of a range of \\ha\\ and H-band surveys. Observationally, relatively little is known about the galaxy population selected by \\ha\\ emission or H-band magnitude at $0.5 < z < 2$. Empirically it is possible to estimate the number density of sources from the available luminosity function data and, on adopting a suitable model, to use the limited clustering measurements currently available to infer the evolution of the number density and bias \\citep{shioya08,morioka08,geach08}. Geach et~al. (2009), in a complementary study to this one, make an empirical estimate of the number density of \\ha\\ emitters, and combine this with the predictions of the clustering of these galaxies presented in this paper to estimate the efficiency with which \\ha\\ emitters can measure the large scale structure of the Universe. The outline of the paper is as follows: in Section~\\ref{sec.models} we give a brief overview of the models. Some general properties of \\ha\\ emitters in the models, such as luminosity functions (LF), equivalent widths (EW) and clustering bias are presented in Section~\\ref{sec.properties} as these have not been published elsewhere. In Section~\\ref{sec.DE} we show how our models can be used to build mock survey catalogues. We analyse the differences in the clustering of \\ha\\ emitters and H-band selected galaxies and present an indication of the efficiency with which different surveys trace large-scale structure (LSS). Finally, we give our conclusions in Section~\\ref{sec.conclusions}. ", "conclusions": "\\label{sec.conclusions} In this paper we have presented the first predictions for clustering measurements expected from future space-based surveys to be conducted with instrumentation sensitive in the near-infrared. We have used published galaxy formation models to predict the abundance and clustering of galaxies selected by either their \\ha\\ line emission or H-band continuum magnitude. The motivation for this exercise is to assess the relative performance of the spectroscopic solutions proposed for galaxy surveys in forthcoming space missions which have the primary aim of constraining the nature of dark energy. The physical processes behind \\ha\\ and H-band emission are quite different. \\ha\\ emission is sensitive to the instantaneous star formation rate in a galaxy, as the line emission is driven by the number of Lyman continuum photons produced by massive young stars. Emission in the observer frame H-band typically probes the rest frame $R$-band for the proposed magnitude limits and is more sensitive to the stellar mass of the galaxy than to the instantaneous star formation rate. The {\\tt GALFORM} code predicts the star formation histories of a wide population of galaxies, and so naturally predicts their star formation rates and stellar masses at the time of observation. Variation in galaxy properties is driven by the mass and formation history of the host dark matter halo. This is because the strength of a range of physical effects depend on halo properties such as the depth of the gravitational potential well or the gas cooling time. This point is most striking in our plot of the spatial distribution of \\ha\\ and H-band selected galaxies, Fig.~\\ref{fig.dm}. This figure shows remarkable differences in the way that these galaxies trace the underlying dark matter distribution. \\ha\\ emitters avoid the most massive dark matter haloes and trace out the filamentary structures surrounding them. The H-band emitters, on the other hand, are preferentially found in the most massive haloes. This difference in the spatial distribution of these tracers has important consequences for the redshift space distortion of clustering. In this paper we have studied two published galaxy formation models, those of \\citet{baugh05} and \\citet{bower06}. The models were originally tuned to reproduce a subset of observations of the local galaxy population and also enjoy notable successes at high redshift. We presented the first comparison of the model predictions for the properties of \\ha\\ emitters, extending the work of \\citet{dell1,dell2} and \\citet{orsi08} who looked at the nature of Lyman-alpha emitters in the models. Observations of \\ha\\ emitters are still in their infancy and the datasets are small. The model predicitions bracket the current observational estimates of the luminosity function of emitters and are in reasonable agreement with the distribution of equivalent widths. The next step towards making predicitions of the effectiveness of future redshift surveys is to construct mock catalogues from the galaxy formation models \\citep[see ][]{baugh08}. Using the currently available data, we used various approaches to fine tune the models to reproduce the observations as closely as possible. The main technique was to rescale the line and continuum luminosities of model galaxies; another approach was to randomly dilute or sample galaxies from the catalogue. This allowed us to better match the number of observed galaxies. The resulting mocks gave reasonable matches to the available clustering data around $z \\sim 2$. Our goal in this paper was to make faithful mock catalogues. The nature of \\ha\\ emitters in hierarchical models will be pursued in a future paper. The ability of future surveys to measure the large scale structure of the Universe can be quantified in terms of their effective volumes. The effective volume takes into account the effect of the discreteness of sources on the measurment of galaxy clustering. If the discreteness noise is comparable to the clustering signal, it becomes hard to extract any useful clustering information. Once this point is reached, although the available geometrical volume is increased by going deeper in redshift, in practice there is little point as no further statistical power is being added to the clustering measurments. The error on a power spectrum or correlation function measurement scales as the inverse square root of the effective volume. In the case of flux-limited samples, the number density of sources falls rapidly with increasing redshift beyond the median redshift. Even though the effective bias of these galaxies tends to increase with redshift, it does not do so at a rate sufficient to offset the decline in the number density. The \\galform\\ model naturally predicts the abundance and clustering strength of galaxies needed to compute the effective volume of a galaxy survey. The differences in the expected performance of \\ha\\ and H-band selected galaxies when measuring the power spectrum is related to the different nature of the galaxies selected by these two methods. \\ha\\ emitters are active star forming galaxies, which makes them have smaller bias compared to H-band selected galaxies. Their redshift distribution is also very sensitive to the details of the physics of star formation: The effect of a top-heavy IMF in bursts in the \\bau\\ model boosts the number density of bright emitters, making the redshift distrubtion of \\ha\\ emitters very flat and slowly decreasing towards high redshifts, in contrast to the predictions of the \\bow\\ model, where a sharp peak at $z\\sim 0.5$ and a rapid decrease for higher redshifts is found. H-band galaxies are less sensitive to this effect, and the redshift distributions are similar in both models. This is why the balance between the power spectrum amplitude (given by the effective bias) and the number density is translated in two different effective volumes for \\ha\\ and H-band selected galaxies. Although there are differences in detail between the model predictions, they give similar bottom lines for the effective volumes of the survey configurations of each galaxy selection. Comparing the spectroscopic solutions in Table~\\ref{table.veff}, a slit based survey down to $H_{\\rm AB}=22$ would sample 4-10 times the effective volume which could be reached by a slitless survey to $\\logfha = -15.4$, taking into account the likely redshift success rate. To match the performance of the H-band survey, an \\ha\\ survey would need to go much deeper in flux, down to $\\logfha = -16$. We have also looked at the accuracy with which \\ha\\ emitters and H-band selected galaxies will be able to measure the bulk motions of galaxies and hence the rate at which fluctuations are growing, another key test of gravity and the nature of dark energy. All of the samples we considered showed a small systematic difference between the measured growth rate and the theoretical expectation, at about the $1 \\sigma$ level. The error on the growth rate from an \\ha\\ survey with $\\logfha > -15.4$ was found to be about three times larger than that for a sample with $H_{AB}<22$." }, "0911/0911.1279_arXiv.txt": { "abstract": "\\noindent Previous work has shown that Lyman break galaxies (LBGs) display a range in structures (from single and compact to more clumpy and extended) that is different from typical local star-forming galaxies. Recently, we have introduced a sample of rare, nearby ($z<0.3$) starburst galaxies that appear to be good analogs of LBGs. These ``Lyman Break Analogs'' (LBAs) provide an excellent training set for understanding starbursts at different redshifts. We present an application of this by comparing the rest-frame UV and optical morphologies of 30 LBAs with those of galaxies at $z\\sim2-4$ in the Hubble Ultra Deep Field. We compare LBAs with star-forming s$BzK$ galaxies at $z\\sim2$, and LBGs at $z\\sim3-4$ at the same intrinsic UV luminosity ($L_{UV}\\gtrsim0.3L^*_{z=3}$). The UV/optical colors and sizes of LBAs and LBGs are very similar, while the $BzK$ galaxies are somewhat redder and larger. LBAs lie along a mass-metallicity relation that is offset from that of typical local galaxies, but similar to that seen at $z\\sim2$. There is significant overlap between the morphologies ($G$, $C$, $A$ and $M_{20}$) of the local and high redshift samples, although the high redshift samples are somewhat less concentrated and clumpier than the LBAs. Based on their highly asymmetric morphologies, we find that in the majority of LBAs the starbursts appear to be triggered by interactions/mergers. When the images of the LBAs are degraded to the same sensitivity and linear resolution as the images of LBGs and BzK galaxies, we find that these relatively faint asymmetric features are no longer detectable. This effect is particularly severe in the rest-frame ultraviolet. It has been suggested that high redshift galaxies experience intense bursts unlike anything seen in the local universe, possibly due to cold flows and instabilities. In part, this is based on the fact that the majority ($\\sim$70\\%) of LBGs do not show morphological signatures of interactions or mergers. Our results suggest that this evidence is insufficient, since a large fraction of such signatures would likely have been missed in current observations of galaxies at $z\\sim2-4$. This leaves open the possibility that clumpy accretion and mergers remain important in driving the evolution of these starbursts, together with rapid gas accretion through other means. ", "introduction": "\\label{sec:intro} One of the key tasks in galaxy evolution is to understand how the young, forming galaxies observed at high redshift relate to the well-defined Hubble sequence observed at the present epoch. The study of sizes and morphologies of large samples of galaxies as a function of redshift would not have been possible without the {\\it Hubble Space Telescope} (HST). Studies of the rest-frame UV sizes of Lyman Break Galaxies (LBGs) at $z\\sim3-7$ indicate that they are mostly very compact objects with a single core ($r_{1/2}\\sim0.7-1.5$ kpc), and that the size distribution develops a tail of larger sized objects of up to several kpc towards lower redshifts \\citep[e.g.][]{bouwens04,ferguson04,papovich05,oesch09b}. Morphological studies performed by means of (a combination of) visual classifications, quantitative morphological parameters and two-dimensional profile fitting \\citep[e.g.][]{abraham96,lotz04,lotz06, delmegreen05,ravindranath06,conselice09,petty09} have shown that this increase in sizes at $z\\sim2-4$ is related to the accumulation of luminous star-forming clumps or cores within a variety of structures, including spheroid- and disk-like objects and irregular objects. These clumpy systems are expected to coalesce and form a spheroid while a surrounding disk may grow through the continued accretion of gas \\citep{belmegreen08,belmegreen09,dekel09}. Individual clumps could in some cases be the nuclei of star-forming objects that are merging together, or they could be giant starburst regions inside a larger gaseous system induced by merging or a plentiful smooth or ``lumpy'' gas supply. Detailed knowledge on the importance of such processes would, in principle, provide powerful constraints on models of galaxy formation \\citep[e.g.][]{birnboim03,somerville01,somerville08,guo08,guo09,dekel09}, but observationally they are hard to ascertain, especially at high redshift. First, estimates of the (major) merger rate of galaxies by means of galaxy pair counts are difficult and critically depend on the merger time-scale, which may evolve with redshift \\citep{kitzbichler08}. Second, our ability to identify galaxy mergers is a strong function of, e.g., the stage of the merger, the viewing angle, and gas fraction \\citep[e.g.][]{lotz08}. Third, at high redshift it is neither possible to directly measure the (HI) gas fractions of galaxies nor to map the distribution of intergalactic hydrogen gas believed to be the supplying reservoir. However, the relatively wide range in kinematic properties of the emission line gas observed in high redshift galaxies \\citep[e.g.][]{law07,law09,lehnert09,forster09,lowenthal09} may indicate that a variety of the basic mechanisms outlined above could be at play. One of the most basic tools that can be used to study the origin of these peculiar morphologies at high redshift is to contrast them against local or lower redshift samples of galaxies \\citep[e.g.][]{hibbard97,burgarella06,lotz06,scarlata07,cardamone09,delmegreen09,ostlin09,petty09,rawat09}. However, it is important to keep in mind that the properties of galaxies in the local universe are generally very different from those at high redshift, making it hard to disentangle actual physical differences from observational biases. To better facilitate the straight comparison, the ``Lyman break analogs'' (LBA) project was designed in order to search for local starburst galaxies that share typical characteristics of high redshift LBGs \\citep{heckman05}. In brief, the UV imaging survey performed by the Galaxy Evolution Explorer (GALEX) was used in order to select the most luminous ($L_{FUV}>10^{10.3}$ $L_\\odot$) and most compact ($I_{FUV}>10^{9}$ $L_\\odot$ kpc$^{-2}$) star-forming galaxies at $z<0.3$. Although such galaxies are very rare, they tend to be much more luminous in the UV than typical local starburst galaxies studied previously \\citep{heckman98,meurer97,meurer99} consistent with high SFRs and relatively little dust extinction. The median absolute UV magnitude of the sample is --20.3, corresponding to $\\simeq0.5L_{z=3}^*$, where $L_{z=3}^*$ is the characteristic luminosity of LBGs at $z\\sim3$ \\citep[][i.e., $M_{1700,AB}=-21.07$]{steidel99}. Analysis of their spectra from the Sloan Digital Sky Survey (SDSS) and their spectral energy distributions from SDSS, GALEX, Spitzer and VLA and follow-up spectroscopy subsequently showed that the LBAs are similar to LBGs in their basic global properties, including: stellar mass, metallicity, dust extinction, SFR, and emission line gas properties \\citep{heckman05,hoopes07,basu-zych07,overzier09}. One of the advantages of LBAs being so bright in the rest-frame UV is that their morphologies can be easily compared with typical galaxies at high redshift without having to artifically brighten them as was done in previous studies \\citep{lotz04,petty09}. In \\citet{overzier08} (Paper I) we analyzed the UV morphologies of a small sample of 8 LBAs observed with HST, finding that most of the UV emission in the LBAs originates in highly compact burst regions in small, clumpy galaxies that appear morphologically similar to LBGs. We argued that if LBGs at high redshift are also small merging galaxies similar to the LBAs, this would be very hard to detect given the much poorer physical resolution and sensitivity. In this paper, we present an analysis of the morphologies of our full data set consisting of rest-frame UV and optical HST images of 30 LBAs and compare with star-forming galaxies at $z\\simeq2,3,4$ in the Hubble Ultra Deep Field (HUDF). The structure of this paper is as follows. In Section 2 we describe our low and high redshift samples, the observations and data reduction methods, and our techniques for producing redshifted simulated images as well as for performing parametrized galaxy morphologies. In Section 3 we compare the rest-frame UV and optical colors, sizes and morphologies of the LBAs, BzKs and LBGs. We discuss our results in Section 4, followed by a summary of the main results. We use the AB magnitude system throughout the paper, and assume a cosmology [$\\Omega_M$,$\\Omega_\\Lambda$,$H_0$] $=$ [0.27,0.73,73.0] (with $H_0$ in km s$^{-1}$ Mpc$^{-1}$) so that the angular scales at $z\\approx0.2$ and 3.0 are about 3 and 8 kpc arcsec$^{-1}$, respectively. ", "conclusions": "\\label{sec:disc} \\subsection{Morphologies of LBAs at different redshifts} In previous works we have shown that LBAs are clumpy starburst galaxies with peculiar morphologies that are most consistent with merging \\citep{overzier08,overzier09}. Although we currently do not know the relative contribution from major and minor mergers, the diversity in LBA morphologies at least suggests a range in merger conditions and stages albeit that they all have in common that their starbursts are all very young (ages of a few tens of Myr), compact and UV-luminous \\citep{overzier09}. We have shown that, both in the UV and optical, asymmetry appears to be a better indicator for this merger activity than $M_{20}$. The median $A$ is larger than the merger criterion of $A>0.35$ from \\citet{conselice09}, while the median $M_{20}$ is much lower than the merger criterion of $M_{20}\\gtrsim-1.1$ \\citep[e.g.][]{lotz04,conselice09}. We redshifted the LBAs to $z=2-4$ and found a significant drop in the UV asymmetries such that the median $A$ fell well below the merger criterion. In the optical, the drop was smaller and less significant. $M_{20}$ became even smaller as the different star-forming clumps blended into single light concentrations at high redshift. We thus conclude that if galaxies similar to LBAs were observed in the UV at high redshift, all but a few would be mistaken as being relatively smooth and symmetric rather than identified as mergers. In the optical, we estimate that $\\sim$50\\% would be identified as symmetric. \\citet{lotz08} investigated in detail the complicated relation between the ability to morphologically classify merging galaxies on one hand, and the physical details of the merger on the other. As expected, the merger observability time-scale (i.e., the time during which a merger is identified as such compared to the total merging time) critically depends on, e.g., orbital parameters, gas fraction, dust, SN feedback, and viewing angle. These results explain, for example, why the $G,C,A,M_{20}$ system is much more effective in identifying the roughly two-thirds of local ULIRGs having more than one nuclei that are seen either just prior to final coalescence or in projection at first passage, than single nucleus ULIRGs \\citep{lotz04,lotz08}. In a similar fashion, this ``merger observability'' as defined by \\citet{lotz08} can also explain why a significant fraction of the LBAs do not appear to be particularly disturbed for some or all of the morphological parameters. However, in addition to these effects, our simple simulations clearly show that for LBA-like galaxies observed at high redshift the decreased physical resolution and sensitivity are of equal or even greater importance than the physical details of the merger. \\subsection{Morphologies of LBAs, BzKs and LBGs} The median $M_{20}$ of LBGs at $z\\sim3-4$ is larger than that of LBAs, while the $A$ is similar. The fractions of LBGs classified as disturbed based on either one of these parameters in the UV is $\\sim$30\\%, consistent with previous works \\citep{lotz04,lotz06,conselice09}. In the optical, a similar result is obtained based on $A$, while a much smaller fraction is derived based on $M_{20}$. It is interesting to note that the differences observed between $BzK$ galaxies and LBAs are much more significant than those between LBGs and LBAs: in the UV, $BzK$s have the highest $A$ and $M_{20}$ and the lowest $C$ and $G$ of all samples considered here. In the optical, $BzK$s had the largest median half-light radii. It must be noted, however, that our way of selecting LBAs based on a UV surface brightness criterion was tuned to select objects having similar UV luminosities and sizes as typical LBGs at $z\\sim3$ \\citep{heckman05}. In principle, the LBA selection could be expanded or modified to search for such objects more similar to $BzK$ galaxies, but one must be more careful as such a selection would include a wider variety of galaxies that are not so suitable analogs of high redshift objects compared to LBAs \\citep{hoopes07}. An example of a particularly clumpy LBA is the luminous blue compact dwarf galaxy Haro 11 at $z=0.02$ \\citep{ostlin09}, which lies close to the edge of our LBA selection criteria based on FUV luminosity and surface brightness \\citep{grimes07}. It is similar to the LBAs in most aspects but with a relatively high degree of clumpiness due to three strong light concentrations in an otherwise amorphous, forming galaxy. Its UV morphology at high redshift is similar to that of double nucleated galaxies identified on the basis of a large $M_{20}$ or even a ``by eye'' classification \\citep{overzier08}. \\subsection{Possible Implications for accretion processes in starbursts at high redshift} Based on the results presented in the previous sections, we can now make a series of statements purely based on the comparison of the morphologies of LBAs and high redshift starbursts (such as BzKs and LBGs):\\\\ 1. LBAs show clumpy star formation. Our redshift simulations suggest that real high redshift samples have a similar (LBGs; $z\\sim3-4$) or perhaps a slightly stronger ($BzK$s; $z\\sim2$) degree of clumpy star formation.\\\\ 2. LBAs show luminous UV clumps and faint optical tidal features that, in the majority of cases, are interpreted by us as being due to mergers based on the visual evidence. Our simulations suggest that this information can not be recovered from the rest-frame UV morphological parameters at high redshift, while in the rest-frame optical perhaps 1 out of every 2 objects would be classified as disturbed. This clearly demonstrates that current estimates for the fraction of galaxies at high redshift having disturbed morphologies (as opposed to smooth/symmetric morphologies) is most likely a (weak) lower limit on the true fraction. This statement is independent of the physical mechanism that is the cause of these disturbed morphologies.\\\\ 3. The observations suggest a scenario where either cold gas accretion and internal instabilities can drive clumpy star formation in LBGs at levels higher than any seen in LBAs in the local universe, or where clumpy cold accretion, possibly at the level of major or minor mergers may be responsible for star formation in LBAs and LBGs alike. We will discuss these possibilities in more detail below.\\\\ The latest observational evidence suggests that the fraction of LBGs at $z\\sim3-4$ having disturbed or distorted morphologies is $\\sim$30\\% \\citep[e.g.,][This Paper]{conselice03,lotz06,conselice09}. Although these disturbances do not necessarily need to be explained by galaxy mergers, it does suggest that bulk material is coming in, perhaps in the form of giant gas clouds or minor mergers. \\citet{conselice09} find that the fraction of $z\\sim4$ $B$-dropouts forming a pair with another $B$-dropout is very similar to the fraction of disturbed morphologies: $\\sim$20\\% within a 20 h$^{-1}$ kpc projected radius. Interpreting the similarity in the pair counts and morphologies as evidence for merging, they estimate that a $\\sim10^{10}$ $M_\\odot$ galaxy will undergo a major merger every 1--2 Gyr at $z\\sim3-4$. Similar high merger rates were obtained for $z\\sim2-3$, possibly with an increase towards the most massive and most luminous galaxies \\citep{conselice03,conselice09,bluck09}. With such high inferred merger rates, the mass growth of LBGs due to these mergers could make a significant contribution to the mass growth at $z\\sim3$ compared to that due to star formation fueled by a more continuous gas accretion and sustained over a period of a Gyr at a rate of $\\sim10$ $M_\\odot$ yr$^{-1}$. Perhaps these two channels of formation are not contradictory given that the high density environments of LBGs are likely to be simultaneously associated with both frequent mergers of small galaxies and rapid cold accretion (likely with some level of lumpiness). \\citet{conselice09} suggest that the remaining 70--80\\% of LBGs, that appear as relatively smooth/symmetric and are forming stars at a similarly high rate as the disturbed objects, could be the result from rapid gas collapse, provided that a sufficiently long amount of time ($\\gtrsim$0.5 Gyr) has passed since their last major merger otherwise this would have been apparent in their morphology given the high (inferred) merger rates. However, our new results based on LBAs present an important caveat to this interpretation: we have shown that, at least for galaxies at high redshift similar to LBAs, in the majority of cases we are not able to detect significant disturbances or asymmetries to their morphologies (mostly because of redshift effects, and not because they are not there). This allows for the possibility that the fraction of galaxies at high redshift that is undergoing interactions could be much higher than currently inferred from the observations. Alternatively, less violent accretion processes coupled with large disk instabilities at high redshift are capable of triggering starbursts at levels only seen in low redshift, merger-induced samples such as the LBAs. Future deep, high resolution observations of rest-frame optical morphologies and kinematics (see below), together with constraints on the LBG pair fractions and small-scale clustering \\citep{conselice09,cooke09} will perhaps allow us to distinguish between these scenarios. \\subsection{Relation to Studies of Gas kinematics} In recent years, the study of high redshift LBGs has advanced considerably beyond studies that are based purely on morphology. Mainly through the use of integral field spectrographs in the near-infrared it has become feasible to determine the basic kinematical properties of the emission line gas as well. Motivated by the high degree of similarity in the properties of our locally selected UV-luminous galaxies and those at high redshift, it is a useful exercise to compare the gas kinematics of LBAs and LBGs. In a first study published by \\citet{basu-zych09}, the bright Pa-$\\alpha$ emission line was used to study the resolved gas dynamics in 3 of our LBAs. In two cases, a mild velocity gradient was found, but in all three cases the kinematics were dominated by the dispersion rather than structured rotation ($v/\\sigma\\sim1-2$). Simulating the data at $z\\sim2$ demonstrated that the (gas) kinematical properties of these objects are similar to the kinematical profiles commonly seen at high redshift \\citep[e.g.][]{law07,law09,genzel08,forster09,lowenthal09}. \\citet{lehnert09} argued that neither the self-gravity of disks fueled by gas accretion flows nor the internal velocity dispersions of massive star-forming clumps can fully explain the high velocity dispersions observed at high redshift. Furthermore, the most massive of clumps observed possibly would not have been formed at all if the gas turbulence in the disk was not high enough to begin with \\citep{belmegreen09}, while mergers alone may not be sufficient to explain those objects dominated by a large number of massive clumps in so-called ``clump-cluster'' or ``chain galaxy'' configurations at high redshift \\citep{bournaud09}. Instead, \\citet{lehnert09} suggest that the kinematic properties of the ISM in starburst galaxies is affected by the mechanical energy input resulting from massive star formation. In this scenario, the starbursts that are associated with each of the clumps drive blast waves from supernovae (SN) and stellar winds that appear sufficient to give rise to the high gas pressures and high velocity dispersions observed. As shown in \\citet{overzier09}, some LBAs show very high pressures and clear evidence for an ISM that is dominated by starburst- and SN-feedback associated with massive star-forming clumps, consistent with such a scenario. If this is correct, then estimates for the merger rates at high redshift estimated from pair counts or morphologies are perhaps less biased than those obtained from the (ionized) gas kinematics because the latter may not always trace galaxy interactions, if present, very well. In addition, \\citet{robertson08} have shown with simulations that at least some merging systems would still be classified as ``disks'' using the methodology applied to observations at $z\\sim2$ by \\citet{shapiro08}. The simulations of \\citet{robertson08} also show how some of the main structural, spectral and chemical properties of certain $BzK$ galaxies can be explained in the context of gas-rich mergers as well, eventhough it has been claimed that such systems cannot be merger remnants. Similar studies of the gas kinematics in LBAs as initiated by \\citet{basu-zych09} will be very useful, and will allow us to carefully test the methods typically applied to high redshift starbursts in a suitable low redshift comparison sample. Results on the kinematical properties in a much larger sample of LBAs are forthcoming \\citep[][in prep.]{goncalves10}." }, "0911/0911.4176_arXiv.txt": { "abstract": "Ultra High Energy Cosmic Rays, UHECR, maybe protons, as most still believe and claim, or nuclei; in particular lightest nuclei as we advocated recently. The two model offer a dramatic different view of UHECR sky because different galactic Lorentz deflection and GZK cut-off. The first (Auger Collaboration) nucleon proposal (2007)\\cite{Auger-Nov07} foresaw to trace clearly the UHECR GZK Universe reaching far (up to $100$ Mpc) Super-Galactic-Plane, with little angular dispersion. On the contrary Lightest Nuclei model (2008)\\cite{Fargion2008}, inspired by observed composition and by nearest $Cen_A$ clustering (almost a quarter of the AUGER events) explains (by cut off) a modest and narrow (few Mpc) Universe view, as well as the puzzling Virgo absence. Lightest nuclei offer a little blurred Astronomy. Here we address to a part of the remaining scattered events in the new up-dated Auger map (March 2009-ICRC09). We found within rarest clustering the surprising imprint of a few remarkable galactic sources. In particular we recognize a first trace of Vela, brightest gamma and radio galactic source, and smeared sources along Galactic Plane and Center. We expect in a near future much more clustering along $Cen_A$ and a few more confirm to those galactic sources. The clustering may imply additional tails of fragments (by nuclei photo-dissociation) at half energies ($2-4 \\cdot 10^{19}$eV). The UHECR light-nuclei fragility and opacity may also reflect into a train of secondaries as gamma and neutrinos UHE events at tens-hundred PeVs. These UHE neutrinos might be detectable in a coming future within nearest AUGER and Array Fluorescence Telescope views,(few $km$ distances) by fast fluorescence flashing of horizontal up-going $\\tau$ Air-showers. ", "introduction": " ", "conclusions": "" }, "0911/0911.3426_arXiv.txt": { "abstract": "Debris discs -- analogous to the Asteroid and Kuiper-Edgeworth belts in the Solar system -- have so far mostly been identified and studied in thermal emission shortward of 100$\\,\\mu{\\rm m}$. The {\\it Herschel} space observatory and the SCUBA-2 camera on the James Clerk Maxwell Telescope will allow efficient photometric surveying at 70 to 850$\\,\\mu{\\rm m}$, which allow for the detection of cooler discs not yet discovered, and the measurement of disc masses and temperatures when combined with shorter wavelength photometry. The SCUBA-2 Unbiased Nearby Stars (SUNS) survey and the DEBRIS {\\it Herschel} Open Time Key Project are complimentary legacy surveys observing samples of $\\sim\\!500$ nearby stellar systems. To maximise the legacy value of these surveys, great care has gone into the target selection process. This paper describes the target selection process and presents the target lists of these two surveys. ", "introduction": "The solar neighbourhood is an ideal testing ground for the study of debris discs and planetary systems. Proximity maximises dust mass sensitivity and can allow systems to be spatially resolved. Systems near the Sun span a wide range of stellar parameters e.g., mass, age, metallicity, multiplicity. Whilst determining these parameters may not be easy, the diversity included in volume limited samples makes them ideal for legacy surveys where one may wish to investigate trends as a function of many system parameters. This paper presents five all-sky volume limited samples of nearby stellar systems with main-sequence primaries of spectral type A,F,G,K,M. These form the basis of the target lists of two complimentary surveys for debris discs using the SCUBA-2 \\citep[Submillimetre Common User Bolometer Array 2;][]{hol03,aud04} camera on the James Clerk Maxwell Telescope (JCMT), and the {\\it Herschel} space observatory \\citep[][]{pil08}. The \\SUNSs\\ \\citep[SUNS;][]{mat07} is a large flux-limited survey of 500 systems at $850\\,\\mu{\\rm m}$. The target flux RMS is $0.7\\,{\\rm mJy}/{\\rm beam}$, equal to the extragalactic confusion limit of the JCMT at $850\\,\\mu{\\rm m}$. Shallow $450\\,\\mu{\\rm m}$ images of varying depth will be obtained simultaneously, and deep images at $450\\,\\mu{\\rm m}$ will be proposed to follow-up $850\\,\\mu{\\rm m}$ detections. The \\DEBRIS\\ (DEBRIS) {\\it Herschel} Open Time Key Program will image 446 systems (356 in common with SUNS) at 110 and $170\\,\\mu{\\rm m}$ using the PACS \\citep[Photodetector Array Camera and Spectrometer;][]{pog08} instrument, with follow-up of around 100 systems at 250, 350 and $500\\,\\mu{\\rm m}$ using the SPIRE \\citep[Spectral and Photometric Imaging Receiver;][]{gri08} instrument. This survey is primarily driven by the $110\\,\\mu{\\rm m}$ band, which has the highest dust mass sensitivity for cold discs such as the Kuiper-Edgeworth belt of our Solar System. The intended flux RMS at $110\\,\\mu{\\rm m}$ is $1.2\\,{\\rm mJy}/{\\rm beam}$, which is twice the predicted extragalactic confusion limit. $170\\,\\mu{\\rm m}$ images are taken simultaneously with a predicted RMS of $1.7\\,{\\rm mJy}/{\\rm beam}$, equal to the predicted extragalactic confusion limit in this band. The primary goals of these surveys are statistical: In general how do debris disc properties vary with stellar mass, age, metallicity, system morphology (multiplicity, component masses, separations), presence of planets etc. To be able to answer so many questions, and to minimise the risk of unforeseen selection effects, large samples and simple, clearly defined target selection criteria are required. Volume limited samples satisfy these requirements, and as well as maximising the proximity of the targets, the stars nearest the Sun are very widely studied. For example, nearby stars are the main targets of radial velocity, astrometry and direct imaging planet searches. The majority of SUNS and DEBRIS targets also have photometry at 24 and $70\\,\\mu{\\rm m}$ from the MIPS (Multiband Imaging Photometer and Spectrometer) instrument on the {\\it Spitzer} space telescope, which ceased operation at the end of March 2009. This large spectral coverage, from 24 to $850\\,\\mu{\\rm m}$ for over 300 systems will be an incredible resource for detailed spectral energy distribution modelling of systems with debris discs. Given that we are considering the closest systems to the Sun, substantial effort was required to compile the samples presented here. Late M-type stars within $10\\,{\\rm pc}$ are still being discovered \\citep[e.g.][]{hen06}, and complete homogeneous datasets covering the spectral type and distance ranges we consider do not exist. We have tried to make our sample selection using the most complete and accurate data available at the time of the DEBRIS proposal submission in October 2007. ", "conclusions": "" }, "0911/0911.2420_arXiv.txt": { "abstract": "{} {A variation of the fundamental constants is expected to affect the thermonuclear rates important for stellar nucleosynthesis. In particular, because of the very small resonant energies of $^8$Be and $^{12}$C, the triple $\\alpha$ process is extremely sensitive to any such variations. } {Using a microscopic model for these nuclei, we derive the sensitivity of the Hoyle state to the nucleon-nucleon potential allowing for a change in the magnitude of the nuclear interaction. We follow the evolution of 15 and 60 $M_{\\sun}$ zero metallicity stellar models, up to the end of core helium burning. These stars are assumed to be representative of the first, Population III stars. } {We derive limits on the variation of the magnitude of the nuclear interaction and model dependent limits on the variation of the fine structure constant based on the calculated oxygen and carbon abundances resulting from helium burning. The requirement that some $^{12}$C and $^{16}$O be present are the end of the helium burning phase allows for permille limits on the change of the nuclear interaction and limits of order $10^{-5}$ on the fine structure constant relevant at a cosmological redshift of $z \\sim 15-20$. } {} ", "introduction": "The equivalence principle is a cornerstone of metric theories of gravitation and in particular of general relativity \\citep{willbook}. This principle, including the universality of free fall, the local position and Lorentz invariances, postulates that the local laws of physics, and in particular the values of the dimensionless constants such as the fine structure constant $\\alpha_{em}\\equiv e^2/4\\pi\\varepsilon_0\\hbar c$, must remain fixed, and thus be the same at any time and in any place. It follows that by testing the constancy of fundamental constants one actually performs a test of General Relativity, that can be extended on astrophysical and cosmological scales~\\citep[for a review, see][]{uzanrevue1,Uzan2009a} We define a fundamental constant as any free parameter of the fundamental theories at hand~\\citep{weinberg83,duff,tri,barrowbook,jpbook}. These parameters are contingent quantities that can only be measured and are assumed constant since (i) in the theoretical framework in which they appear, there is no equation of motion for them and they cannot be deduced from other constants and (ii) if the theories in which they appear have been validated experimentally, it means that, at the precision of the experiments, these parameters have indeed been checked to be constant. Hence, by testing for their constancy we extend our knowledge of the domain of validity of the theories in which they appear. In that respect, astrophysics and cosmology allow one to probe larger time-scales, typically of the order of the age of the universe. One can, however, question the constancy of these dimensionless numbers and the physics which determines their value. This sends us back to the phenomenological argument by~\\citet{dirac37}, known as the `Large Number Hypothesis', according to which the dimensionless ratio $Gm_{\\rm e}m_{\\rm p}/\\hbar c$, or simply $G$ in atomic units, should decrease as the inverse of the age of the universe, followed by~\\citet{jordan37} who formulated a field theory in which both the fine structure constant and the gravitational constant were replaced by dynamical fields. It was soon pointed out by~\\citet{fierz56} that astronomical observations can set strong constraints on the variations of these constants. This paved the way to two complementary directions of research on the fundamental constants. On the one hand, from a theoretical perspective, many theories involving ``varying constants'' have been designed. This is in particular the case of theories involving extra-dimensions, such as the Kaluza-Klein mechanism~\\citep{kaluza21,Klein26} and string theory, in which all the constants (including gauge, Yukawa and gravitational couplings) are dynamical quantities \\citep{Wu:1986ac,Wetterich:1987fk, taylor, witten}, or in theories such as scalar-tensor theories of gravity~\\citep{Jordan49,bd61,gefdam} and in many models of quintessence~\\citep{uzan99,dam02, rundil,wetterich,Lee:2004vm, riaz}, aiming at explaining the acceleration of the universe by the dynamics of a scalar field. It is impingent on these models to explain why the constants are so constant today and provide a mechanism for fixing their value~\\citep{dn,dp}. In this respect, testing for the constancy of the fundamental constants is one of the few windows on these theories. On the other hand, from an experimental and observational perspective, the variation of various constants have been severely constrained. This is the case of the fine structure constant for which the constraint $\\dot\\alpha_{em}/\\alpha_{em}=(-1.6\\pm2.3)\\times10^{-17}\\,{\\rm yr}^{-1}$ at $z=0$ has been obtained from the comparison of aluminium and mercury single-ion optical clocks~\\citep{rosenband}. Over a longer timescale, it was demonstrated that $\\alpha_{em}$ cannot have varied by more than $10^{-7}$ over the last 2 Gyr from the Oklo phenomenon~\\citep{oklo0,oklo1,Fujii:1998kn,oklo2,oklo3} and over the last 4.5 Gyr from meteorite dating~\\citep{meteo1,dyson,meteo2,meteo3}. At higher redshift, $0.4 < z < 3.5$, there are conflicting reports of a an observed variation of $\\alpha_{em}$ from quasar absorption systems. Using the many-multiplet method, \\citet{webb01} and \\cite{murph03,murphy07} claim a statistically significant variation $\\Delta\\alpha_{em}/\\alpha_{em} = (-0.54 \\pm 0.12)\\times 10^{-5}$, indicating a smaller value of $\\alpha_{em}$ in the past. More recent observations taken at VLT/UVES using the many multiplet method have not been able to duplicate the previous result \\citep{chand,sri04,quast04,srianand2007}. The use of Fe lines in \\citet{quast04} on a single absorber found $\\Delta \\alpha_{em} / \\alpha_{em} = (-0.05 \\pm 0.17) \\times 10^{-5}$. However, since the previous result relied on a statistical average of over 100 absorbers, it is not clear that these two results are in contradiction. In \\citet{chand}, the use of Mg and Fe lines in a set of 23 systems yielded the result $\\Delta \\alpha_{em} / \\alpha_{em} = (0.01 \\pm 0.15) \\times 10^{-5}$ and therefore represents a more significant disagreement and can be used to set very stringent limits on the possible variation in $\\alpha_{em}$. A purely astrophysical explanation for these results is also possible \\citep{amo,amo2}. At larger redshifts, constraints at the percent level have been obtained from the observation of the temperature anisotropies of cosmic microwave background at ($z\\sim10^{3}$) \\citep[\\textsl{e.g. }][]{cmb4,cmb3,cmb1,cmb2} and from big bang nucleosynthesis (BBN) ($z\\sim10^{10}$) \\citep[\\textsl{e.g. }][]{kpw,campolive95,Bergstrom,Nollett,Ichikawa02,bbn2,bbn3,Ichikawa04,bbn4,coc07,bbn1}. We refer to \\citet{uzanrevue1,uzanrevue2,uzanrevue3} for recent reviews on this topic. For the time being, there is no constraint on $\\alpha_{em}$ for redshifts ranging from 4 to $10^3$ although it has been proposed that 21~cm observations may allow one to fill in the range $305$ sources. We test the robustness of our results against different dust attenuations and, most importantly, against the inclusion of TP-AGB stars in Simple Stellar Populations (SSPs) used to generate galaxy spectra, and find that the inclusion of TP-AGBs has a relevant effect, in that it allows to increase by a large factor the number of very red active objects at all color cuts. We find that though the most passive and the most obscured active galaxies have a higher probability of being selected as EROs, many EROs have intermediate properties and the population does not show bimodality in specific star formation rate (SSFR). We predict that deep observations in the Far-IR, from 100 to 500 $\\mu$m, are the most efficient way to constrain the SSFR of these objects; we give predictions for future {\\it Herschel} observations, and show that a few objects will be detected in deep fields at best. Finally, we test whether a simple evolutionary sequence for the formation of $z=0$ massive galaxies, going through a sub-mm-bright phase and then a ERO phase, are typical in this galaxy formation model. We find that this sequence holds for $\\sim25$ per cent of $z = 0$ massive galaxies, while the model typically shows a more complex connection between sub-mm, ERO and massive galaxies. ", "introduction": "The wealth of data coming from recent multiwavelength surveys has made it possible to study in detail distant galaxy populations from their UV-optical-Infrared broad-band spectral energy distributions (SED). Among the various categories that have been observationally defined by means of color selection techniques, the Extremely Red Objects (EROs) have attracted wide interest. Defined on the basis of of their very red $R-K$ color (in the Vega system $R-K>5$ at least), they were first addressed in the context of deep near-infrared surveys \\citep{Elston88, McCarthy92}. Such red colors are expected to arise as a combination of an intrinsically red SED and a large K-correction at $12$ as their formation epoch. The correct prediction of such an early and efficient formation of massive galaxies at high redshift is a well known challenge for models of galaxy formation based on the CDM cosmology: in fact early implementations predicted lower space densities of EROs than observed \\citep[e.g.,][]{Firth02,Smith02}. On the other hand, it soon became evident that the same red $R-K$ colors might be linked to a completely different class of objects such as the active (starburst) galaxies embedded in dust shells that cause high extinction of the SEDs \\citep{Cimatti02a,PozzettiMannucci00,Smail02}. Interestingly enough, the relative contribution to the total space density by the two sub-populations is similar, with a passive fraction of $48$ per cent at $K<19.2$ (\\citealt{Kong09}, but a ratio between 30 and 70 per cent has been reported also by \\citealt{Cimatti02a} and \\citealt{Smail02}), and $58$ per cent at $K<20.3$ \\citep{Miyazaki03}. However, splitting the observed ERO population in two is a difficult task. The cleanest way relies on high signal-to-noise spectroscopy, able to reveal either emission lines linked to ongoing star formation or the spectral breaks typical of aged stellar populations (see e.g., \\citealt{Cimatti02a}). Alternatively, the presence of a starburst may be revealed by additional imaging in the far-infrared, where the absorbed light is reprocessed \\citep{Kong09}, or indirectly through the morphology of these objects \\citep{StiavelliTreu01,Cimatti03}. Presently, all these techniques can be successfully applied only to a small fraction of the overall population, typically limited to the brightest sources. As mentioned above, the redshift distribution and the expected physical properties of the two ERO sub-populations have been a long-standing problem for theoretical models of galaxy formation \\citep{Cimatti02a,Somerville04b}. More recent models, based on improved treatment of baryonic physics \\citep{DeLucia06,Bower06,Menci06,Monaco07,Lagos08}, are able to give a better representation of the formation and assembly of massive galaxies at $z<3$. Only a few of these models have been directly tested by comparing the predicted properties of EROs with data: in most cases a reasonable success in reproducing the abundance of passive EROs has been claimed, giving support to the interpretation of these objects as passive massive galaxies at $z\\sim1-2$ that are the progenitors of local elliptical galaxy. At the same time, the same models do not produce a sufficient number of active EROs. More specifically, \\citet{Nagamine05} implemented sub-grid baryonic physics in both a Lagrangian and Eulerian code; they found the correct abundance of massive galaxies at $z\\sim 2$, but could find almost no passive ERO, and in order to produce enough red galaxies they were forced to assume a dust attenuation as high as an equivalent $E(B-V)=0.4$ for the whole mock catalogue. \\citet{Kang06} correctly reproduced the space density of massive galaxies at $z\\sim 2$, but underpredicted the number of very red galaxies, in particular of the active class. \\citet{Menci06} proposed a more successful model able to reproduce the color distribution of galaxies at $z<1$ and the space density of massive galaxies at $17$) was used, but again the predicted ERO population was dominated by passive objects, with only a small fraction of active galaxies. On the other hand the model by \\citet{Baugh06}, which gives a good fit of the sub-mm galaxy population, was found to significantly underestimate ERO number counts. A remarkable attempt to put quasars, bright sub-mm galaxies, EROs, and low-$z$ ellipticals into a unified scenario was proposed by \\citet{Granato04}. In their model a typical low-z elliptical forms the bulk of its mass in an intense obscured starburst phase at $z\\sim2$, when it is visible in the sub-mm band. In the galaxy nucleus, star formation triggers accretion onto a seed black hole; the end of the star formation phase is caused by feedback from the resulting quasar, able to wipe away the interstellar medium from the galaxy and quench star formation. After $\\sim1$ Gyr the galaxy gets very red and becomes an ERO. It then evolves passively into a typical local elliptical galaxy. The above mentioned papers have clarified the role of the ERO population as a powerful test for theoretical models of galaxy formation. These should reproduce both their global number and the diversity of their physical properties. A successful model would be helpful in investigating the relation of EROs with other observationally selected galaxy populations, in particular with the sub-mm galaxies. Apparently, the most challenging requirement for published model is to produce a high fraction of active EROs, especially when deep samples are concerned. In principle, active EROs should be easy to produce in a hierarchical model of galaxy formation, because the redshift range where we observe them corresponds to the peak of the cosmic star formation rate density. \\citet{Kong06} and \\citet{Nagamine05} interpret the tension between their models and data as a sign of over-simplified recipes in treating baryonic physics or dust attenuations. Regarding the latter point, \\citet{Fontanot09a} showed that currently used recipes for dust attenuation, calibrated using low-$z$ samples, provide very scattered extinction values at $z\\sim 2$ and, more interestingly, do not match the predictions of the {\\gs} code \\citep{Silva98} that explicitly solves the equations of radiative transfer. Another important point is the contribution of the TP-AGB stars to the synthetic SEDs of model galaxies (see e.g. \\citealt{Maraston05}), which is not included in the above mentioned models. It has been shown by \\citet{Tonini08, Tonini09} that the inclusion of this extreme phase of stellar evolution has a strong impact on the expected photometry and colors of simulated galaxies at $1 \\la z \\la 4$, where intermediate-age stellar populations dominate the restframe $K$-band starlight. In the case of EROs, an increase of their number density is expected when TP-AGBs are included due to the reddening of rest-frame $V-K$. All these elements critically affect the modeling of galactic SEDs: they are therefore a primary concern for the interpretation of the physical ingredients of models. In this paper we use the {\\mor} model (\\citealt{Monaco07}, as updated in \\citealt{LoFaro09}), interfaced with the {\\gs} spectro-photometric + radiative transfer code \\citet{Silva98} and to Infrared (IR) template library \\citep{Rieke09}, to produce mock samples of EROs. We both use the {\\it Padova} \\citep{Bertelli94} and the \\citet[][M05 hereafter]{Maraston05} library of SSPs. We find reasonable agreement with data in terms of number counts, redshift distribution and fraction of active objects, especially when M05 library is used (details on the effect of this modification on the whole galaxy population will be the subject of a forthcoming paper). This prompts us to use the model to investigate the physical properties of these object. We address whether there is a true active/passive bimodality of the ERO population and what is the connection between EROs, sub-mm galaxies and massive galaxies at $z=0$. We finally give predictions of ERO properties for future surveys with the {\\it Herschel} satellite. This paper is organized as follows. In sec.~\\ref{sec:model} we recall the main feature of the theoretical models {\\mor} and {\\gs}. In sec.~\\ref{sec:results} we present our results: in sec.~\\ref{sec:ero_prop} we analyze the properties of the model ERO sample and in sec.~\\ref{sec:sequence} we discuss the connection between the ERO population, the sub-mm bright galaxies and the massive galaxies at $z=0$. Finally in sec.~\\ref{sec:final} we give our conclusions. Throughout the paper we assume magnitudes are in the Vega system (unless otherwise stated), and a cosmological model consistent with WMAP3 results ($\\Omega_0=0.24$, $\\Omega_\\Lambda=0.76$, $h=0.72$, $\\sigma_8=0.8$, $n_{\\rm sp}=0.96$, \\citealt{Komatsu09}). ", "conclusions": "\\label{sec:final} \\begin{figure} \\centerline{ \\includegraphics[width=9cm]{flux_evo.ps} } \\caption{The redshift evolution of the total mass ({\\it panel a}), SSFR ({\\it panel b}), observer frame R-K color ({\\it Panel c}), and observer frame $850 \\mu$m flux ({\\it Panel d}) for 5 representative galaxies. In each panel the same greyscale corresponds to the same object. Empty symbols correspond to the $z=0$ mass of the 5 galaxies.} \\label{fig:flux_evo} \\end{figure} In this paper we have discussed the statistical properties of the ERO population as predicted by the {\\mor} galaxy formation model, in terms of number counts, redshift distributions and active fractions of $R-K>5$ sources. We generated mock photometric catalogues by interfacing {\\mor} with the spectro-photometric + radiative transfer code {\\gs}, and created large samples of model galaxies (so as to have good statistics on the rare EROs). We used both the {\\it Padova} and the M05 SSP libraries; the latters include the contribution of TP-AGB stars. We combined the resulting UV-to-Near-IR synthetic SEDs, computed without the expensive Far-IR emission, with the \\cite{Rieke09} IR template library to estimate IR fluxes longward $3 \\mu m$. We then selected samples of EROs (and sub-mm galaxies) by applying the corresponding color and flux cuts. We found that our standard model with {\\it Padova} SSP library is able to reproduce the overall number counts and redshift distribution of the $R-K>5$ sources; the agreement worsens when redder cuts are considered. Other discrepancies, like the excess of faint EROs at deep $K$ fluxes and the underestimate of the number of distant passive EROs, are consequences of well-known points of tension of the model with data, like the dearth of massive $z>2$ galaxies \\citep{Fontanot07b,Cirasuolo08,Marchesini09}, or the excess of $z\\sim1$ small galaxies \\citep{Fontanot09b}; this analysis does not reveal new discrepancies. Interestingly, {\\mor} is able to roughly reproduce the substantial contribution (of the order of $\\sim 50\\%$, see e.g. \\citealt{Cimatti02a,Miyazaki03}) of active galaxies to the global ERO population; here we define active galaxies as those with ${\\rm SSFR} > 10^{-11}$ Gyr. We ascribe this success to the higher SSFR predicted by the model at higher redshifts \\citep{Fontanot07b,Santini09}. We showed that the ERO number counts increase considerably when using the M05 SSP library. The addition of TB-AGBs gives a strong boost to the restframe SEDs of active galaxies longward $6000$ \\AA, reddening them and thus increasing the number of active EROs. This leads to a good level of agreement even at very red cuts ($R-K>6$), though it does not completely solve the lack of very red passive galaxies at higher redshift. This highlights the importance of an accurate modeling of galactic synthetic SEDs when comparing the prediction of theoretical models of galaxy formation to observations, and confirms that TP-AGBs are a relevant ingredient for the correct reproduction of the high-redshift Universe \\citep{Maraston06}. The details of the modeling of dust attenuation also play a role: we checked that both increasing the fraction of cold gas locked into molecular clouds (as suggested by \\citealt{LoFaro09} to reproduce the properties of Lyman break Galaxies at $z \\ga 4$) or using the analytical approach proposed by \\citet{DeLucia07b} based on an universal attenuation law slightly reduces the number of active EROs ($\\lesssim 0.3$ dex at $K \\sim 20$). We then considered the physical properties of the predicted ERO population with respect to a control sample of non color-selected galaxies with $K<22$ and $10.2$) active galaxies, but this trend is not visible in the overall ERO distribution. This result is in agreement with \\citet{GonzalezPerez09}. Deep fields in the Far-IR, from 100 to 500 $\\mu$m, are the best way to constrain the SSFR of these objects, but presently planned surveys with {\\it Herschel} will detect a small number of EROs at best. We also tested the possible relation between EROs and both the sub-mm ($f_{850 \\mu m}>0.5$ mJy) galaxies and the local massive ($M_\\ast > 10^{11}$ M$_\\odot$) galaxies. This was done to test the suggestion that these different populations reflect well-defined stages of a typical evolutionary sequence of massive galaxies, like sub-mm galaxies evolving into EROs and later into massive passive galaxies. We found that most massive EROs have a sub-mm bright main progenitor and evolve into a massive galaxy at $z=0$, and that many $z=0$ massive galaxies have an ERO or a sub-mm main progenitor, but only a minor fraction, $\\sim25$ per cent, follow the sub-mm -- ERO -- massive galaxy sequence. Moreover, massive galaxies may easily have more than one sub-mm and/or ERO progenitor. In fact, star formation histories predicted by {\\mor} are far more complex and exhibit a great degree of diversity, with only a small sample following the well defined sequence described above. Our work shows that EROs can be used as a powerful constraint for theories of galaxy formation and evolution, but a better and deeper understanding of the distribution of EROs in a configuration space defined by the physical quantities, like star formation activity, stellar mass, metallicity, stellar population ages and dust obscuration, is needed. Unfortunately, planned surveys with {\\it Herschel} may not reach the combination of depth and area required for detecting and analysing statistically significant samples of EROs. On the other hand, the same instruments may provide the required spectroscopic and photometric follow-up in order to better understand the real nature of already detected EROs." }, "0911/0911.5352_arXiv.txt": { "abstract": "We present MMT/Megacam imaging of the Leo~IV dwarf galaxy in order to investigate its structure and star formation history, and to search for signs of association with the recently discovered Leo~V satellite. Based on parameterized fits, we find that Leo~IV is round, with $\\epsilon < 0.23$ (at the 68\\% confidence limit) and a half-light radius of $r_{h} \\simeq 130$ pc. Additionally, we perform a thorough search for extended structures in the plane of the sky and along the line of sight. We derive our surface brightness detection limit by implanting fake structures into our catalog with stellar populations identical to that of Leo~IV. We show that we are sensitive to stream-like structures with surface brightness $\\mu_{r}\\lesssim29.6$ mag arcsec$^{-2}$, and at this limit, we find no stellar bridge between Leo IV (out to a radius of $\\sim$0.5 kpc) and the recently discovered, nearby satellite Leo V. Using the color magnitude fitting package StarFISH, we determine that Leo~IV is consistent with a single age ($\\sim$14 Gyr), single metallicity ($[Fe/H]\\sim-2.3$) stellar population, although we can not rule out a significant spread in these value. We derive a luminosity of $M_{V}=-5.5\\pm0.3$. Studying both the spatial distribution and frequency of Leo~IV's 'blue plume' stars reveals evidence for a young ($\\sim$2 Gyr) stellar population which makes up $\\sim$2\\% of its stellar mass. This sprinkling of star formation, only detectable in this deep study, highlights the need for further imaging of the new Milky Way satellites along with theoretical work on the expected, detailed properties of these possible 'reionization fossils'. ", "introduction": "Since 2005, 14 satellite companions to the Milky Way have been discovered (see \\citet{willman09} and references therein). Despite the fact that many of these objects are less luminous than a typical globular cluster ($-1.5 < M_V < -8.6$), these 14 objects have a range of properties that encompass the most extreme of any galaxies, including: the highest inferred dark matter content \\citep[e.g.][]{simongeha,geha09}, the lowest [Fe/H] content \\citep{Kirby08}, unusually elliptical morphologies \\citep[e.g. Hercules; ][]{Coleman07,sandherc}, and in some cases evidence for severe tidal disturbance \\citep[e.g. Ursa Major II; ][]{Zucker06,munoz09}. The varied properties of these lowest luminosity galaxies are valuable probes for understanding the physics of dark matter and galaxy formation on the smallest scales. Of the newly discovered Milky Way (MW) satellites, Leo~IV ($M_V = -5.5$, $r_h \\sim 130$ pc) is among the least studied, despite several signs that it is an intriguing object. Leo IV appears to be dominated by an old and metal-pool stellar population \\citep{sdsssfh}. However it also has an apparently complex color magnitude diagram (CMD), with a 'thick' red giant branch, possibly caused by either multiple stellar populations and/or depth along the line of sight \\citep{Belokurov07}. Leo IV also may have a very extended stellar distribution, despite its apparently round ($\\epsilon=0.22^{+0.18}_{-0.22}$) and compact \\citep[$r_{h}=2.5^{+0.5}_{-0.7}$ arcmin;][]{sdssstruct} morphology. A search for variable stars was recently performed by \\citet{Moretti09}, who used the average magnitude of three RR Lyrae stars to find a distance modulus of $(m-M)=20.94\\pm0.07$ mag, corresponding to $154\\pm5$ kpc. Interestingly, one of the three RR~Lyrae variables lies at a projected radius of $\\sim$10 arcmin, roughly three times the half light radius, leading to the suggestion that Leo~IV may actually possess a 'deformed morphology'. Based on Keck/DEIMOS spectroscopy of 18 member stars, Leo~IV has one of the smallest velocity dispersions of any of the new MW satellites, with $\\sigma=3.3\\pm1.7$ km/s \\citep{simongeha}. A metallicity study of 12 of these spectra \\citep{Kirby08} showed Leo~IV to be extremely metal poor, with $\\langle [Fe/H] \\rangle=-2.58$ with an intrinsic scatter of $\\sigma_{[Fe/H]}=0.75$ -- the highest dispersion among the new dwarfs. Recently, the MW satellite Leo~V ($M_V \\sim -4.3 \\pm 0.3$) has been discovered, separated by only $\\sim$2.8 degrees on the sky and $\\sim$40 km/s from Leo~IV \\citep{leov}. With a Leo~V distance of $\\sim$180 kpc, this close separation in phase space led \\citet{leov} to suggest that the Leo~IV/Leo~V system may be physically associated. This argument was bolstered by \\citet{leovspec}, who spectroscopically identified two possible Leo~V members 13 arcminutes from the satellite's center (Leo~V's $r_{h}$ is $\\sim$0.8 arcminutes) along the line connecting Leo~IV and Leo~V, suggesting that Leo~V is losing mass. A recent analysis by \\citet{leoivleov} of two 1 square degree fields situated between Leo~IV and Leo~V shows tentative evidence for a stellar 'bridge' between the two systems with a surface brightness of $\\sim$32 mag arcsec$^{-2}$. Motivated by all the above, we obtained deep photometry of Leo~IV with Megacam on the MMT. In this paper, we use these data to present a detailed analysis of both the structure and SFH of Leo~IV. We also search for any signs of disturbance in Leo~IV which may hint at a past interaction with the recently discovered, nearby Leo~V. In \\S~\\ref{sec:observations} we describe the observations, data reduction and photometry. We also present our final catalog of Leo~IV stars. In \\S~\\ref{sec:struct} we derive the basic structural properties of Leo~IV, and search for signs of extended structure. We quantitatively assess the stellar population of Leo~IV in \\S~\\ref{sec:starform} using both CMD-fitting software and an analysis of its blue plume population. We discuss and conclude in \\S~\\ref{sec:discuss}. ", "conclusions": "\\label{sec:discuss} In this work we have presented deep imaging of the Leo~IV MW satellite with Megacam on the MMT and study this galaxy's structure and SFH. In particular, we assess reports in the literature concerning both its stellar population and its possible association with the nearby satellite, Leo~V. Leo~IV's SFH is dominated by an old ($>12$ Gyr), metal poor ($[Fe/H]\\lesssim-2.0$) stellar population, although we uncover evidence for a young sprinkling of star formation 1-2 Gyrs ago. Our best-fit StarFISH results indicate that a single metal poor population dominates, although the data is also compatible with a spread in metallicities. The old population is consistent with the emerging picture that the faintest MW satellites are 'reionization fossils' \\citep[e.g.~][]{Ricotti05,Gnedin06}, who formed their stars before reionization and then lost most of their baryons due to photoevaporation. The apparent sprinkling of young stars begs the question of what has enabled Leo~IV to continue forming stars at a low level. There is no sign of HI in Leo~IV, with an upper limit of 609 $M_{\\odot}$ \\citep{Grcevich09}, although we note that this limit is still a factor of $\\sim$2 larger than the stellar mass associated with the young stellar population studied in \\S~\\ref{sec:BP}. One possible mechanism for late gas accretion, and subsequent star formation, among the faint MW satellites was recently discussed by \\citet{Ricotti09} to help explain the complex SFH and gas content of Leo~T \\citep{leot,leotgas}. In this scenario, the smallest halos stop accreting gas after reionization as expected, but as their temperature decreases and dark matter concentration increases with decreasing redshift they are again able to accrete gas from the intergalactic medium at late times, assuming they themselves are not accreted by their parent halo until $z\\lesssim1-2$. This can lead to a bimodality in the SFH of the satellite, with both a $>12$ Gyr population and one that is $<10$ Gyr, as we see in Leo~IV. One stringent requirement of this model is that the satellite can not have been accreted by its host halo until $z\\lesssim1-2$ (and thus not exposed to tidal stirring and ram pressure stripping until late times, allowing the satellite to retain its newly accreted gas). Future proper motion studies will be able to test if Leo~IV is compatible with this late gas accretion model. Another prediction of this model is the possible existence of gas-rich minihaloes that never formed stars, but could serve as fuel for star formation if they encountered one of the luminous dwarfs. More detailed study of this late gas accretion mechanism will be necessary to understand the possible diversity of SFHs in the faint MW satellites. Additionally, we note that if the apparent segregation of young stars in Leo~IV is real, then it is not an isolated case among the MW satellites. As has been mentioned in \\S~\\ref{sec:BP}, \\citet{Martin08} noted that Canes~Venatici I has a compact star forming region clearly offset from the galaxy as a whole, with an age of 1-2 Gyrs, similar to Leo~IV. Additionally, Fornax has several compact clumps and shells that house young stellar populations roughly $\\sim$1.4 Gyrs old \\citep{coleman04,Coleman05,Olszewski06}. It has been suggested that these could have been the result of a collision between Fornax and a low-mass halo, which was possibly gas-rich. \\citet{pennarubia09} investigated the disruption of star clusters in triaxial, dwarf-sized halos and found that segregated structures can persist depending on the orbital properties of the cluster, providing yet another viable mechanism. More work is needed to distinguish between all of the above scenarios and to properly model the emerging diversity of SFHs among the new, faint MW satellites. Structurally, Leo~IV appears to be very round, with $\\epsilon \\lesssim 0.23$ (at the 68\\% confidence limit) and a half light radius ($\\sim 130$ pc) which is typical of the new MW satellites. An exhaustive search for signs of extended structure in the plane of the sky has ruled out any associated streams with surface brightnesses of $\\mu_{r}\\lesssim29.6$. The extent of Leo~IV along the line of sight is less than $\\sim$15 kpc, a limit that will be difficult to improve upon given the inherent limitations of using the spread in BHB magnitudes to measure depth. We find no evidence for structural anomalies or tidal disruption in Leo~IV. We do not have the combination of image depth and area necessary to confirm the stellar bridge, with a surface brightness of 32 mag arcsec$^{-2}$, recently reported in between Leo~IV and Leo~V \\citep{leoivleov}. Indeed, Leo~V is almost certainly disrupting, as discussed by \\citet{leovspec}, due to the presence of two member stars $\\sim$13 arcminutes ($>13 r_{h}$) from Leo~V's center along the line connecting the putative Leo~IV/Leo~V system. The nature of Leo~V is still very ambiguous, with the kinematic data being consistent with it being dark matter free -- suggesting that perhaps Leo~V is an evaporating star cluster \\citep{leovspec}. It is thus critical to obtain yet deeper data on these two systems and the region separating them to uncover their true nature. The probable detection of a small population of young stars illustrates once again that it is crucial to obtain deep and wide field follow up for all of the newly detected MW satellites. Every new object has a surprise or two in store upon closer inspection." }, "0911/0911.5164_arXiv.txt": { "abstract": "New astrophysical instruments such as skA (square kilometer Array) and IXO (formerly Constellation X) promise the discovery of tens of thousands of new isolated rotating neutron stars (pulsars), neutron stars in low-mass X-ray binaries (LMXBs), anomalous X-ray pulsars (AXPs), and soft gamma repeaters (SGRs). Many of these neutron stars will experience dramatic density changes over their active lifetimes, driven by either stellar spin-up or spin-down, which may trigger phase transitions in their dense baryonic cores. More than that, accretion of matter onto neutron stars in LMXBs is believed to cause pycno-nuclear fusion reactions in the inner crusts of neutron stars. The associated reaction rates may be drastically altered if strange quark matter would be absolutely stable. This paper outlines the investigative steps that need to be performed in order to explore the thermal response of neutron stars to rotationally-driven phase transitions in their cores as well as to nuclear burning scenarios in their crusts. Such research complements the exploration of the phase diagram of dense baryonic matter through particle collider experiments, as performed at RHIC in the USA and as planned at the future Facility for Antiproton and Ion Research (FAIR) in Darmstadt, Germany. ", "introduction": "On the Earth, particle collider experiments enable physicists to cast a brief glance at the properties of ultra-dense and hot baryonic matter (Fig. \\ref{fig:FAIR}). On the other hand, it is estimated that galaxies like our Milky Way contain between $10^8$ and $10^{10}$ collapsed stars known as neutron stars, which harbor ultra-dense matter permanently in their cores. This key feature together with the unprecedented progress in observational astrophysics, which is expected to be excelled by future observatories such as the square kilometer Array (skA) and Constellation-X, make neutron stars superb astrophysical laboratories for a wide range of physical studies. These studies concern nuclear fusion processes on the stellar surface, pycnonuclear reactions in electron degenerate matter at sub-nuclear densities, and the possible formation of boson condensates and other novel states of baryonic matter--like color superconducting quark matter--at super-nuclear densities. (For overviews, see, for instance \\cite{glen97:book,weber99:book,blaschke01:trento,lattimer01:a,rajagopal01:a,% weber05:a,klahn06:a_short,page06:a,sedrakian07:a,alford08:a}.) More than that, there is the very intriguing theoretical suggestion that strange quark matter could be more stable than atomic nuclei, known as the strange quark matter hypothesis, in which case neutron stars should be largely composed of pure strange quark matter \\cite{alcock86:a,alcock88:a,madsen98:b}. If quark matter exists in neutron stars it ought to be a color superconductor \\cite{rajagopal01:a,alford01:a,alford98:a,rapp98+99:a}. Other testable implications of the strange quark matter hypothesis concern the \\begin{figure}[tb] \\begin{center} \\includegraphics*[width=0.60\\textwidth,angle=0]{FAIR_FWeber3.eps} \\caption[]{Schematic phase diagram of strongly interacting matter \\cite{FAIR09:phasediagram}. The dashed circle indicates the portion of the phase diagram that can be explored by studying either hot and newly formed proto neutron stars (arrow pointing downwards), or cold rotating neutron stars (vertical double-headed arrow) such as millisecond pulsars and neutron stars in low-mass X-ray binaries (LMXBs).} \\label{fig:FAIR} \\end{center} \\end{figure} possible existence of a new class of white dwarfs, known as strange white dwarfs \\cite{glen92:crust,glen94:a}, and the drastic alteration of heavy-ion reaction rates in the deep crustal layers of neutron stars \\cite{golf09:a}, if strange quark matter nuggets should be present in these layers. ", "conclusions": "" }, "0911/0911.2328_arXiv.txt": { "abstract": "Gamma-ray bursts (GRBs) are the brightest events in the universe. They have been used in the last five years to study the cosmic chemical evolution, from the local universe to the first stars. The sample size is still relatively small when compared to field galaxy surveys. However, GRBs show a universe that is surprising. At $z>2$, the cold interstellar medium in galaxies is chemically evolved, with a mean metallicity of about 1/10 solar. At lower redshift ($z<1$), metallicities of the ionized gas are relatively low, on average 1/6 solar. Not only is there no evidence of redshift evolution in the interval $0 20.2$ and $\\log N_{\\rm ZnII} > 12.4$. The completeness level for GRB-DLAs is not well determined. It is apparent that column densities in GRB-DLAs are generally higher than in QSO-DLAs. This can either indicate that GRB-DLAs originate in bigger galaxies, or that the volume density of the gas is higher (e.g., the GRB sightline is crossing a region closer to the galaxy center) than QSO-DLAs, or both.} \\label{f1} \\end{center} \\end{figure} \\begin{figure}[b] \\begin{center} \\includegraphics[width=4in]{fig2.ps} \\caption{Redshift evolution of the metallicity relative to solar values, for 17 GRB-DLAs at $z>2$, 16 GRB hosts at $z<1$ and $\\sim250$ QSO-DLAs in the interval $0 2$ is not low. The measured average value is 1/10 solar, and the dispersion is large, about a factor of five. Relatively high metallicity is confirmed also for the highest redshift detections (Savaglio 2006; Totani et al.\\ 2006; Price et al.\\ 2007), which means that there is no indication of redshift evolution. Low metallicities of the GRB progenitor are theoretically predicted (no mass loss) in order to keep a high angular momentum, and have a highly collimated jet. Relatively high metallicities, in this case from ionized gas of the host galaxies, are confirmed also at $z<1$. The average value is 1/6 solar, with a dispersion of a factor of two, indicating that the metal content in host galaxies is not evolving so fast. The sample is not very large and systematic uncertainties are still not totally under control, therefore more observations and detections are very important. Nevertheless, the lack of evidence of redshift evolution and the observed large dispersion suggest that GRBs do not happen necessarily in metal poor galaxies. Star formation, on the other hand, might be a more important physical trigger. GRBs are extremely important for our understanding of the primordial universe and the formation and evolution of heavy elements. The enlightening discoveries of the last few years are a clear indication that the investigation is affected by our technical capabilities which have dramatically improved recently. Dedicated instruments and observational programs have opened a new window in the hidden universe and show that this is more surprising and fascinating than expected." }, "0911/0911.2434_arXiv.txt": { "abstract": "The observed super-massive black hole (SMBH) mass -- galaxy velocity dispersion ($M_{\\rm bh} - \\sigma$) correlation may be established when winds/outflows from the SMBH drive gas out of the potential wells of classical bulges. Here we present numerical simulations of this process in a static isothermal potential. Simple spherically symmetric models of SMBH feedback at the Eddington luminosity can successfully explain the $M_{\\rm bh} - \\sigma$ and nuclear cluster mass $M_{\\rm NC}-\\sigma$ correlations, as well as why larger bulges host SMBHs while smaller ones host nuclear star clusters. However these models do not specify how SMBHs feed on infalling gas whilst simultaneously producing feedback that drives gas out of the galaxy. More complex models with rotation and/or anisotropic feedback allow SMBHs to feed via a disc or regions not exposed to SMBH winds, but in these more realistic cases it is not clear why a robust $M_{\\rm bh} - \\sigma$ relation should be established. In fact, some of the model predictions contradict observations. For example, an isotropic SMBH wind impacting on a disc (rather than a shell) of aspect ratio $H/R \\ll 1$ requires the SMBH mass to be larger by a factor $\\sim R/H$, which is opposite to what is observed. We conclude that understanding how a SMBH feeds is as important a piece of the puzzle as understanding how its feedback affects its host galaxy. Finally, we note that in aspherical cases the SMBH outflows induce differential motions in the bulge. This may pump turbulence that is known to hinder star formation in star forming regions. SMBH feedback thus may not only drive gas out of the bulge but also reduce the fraction of gas turned into stars. ", "introduction": "It is believed that the centres of most galaxies contain super-massive black holes (SMBHs) whose mass $\\mbh$ correlates with the velocity dispersion $\\sigma$ of the host galaxy \\citep{Ferrarese00,Gebhardt00,Tremaine02}. Similarly, there is a correlation between $\\mbh$ and the mass of the bulge, $M_{\\rm bulge}$ for large SMBH masses \\citep{Magorrian98,Haering04,GultekinEtal09}. Observations also suggest that the masses of nuclear star clusters (NC) ($10^5 \\msun \\simlt M_{\\rm NC} \\simlt 10^8 \\msun$) correlate with the properties of their host dwarf ellipticals \\citep{FerrareseEtal06,Wehner06} in a manner that is analogous to the one between SMBHs and their host ellipticals. These empirical relations can be explained in a very natural way if the growth of host galaxies and their central SMBHs or NCs are linked by feedback. This was first pointed out by \\cite{SilkRees98}, who highlighted the potential importance of SMBH heating and outflows before any robust observational evidence for such a link had been found. Subsequently, \\cite{Fabian99} argued that radiation pressure acting on cold gaseous clouds in the bulge could give rise to the observed correlations. \\cite{KP03} argued for the existence sub-relativistic outflows from the very central regions of AGN, which prompted \\cite{King03,King05} to study the motion of a shell of gas swept up by a wind/outflow from a central black hole in a galactic isothermal dark matter potential. \\cite{King05} demonstrated that the shell will be expelled from the potential provided the black hole mass exceeds a critical value that, as a function of $\\sigma$, turns out to be close to the observed $M_{\\rm BH}$--$\\sigma$ relation. The main result of \\cite{King03,King05} can be deduced using a simpler order of magnitude ``weight argument''. According to this argument, the SMBH luminosity is assumed to be limited by the Eddington value. Radiation pressure drives a wind. \\cite{KP03} argue that wind velocity is comparable to the escape velocity from the inner accretion disc, e.g., $v \\sim 0.1 c$. The momentum outflow rate is assumed to be \\begin{equation} \\dot P_{\\rm SMBH} \\approx {L_{\\rm Edd}\\over c} = \\frac{ 4 \\pi G M_{\\rm BH}}{\\kappa}\\;; \\label{pismbh} \\end{equation} \\noindent here $\\kappa$ is the electron scattering opacity and $M_{\\rm BH}$ is the SMBH mass. This result is natural to the order of magnitude: $L_{\\rm Edd}/ c$ is the radiation momentum flux which is presumably passed to the wind as radiation accelerates the outflow \\citep[cf.][]{KP03}. The argument also assumes that the black hole wind is optically thin at the point of interaction with the ambient gas. Because the cooling time of the shocked gas is short on scales appropriate for observed bulges \\citep{King03,King05}, the bulk energy of the outflow is thermalised and quickly radiated away. It is then only the momentum push (equation \\ref{pismbh}) of the outflow on the ambient gas that is important since it is this that produces the outward force on the gas\\citep[as in the earlier model by][]{Fabian99}. The weight of the gas is $W(R) = GM(R)[M_{\\rm total}(R)]/ R^2$, where $M_{\\rm gas}(R)$ is the enclosed gas mass at radius $R$ and $M_{\\rm total}(R)$ is the total enclosed mass including dark matter. For an isothermal potential, $M_{\\rm gas}(R)$ and $M_{\\rm total}(R)$ are proportional to $R$, so the result is \\begin{equation} W = {4f_g\\sigma^4\\over G}\\;. \\label{w} \\end{equation} Here $f_g$ is the baryonic fraction and $\\sigma^2 = GM_{\\rm total}(R)/2R$ is the velocity dispersion in the bulge. By requiring that momentum output produced by the black hole just balances the weight of the gas, it follows that \\begin{equation} M_{\\sigma} = {f_g\\kappa\\over \\pi G^2}\\sigma^4, \\label{msigma} \\end{equation} \\noindent which is consistent with the observed the $M_{\\rm BH}$--$\\sigma$ relation. To order of magnitude, relation \\ref{w} should hold for any potential at the virial radius. Therefore the model by \\cite{King03,King05} appears to be a promising explanation of the observed correlations between SMBHs and their host galaxies. However, additional complications, not considered in \\cite{King03,King05}, may be important. Amongst these are (1) the self-gravity of the gas, because gas that accumulates at the centre of the potential may begin to dominate it and therefore alter the result; (2) finite angular momentum of the gas may lead to formation of a disc; (3) collimated and variable outflows. \\\\ The goal of this paper is to verify that the predictions of the analytical model of \\cite{King03,King05} hold and to explore more realistic settings, of the kind just described. Note that we concentrate on the early stages of galaxy evolution, when we would expect galaxies to be gas-rich and the baryonic mass within dark matter haloes is dominated by gas rather than stars. At later times, when the SMBH is likely to be less luminous and the galaxy is less gas-rich, we would expect different forms of feedback to become important, such as relativistic jets and the associated radio bubbles \\citep[e.g.,][]{ChurazovEtal02} or pre-heating due to the inverse Compton effect \\citep[e.g.,][]{SazonovEtal04}. However, we do not include these forms of feedback in our current simulations. ", "conclusions": "Our main conclusions are \\begin{itemize} \\item Predictions of spherically symmetric models with black hole feedback tied to Eddington limit luminosity as in the models of \\cite{King03,King05} are confirmed numerically. Self-gravity of the gas complicates evolution of the system, and a self-consistent treatment of star formation and its feedback is necessary for further progress. \\item As suggested earlier based on analytical arguments, it is the dynamical time in the bulge, $R_{\\rm b}/\\sigma$, that determines whether or not the SMBH can reach its limiting $M_\\sigma$ value. Central regions of smaller bulges, where the dynamical time is observed to be shorter than the Salpeter time, could be smothered with infalling gas despite the SMBH feedback. This process may be the origin of the nuclear star clusters in ``smaller'' galaxies. \\item A net angular momentum in the shell is essential in determining the fate of the shell and the SMBH feeding. There is no well defined $M_\\sigma$ mass in that case since the momentum thrust required is different in different directions. Gas near the symmetry axis is blown out easier than in the spherically symmetric case, whereas gas settled into a disc requires $\\sim R/H$ more thrust than in the latter case. \\item If the SMBH is fed through a several kpc scale disc, the SMBH mass would have to be $\\sim R/H$ larger than in the spherical case to expel the feedback-resistant self-shielding disc. This directly contradicts the recent observations of \\cite{Hu09} that show that pseudo-bulges have lighter SMBHs than their classical counter-parts. This may imply that black holes are not fed by large (kpc or larger) discs but rather by flows with a small specific angular momentum. \\item We also noted that SMBH outflows in a realistic, i.e., aspherical, situation may pump turbulence (differential motions) in the bulge. Turbulence is known to hinder star formation. Therefore we believe that AGN feedback not only expels the gas from the galaxy but it may also reduce the amount of mass turned into stars during bulge formation. \\end{itemize}" }, "0911/0911.3118_arXiv.txt": { "abstract": "We propose a mechanism for the creation of cosmic string loops with dynamically stabilised windings in the internal space. Assuming a velocity correlations regime in the post-inflationary epoch, such windings are seen to arise naturally in string networks prior to loop formation. The angular momentum of the string in the compact space may then be sufficient to ensure that the windings remain stable \\emph{after} the loop chops off from the network, even if the internal manifold is simply connected. For concreteness we embed our model in the Klebanov-Strassler geometry, which provides a natural mechanism for brane inflation, as well a being one of the best understood compactification schemes in type IIB string theory. We see that the interaction of angular momentum with the string tension causes the loop to oscillate between phases of expansion and contraction. This, in principle, should give rise to a distinct gravitational wave signature, the future detection of which could provide indirect evidence for the existence of extra dimensions. ", "introduction": "The existence of string loops with dynamically stabilised winding in a compact space was first demonstrated by Iglesias and Blanco-Pillado \\cite{BlancoPillado:2005dx}, using the Klebanov-Strassler geometry \\cite{Klebanov:2000hb}. They considered strings at the tip of the throat, with geodesic wrappings in the $S^3$ which regularises the conifold singularity. Although they derived a lower bound for the angular momentum of a loop with a given number of windings - below which the windings became unstable - the result must remain of purely theoretical interest to cosmology as long as a specific mechanism for winding formation (and hence for the formation of the string angular momentum) is not considered. The purpose of this paper is to propose such a mechanism, which leads to the formation of string configurations such as those investigated in \\cite{BlancoPillado:2005dx} and to investigate the resulting string dynamics with specific reference to their cosmologically observable consequences. \\n Assuming that a velocity correlations regime in the post-inflationary epoch as in \\cite{Avgoustidis:2005vm} leads naturally to the formation of geodesic windings we show that the winding number ($n$), total energy ($E$) and angular momentum ($l$) of the string are specified precisely by the model parameters. That is, by the parameters which define the Klebanov-Strassler geometry (in this case specifically by the value of the warp factor $a_0$) and those which determine the scale of the string network ($\\alpha$, $t_i$). Substituting for $l$ and $n$ in the bound referred to above then demonstrates the stability of these windings, at least under the assumption that $l$ remains approximately constant over small time scales after the moment of initial loop formation. By assuming also that the total energy of the string remains approximately constant (i.e. by neglecting the loss of $E$ and $l$ via gravitational wave emission over small time scales), we then determine the equation of motion for the four dimensional string radius $r(t)$, and solve to find a (generically) oscillating solution. Crucially we observe that the qualitative behaviour of the loop depends on the value of the warp factor $a_0$ with $a_0^2<1/2$ leading to an initial phase of expansion and $a_0^2>1/2$ leading to an initial phase of contraction. The fixed point solution $a_0^2 = 1/2$ is a static, non-oscillatory solution. In both oscillatory modes (initially contracting \\emph{and} initially expanding) we find that the period of the oscillation is inversely proportional to $a_0^2$, and proportional to the initial size of the loop $(\\alpha t_i)$. Following \\cite{Avgoustidis:2005vm} we refer to these objects as non-topological cycloops \\footnote{The term 'cycloop' was first coined in \\cite{Avgoustidis:2005vm} to refer to cosmic string loops with \\emph{smooth} windings wrapping cycles in the internal space - as opposed, for example, to non-smooth, step-like windings which give rise to necklace configurations from a four dimensional perspective \\cite{Matsuda:2006ju, Matsuda:2005ez, Lake:2009nq}. However in their original conception Avgoustidis and Shellard used it only to refer to windings which are topologically trapped. We therefore propose the term 'non-topological cycloops' to refer to string loops with dynamically stabilised smooth windings (in this case geodesics) around a simply connected compact manifold.}. \\n The layout of the paper is then as follows: In Section 2 we briefly review the relevant Klebanov-Strassler background, focusing on the geometry of the conifold tip. In Section 3 we show how the assumption of a velocity correlations regime yields a dynamical model of winding formation after the end of inflation. Section 4 recaps the generic results of \\cite{BlancoPillado:2005dx} which we then combine with the results of the previous section to give explicit expressions for the winding number, energy and angular momentum of the string. The equation of motion for the loop radius is then derived and solved in Section 5 and a brief summary of the main results together with a discussion of their cosmological implications and possible consequences for experimental observations is presented in section 6. Finally the two appendices at the end of this work deal with issues which arise within the text: Appendix I outlines the method of Eulerian substitution of the third kind which is used to integrate the differential equation involving $\\dot{r}(t)$ and $r(t)$ derived in Section 5. Appendix II gives a detailed description of the Hopf fibration of the 3-sphere, which is introduced briefly in section 2 and used throughout the following analysis. ", "conclusions": "We have argued that a velocity correlations regime in the post-inflationary epoch leads naturally to the formation of string loops with geodesic windings in the compact space. For strings at the tip of the conifold throat of the Klebanov-Strassler geometry we were able to show that the quantities which determine the dynamical evolution of the loop (i.e. the initial winding number $n(t_i)$, energy $E(t_i)$ and string velocity/angular momentum in the compact space, $\\dot{s}(t_i)$/ $l(t_i)$) are \\emph{uniquely} determined by the parameters $a_0^2$, $\\alpha$ and $t_i$. Crucially these windings were found to have sufficient angular momentum in the compact directions to remain stable, even after the string chops off from the network to form a loop. \\n The interaction between the tension and the angular momentum in the compact space was found to play a significant role in the dynamical evolution of the string, including - perhaps surprisingly - the evolution in four dimensions. By assuming energy loss (and angular momentum) via gravitational radiation to be negligible, we determined equations of motion for the four-dimensional radius $r(t)$ and the string velocity $v(t) \\sim \\dot{s}(t)R$, which we believe to be valid over small time scales after the moment of loop formation, $t=t_i$. \\n We found that the qualitative behaviour of the string depends crucially on the square of the warp factor, $01/2$ leads to oscillations with an initial contracting phase. In each case the string was seen to oscillate between it's initial radius $r(t_i)=(\\alpha t_i)$ and a secondary critical value defined by $ r_{c2}=\\frac{(1-a_0^2)}{a_0^2}(\\alpha t_i)$, with the period of oscillation given by $T = \\frac{(\\alpha t_i)} {a_0^2}$. \\n As noted above, in the two oscillatory regimes it is the interaction of the angular momentum with the string tension which \"drives\" the dynamical evolution converting kinetic energy into rest mass during expansion (with the inverse process occurring in the contracting phase). The string is seen to evolve \\emph{towards} a static, minimum energy configuration where $l=2\\pi a_0^2 T_1$ and $\\dot{s}^2 = a_0^2/R^2$, but is unable to satisfy these two conditions simultaneously at a point where $\\dot{r}(t)=0$. By contrast, in the $a_0^2=1/2$ scenario we find that the energy minimising conditions \\emph{are} satisfied simultaneously at the moment of loop formation. In this case the tension exactly offsets the effect of angular momentum so that the string remains static at it's original radius $r(t_i)=(\\alpha t_i)$ and is non-oscillatory. \\n The meaning of the term \"small\" above is somewhat ambiguous, but it seems reasonable to assume that our solution will provide a valid approximation over at least one full oscillatory cycle of the loop; that is, over a time period $\\Delta t \\sim T = \\frac{(\\alpha t_i)}{a_0^2}$. For time periods $\\Delta t >> T$ our initial analysis must be extended to include the effects of gravitational wave emission on $E(t)$ and $l(t)$ (or equivalently on $r(t)$ and $\\dot{s}(t)$). We may expect the qualitative effect of energy and angular momentum loss to be relatively simple, as long as the string retains sufficient angular momentum for the extra-dimensional windings to remain stable. In this case it is likely that the loss of $E$ and $l$ due to gravitational wave emission will simply act to damp the oscillations of $r(t)$ and $\\dot{s}(t)$. What is unclear however is whether or not the damping coefficient itself is likely to be time-dependent; For example, it is possible that smaller oscillations lead to greater rates of emission per unit length (as the rate of acceleration $\\ddot{r}(t)$ is higher in this case) so that the damping itself increases with time (for an individual loop). \\n In the case that the windings eventually become unstable however (as $l(t)$ drops below the threshold for ensuring their stability) the string dynamics are likely to become complicated, and it is not clear whether the process of winding contraction (i.e. of windings \"falling off\" the $S^3$) may even be accommodated within an analysis which uses an ansatz of the form (\\ref{eq:ansatz}) to describe the string configuration. This is because the coordinates $r(t)sin\\sigma$, $r(t)cos\\sigma$, $\\psi(\\sigma,t)$, $\\theta(\\sigma,t)$ and $\\phi(\\sigma,t)$ are treated as \\emph{independent} variables with respect to the determination of the equations of motion. We are therefore unable to take account of the continuously connected nature of the string when string sections \"move\" from one direction to another (e.g. in \"falling off\" the $S^3$ to form part of the four dimensional rest mass of the loop). \\n The present analysis could of course still be improved by accounting for the effects of gravitational wave emission under the assumption that that loops retain their windings, which would indeed be valid up to the point where $l(t)$ drops below the critical value (which may also be calculated). Such an improved analysis could proceed as follows: one could compute the stress energy tensor for an oscillating loop and look for solutions with this as a source to the Einstein equations. This should allow us to estimate the rate of loss of $E$ and $l$ via gravitational wave emission and, as stated, we expect this to produce a damping term in the equations for $r(t)$ and $\\dot{s}(t)$. \\n Calculating the emission spectrum for an oscillating loop would also be of immense practical interest. The gravitational wave signature of such a loop, whose self-oscillation is caused by the presence of angular momentum in the compact space, may differ significantly from that of a string whose oscillations, though superficially similar, do not result from self-interaction. Gravitational wave emission from loops oscillating with period $\\omega \\sim L^{-1}$ (where $L$ is the loop size) have been intensively studied in four dimensions \\cite{Vaschpati+Vilenkin:1985, Burden:1985, Garfinkle+Vaschpati:1987, Allen+Shellard:1992}. However in such cases the loop is not undergoing \\emph{genuine} phases of expansion and contraction, but rather experiencing \"wiggles\" of a size comparable to it's own length. Although Weinberg \\cite{Weinberg:1972} has has shown that - in an FRW universe - the power of a weak, isolated, periodic source (to lowest order in $G$) may be given by a single formula \\emph{regardless} of the exact nature of the source, it is not immediately clear that this should hold in extra-dimensional scenarios. Additionally such sources (i.e. loops) have no angular momentum to shed in the process of emission. In fact even if we were to study loops whose self-oscillation was due to their \"rotational\" motion in Minkowski space (c.f. \\cite{Durrer:1989, Vilenkin+Shellard:2000}), the angular momentum which would be carried away by gravitational radiation would be very different from that lost via emission from oscillating cycloops \\footnote{We note also that one possible further extension of the current analysis would be to consider a wound string with rotational motion in both the compact \\emph{and} Minkowski directions.}. The detection of such a signature by the future generation gravitational wave detectors could then provide indirect evidence of the presence of compact dimensions. We hope therefore to be able to provide such an analysis in the near future. \\n Although the full string theory compactification favours exponentially small values of $a_0^2$, it is also interesting to note that the possibility of obtaining evidence from the gravitational wave signature of wound strings exists even in the case of static loops (i.e. the $a_0^2 = 1/2$ case). These loops, which form automatically with a configuration which meets the energy minimising conditions, look just like ordinary string loops from a four-dimensional perspective. However they contain an \"unseen\" angular momentum which is not directly manifested in their dynamical evolution in Minkowski space. As stated above, we expect that even \"standard\" four dimensional string loops may undergo periodic oscillations with $\\omega \\sim L^{-1}$ creating ripples in space-time and giving rise to gravitational waves. Such fluctuations \\emph{along} the length of the string - though not in the total four dimensional length itself - would likely still occur in this case (though it is possible that the existence of the angular momentum term may lend a certain 'rigidity' to the circular string configuration, making it more resistant to deformation) but the string must also now shed it's angular momentum. The gravitational wave signature of even a non-self-oscillating loop in the extra-dimensional scenario is therefore also likely to differ significantly from the standard case of an un-wound string, though perhaps to a less significant degree than in the self-oscillating case. However the possibility of the indirect detection of compact dimensions from cosmic strings therefore remains, even in the absence of self-oscillating loops. We hope also to be able to offer an analysis of this interesting possibility in a future letter. \\n It remains for us to outline some of the possible limitations of the analysis we have attempted: We have assumed throughout the present work that the velocity correlations regime leads naturally to a) geodesic windings and b) movement of the string parallel to itself (i.e. along the geodesics). However these assumptions may be questioned. The rationale for adopting such an approach (which simplified the resulting analysis considerably) was simple - we expect velocity correlations to impart a constant angular momentum density to each point along the string. Therefore we expect each \"point\" (or infinitesimal string segment) to travel along a geodesic curve, both before and after loop formation. However it has not been proved that each point along the string traversing a separate geodesic, leads to a winding configuration that is itself geodesic. Likewise it does not necessarily follow that the resulting motion of the string as a whole is parallel to itself. In general, we may expect that the geometry of the internal space plays a role in determining the exact nature of the winding configuration and the resulting string motion. Moreover we have not included the contribution from the Ramond-Ramond (RR) sector, which in principle will couple to the string. This charge term is ultimately crucial for distinguishing between cosmic strings and cosmic super-strings. \\n Whilst, in principle, motion of the string perpendicular to its length may easily be accounted for (see end of Appendix II) the most significant problem in the analysis of string loops with non-geodesic windings arises from the resulting $\\sigma$-dependence of the expressions for $E$ and $l$. We believe that the present analysis may therefore be improved by a more thorough investigation of the winding process itself, and though we expect our expressions for $n(t_i)$ (\\ref{eq:n2}) and $v(t_i) \\sim \\dot{s}(t_i)R$ (\\ref{eq:omega_l_v}) to remain valid (in the KS case), the appropriate string ansatz may well be more complicated. Although it seems unlikely that the qualitative behaviour of the string will differ significantly than described in the scenarios above, finding analogous (and \\emph{quantitatively} different) results may be extremely difficult. But ultimately exact \\emph{quantitative} predictions will be needed for any comparison with future experimental data, and we believe such a project to be worthwhile. \\begin{center} {\\bf Acknowledgments} \\end{center} M.L is supported by an STFC studentship and the Queen Mary EPSTAR consortium. This work was supported in part by NSERC of Canada. \\renewcommand{\\theequation}{A-\\arabic{equation}} \\setcounter{equation}{0} %" }, "0911/0911.0722.txt": { "abstract": "The Extreme ultraviolet SpectroPhotometer (ESP) is one of five channels of the Extreme ultraviolet Variability Experiment (EVE) onboard the NASA {\\it Solar Dynamics Observatory} (SDO). The ESP channel design is based on a highly stable diffraction transmission grating and is an advanced version of the Solar Extreme ultraviolet Monitor (SEM), which has been successfully observing solar irradiance onboard the {\\it Solar and Heliospheric Observatory} (SOHO) since December 1995. ESP is designed to measure solar Extreme UltraViolet (EUV) irradiance in four first order bands of the diffraction grating centered around 19 nm, 25 nm, 30 nm, and 36 nm, and in a soft X-ray band from 0.1 to 7.0~nm in the zeroth order of the grating. Each band's detector system converts the photo-current into a count rate (frequency). The count rates are integrated over 0.25~sec increments and transmitted to the EVE Science and Operations Center for data processing. An algorithm for converting the measured count rates into solar irradiance and the ESP calibration parameters are described. The ESP pre-flight calibration was performed at the Synchrotron Ultraviolet Radiation Facility of the National Institute of Standards and Technology. Calibration parameters were used to calculate absolute solar irradiance from the Sounding Rocket flight measurements on 14 April 2008. These irradiances for the ESP bands closely match the irradiance determined for two other EUV channels flown simultaneously, EVE's Multiple Euv Grating Spectrograph (MEGS) and SOHO's Charge, Element and Isotope Analysis System / Solar EUV Monitor (CELIAS/SEM). ", "introduction": "\\label{S-Introduction} The {\\it Solar Dynamics Observatory} (SDO) is the first NASA Living With a Star mission with its launch planned for February 2010. It will provide accurate measurements of the solar atmosphere characteristics with high spatial and temporal resolution at many wavelengths simultaneously. These measurements will help us understand the solar activity cycle, the dynamics of energy transport from magnetic fields to the solar atmosphere, and the influences of this energy transport on the Earth's atmosphere and the heliosphere. Dynamic changes of the solar radiation in the extreme ultraviolet (EUV) and X-ray regions of the solar spectrum are efficient drivers of disturbances in the Earth's space weather environment. The Extreme ultraviolet Variability Experiment (EVE) is one of three instrument suites on SDO. EVE measures the solar EUV irradiance with unprecedented spectral resolution, temporal cadence, accuracy, and precision. Furthermore, the EVE program will incorporate physics-based models of the solar EUV irradiance to advance the understanding of solar dynamics based on short- and long-term activity of the solar magnetic features \\cite{Woods06, Woods09}. ESP is one of five channels \\cite{Woods06} in the EVE suite. It is an advanced version of the SOHO/CELIAS Solar Extreme ultraviolet Monitor (SEM) \\cite{Hovestadt95, Judge98}. SEM measures EUV solar irradiance in the zeroth diffraction order (0.1 to 50.0~nm bandpass) and two (plus and minus) first order diffraction bands (26 to 34~nm bandpasses) centered at the strong He II 30.4~nm spectral line. More than 13 years of EUV measurements have shown that the SEM is a highly accurate and stable EUV spectrometer \\cite{Judge08} that has suffered only minor degradation, mainly related to deposition of carbon on the SEM aluminum filters. The ESP design is based on a highly stable diffraction transmitting grating \\cite{Schattenburg90, Scime95}, very similar to the one used in SEM. ESP has filters, photodiodes, and electronics with characteristics (transmission, sensitivity, shunt resistance, {\\it etc.}) that are susceptible to some change over the course of a long mission. Because of such possible degradation ESP shall be periodically calibrated throughout the mission. The calibration program includes a pre-flight calibration followed by a number of sounding rocket under-flights with a nominally identical prototype of the flight instrument that is typically calibrated shortly before and shortly after each under-flight. Calibration of the SDO/EVE and EVE/ESP soft X-ray and EUV flight instruments was performed at the National Institute of Standards and Technology (NIST) at the Synchrotron Ultraviolet Radiation Facility (SURF-III). NIST SURF-III has a number of Beamlines (BL) with different specifications and support equipment ({\\it i.e.} monochromators, translation stages and calibration standards). The beamline used for a given instrument depends on the instrument's characteristics and calibration requirements such as spectral ranges, entrance aperture, alignment, {\\it etc.} Before the flight ESP was integrated into the EVE package, it was first calibrated on SURF BL-9. BL-9 is equipped with a monochromator capable of scanning through a wide EUV spectral band which allows the instrument efficiency profile for ESP to be measured in increments of 1.0~nm over a spectral window of about 15 to 49~nm. After ESP was mounted onto EVE, the second ESP calibration was performed on SURF BL-2 \\cite{Furst93}. BL-2 illuminates the instrument with the whole synchrotron irradiance spectrum, which can be shifted in its spectral range by changing the energy of the electrons circulating in the electron storage ring producing the synchrotron radiation. The intensity of the EUV beam is controlled by the current (and therefore number) of circulating electrons. Because the EUV beam consists of a wide spectrum of photon energies, this type of calibration is a radiometric calibration. A major part of this paper is related to the results of the ESP BL-9 and BL-2 calibrations. The scientific objectives of ESP are given in Section 2. Section 3 presents a short overview of the ESP channel. Section 4 describes an algorithm to calculate solar irradiance from count-rates measured on each of ESP's photometer bands, measurement conditions ({\\it i.e.} ESP temperature, dark current, band response, {\\it etc}.), ESP calibration data and orbit parameters. Section 5 shows results from ESP ground tests. Section 6 summarizes ESP pre-flight calibration results. A comparison of ESP solar irradiance measurements from the sounding rocket flight of 14 April 2008 with data from other EUV channels (EVE/MEGS sounding rocket instrument and SOHO/SEM) are given in Section 7. Concluding remarks are given in Section 8. ", "conclusions": "\\label{S-Conclusions} As part of the SDO/EVE suite of channels, ESP will provide highly stable and accurate measurements of absolute solar irradiance in five EUV wavelength bands. Its high temporal cadence, low latency, and spectral overlap with the other SDO instruments, will provide important details of the dynamics of rapidly changing irradiance related to impulsive phases of solar flares. An algorithm to convert measured count rates and other ESP data into solar irradiance was described. ESP has many significant design improvements over its SOHO/SEM predecessor, including the possibility of measuring on orbit changes of dark currents, electronics gain changes, contamination degradation and pin-holes of thin-film filters, visible light scatter, and energetic particles background. All optical components of ESP (diodes, filters, grating) were also separately tested and calibrated prior to an end-to-end calibration. The ESP was calibrated at SURF BL-9 using quasi-monochromatic wavelengths, and at BL-2 to obtain a radiometric calibration. Efficiency profiles determined during BL-9 calibration were used as reference data for calculation of irradiance in each band during BL-2 calibration. Measurements of solar irradiance from ESPR during the EVE sounding rocket flight of 14 April 2008 were compared with corresponding EVE/MEGS spectra and SOHO/SEM irradiance measurements. These comparisons show good agreement. Both the ESP flight and rocket instruments are fully calibrated and ready for solar measurements. %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \\begin{acks} The authors thank Frank Eparvier, Mike Anfinson, Rick Kohnert, Greg Ucker, Don Woodraska, Karen Bryant, Gail Tate, Matt Triplett, and the rest of the LASP EVE Team at the University of Colorado at Boulder for their many contributions during ground tests of the ESP. We also thank Don McMullin of the Space Systems Research Corporation for his discussions and long-term support of this mission, and the NIST SURF Team: Rob Vest, Mitch Furst, Alex Farrell, and Charles Tarrio for support of ESP calibration at BL-9 and BL-2 and diffraction grating transmission measurements at the SURF reflectometer facility. This work was supported by the University of Colorado award 153-5979. \\end{acks} %%% BIBLIOGRAPHY %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%" }, "0911/0911.4448_arXiv.txt": { "abstract": "A digital frequency multiplexing (DfMUX) system has been developed and used to tune large arrays of transition edge sensor (TES) bolometers read out with SQUID arrays for mm-wavelength cosmology telescopes. The DfMUX system multiplexes the input bias voltages and output currents for several bolometers on a single set of cryogenic wires. Multiplexing reduces the heat load on the camera's sub-Kelvin cryogenic detector stage. In this paper we describe the algorithms and software used to set up and optimize the operation of the bolometric camera. The algorithms are implemented on soft processors embedded within FPGA devices operating on each backend readout board. The result is a fully parallelized implementation for which the setup time is independent of the array size. ", "introduction": "\\label{sec-introduction} A new generation of mm-wavelength experiments use (or will use) arrays of hundreds or thousands of transition edge sensor (TES) bolometers operating at sub-Kelvin temperatures to image the cosmic microwave background (CMB) to unprecedented precision in the search for new galaxy clusters or the polarization imprint left by inflation. One key element in the cryogenic design for these experiments is the reduction of readout wiring, which is accomplished for APEX-SZ~\\cite{2008JLTP..151..697M}, the South Pole Telescope (SPT)~\\cite{2004astro.ph.11122R}, EBEX \\cite{2008SPIE.7020E..68G} and Polarbear~\\cite{2008AIPC.1040...66L} by multiplexing the bolometer sky-signals in the frequency domain. In this paper we describe tuning algorithms and software developed for the new digital backend electronics~\\cite{2008ITNS...55...21D} in use for EBEX and POLARBEAR. This system is the evolution of the analog backend~\\cite{2005ApPhL..86k2511L} deployed on APEX-SZ and the SPT. The new system provides an order of magnitude reduction in power consumption and size, improvements in low-frequency noise performance, and a faster, more robust setup. As most of the system complexity is encapsulated in reprogrammable firmware, it is easily adapted to the requirements of specific experiments. \\begin{figure}[htbp] \\includegraphics[width=7cm]{schematic} \\caption{Schematic diagram showing $N$ bolometers multiplexed through a single set of wires and SQUID in the Digital frequency domain multiplexer system.} \\label{fig-schematic} \\end{figure} Figure \\ref{fig-schematic} shows a circuit with $N$-bolometers multiplexed on a single set of wires and read out through a SQUID operating in shunt feedback~\\cite{2008ITNS...55...21D, 2005ApPhL..86k2511L}. The box on the left labelled `DfMUX' (for digital frequency multiplexer') represents firmware implemented on a Field Programmable Gate Array (FPGA) that performs the signal processing for the system digitally. A digital multifrequency synthesizer (DMFS) produces a sine-wave bias carrier for each channel and sums them to form a ``comb'' in frequency space. Each bolometer in the multiplexer module is in series with an inductor-capacitor (LC) filter that selects a single bias that is uniquely positioned in frequency-space. Intensity variations from the sky-signal change the bolometer resistance, amplitude modulating the bolometer current, and appearing as side-bands adjacent to the carriers. These currents are summed at the input of the SQUID, which operates as a transimpedance amplifier. To avoid flux-burdening the SQUID with the large bias carriers, an inverted copy of the carrier ``comb,'' called the nuller, is injected at the SQUID input from a second DMFS. The sky-signal modulated comb is digitized after the SQUID circuit and each bolometer signal is separately demodulated in the digital multifrequency demodulator (DMFD), implemented in firmware. Each detector channel has its own demodulator, which provides only the in-phase ``I'' component. Including a second demodulator for each channel doubles the usage of FPGA fabric, increasing the power consumption, and does not provide any useful additional information. To align the demodulator phase with the carrier, accounting for phase shifts in the cryostat, each module has an extra $N^{th}+1$ demodulator that can be used to momentarily measure the out-of-phase ``Q'' component of each detector. In addition to aligning the demodulator with the carrier, this ``helper'' demodulator is employed to provide extra information in the algorithms described below. Frequency domain multiplexing permits the reduction of cryogenic wires from two per bolometer to two per $N$ bolometers, where $N$ is the multiplexing factor. The system is currently implemented with $N=8$ and $N=16$. Extension to much higher multiplexing factors is feasible but requires a substantial improvement in SQUID bandwidth, using methods such as those demonstrated in Ref.~\\cite{2009arXiv0901.1919L}. ", "conclusions": "The algorithms used to setup and tune large arrays of TES bolometers for observations of the Cosmic Microwave Background radiation with mm-wavelength experiments using a digital frequency domain SQUID multiplexer system are described. The system provides substantial performance and practical improvements over previous generation analog systems. This work is supported by the Natural Sciences and Engineering Research Council of Canada, Canadian Institute for Advanced Research, Canada Research Chairs program, a NASA GSRP grant, a Minnesota Graduate School Dissertation Fellowship, and the Canadian Foundation for Innovation. We thank Xilinx Canada for their contribution of the FPGA devices." }, "0911/0911.4481_arXiv.txt": { "abstract": "Massive black hole (BH) mergers can result in the merger remnant receiving a ``kick,'' of order 200 km~s$^{-1}$ or more, which will cause the remnant to oscillate about the galaxy centre. Here we analyze the case where the BH oscillates through the galaxy centre perpendicular or parallel to the plane of the galaxy for a model galaxy consisting of an exponential disk, a Plummer model bulge, and an isothermal dark matter halo. For the perpendicular motion we find that there is a strong resonant forcing of the disk radial motion near but somewhat less than the ``resonant radii'' $r_R$ where the BH oscillation frequency is equal one-half, one-fourth, ($1/6$, etc.) of the radial epicyclic frequency in the plane of the disk. Near the resonant radii there can be a strong enhancement of the radial flow and disk density which can lead to shock formation. In turn the shock may trigger the formation of a ring of stars near $r_R$. As an example, for a BH mass of $10^8~M_\\odot$ and a kick velocity of $150$ km s$^{-1}$, we find that the resonant radii lie between $0.2$ and $1$ kpc. For BH motion parallel to the plane of the galaxy we find that the BH leaves behind it a supersonic wake where star formation may be triggered. The shape of the wake is calculated as well as the slow-down time of the BH. The differential rotation of the disk stretches the wake into ring-like segments. ", "introduction": "Recent breakthroughs in numerical General Relativity have led to predictions of large recoil velocities of merged binary supermassive black-holes (BH) as the binary radiates away linear momentum as gravitational waves during the final stages of merger (Gonz\\'alez et al. 2007; Campanelli et al. 2007; Lousto, Campanelli, \\& Zlochower 2009). Typical kick velocities are of order $\\sim 200$~km~s$^{-1}$ (Bogdanovi\\'c, Reynolds, \\& Coleman 2007). One possible result of these mergers is to cause the resulting remnant black hole (BH) to oscillate through the stellar disk on timescales on the order of Gyr (Kornreich \\& Lovelace 2008, hereafter KL08; Blecha \\& Loeb 2008; Fujita 2009). If the merged BH receives an impulse from the merger, some of that motion will be transmitted to the galaxy by dynamical friction. It is clearly of interest to know whether the motion of the BH through the disk generates observable changes in morphology and dynamics in the galaxy. Binary black-holes have already been observed as a double nucleus in a quasar (e.g., Decarli et al. 2009). Further, Comerford et al. (2009) show that BH binaries resulting from recent mergers are observable in the spectra of as many as $40\\%$ of Seyfert 2 galaxies. The possible observable effects of free and ``wandering'' BH merger remnants on disk galaxies, however, have not yet been fully analyzed. de la Fuente Marcos \\& de la Fuente Marcos (2008) discuss the formation of stars in the wakes of runaway black holes ejected from the host galaxy. Here we consider more slowly moving BHs which remain bound to the host galaxy. This work first analyzes the case where the ejected BH oscillates with angular frequency $\\Omega_{bhz}$ through the galaxy center in a direction normal to the disk. Later, we discuss the case where the BH is ejected parallel to the plane of the galaxy. Sufficiently large gas accretion by the binary BH system is predicted to drive the orbital and BH spins into alignment normal to the plane of the galaxy (Bogdanovi\\'c et al. 2007). However, the amount is uncertain due to the uncertainty in the ratio of the two viscosity coefficients for the $(r,z)$ and $(r,\\phi)$ motion in the disk (e.g., Natarajan \\& Armitage 1999). This alignment favors BH ejection in the plane of the galaxy (Campanelli et al. 2007). The merging process preceding the ejection is assumed to be sufficiently slow that the host galaxy has returned to an axisymmetric equilibrium state. The BH oscillations are observed in simulations to persist for many periods (KL08; Blecha \\& Loeb 2008; Fujita 2009). Under these conditions we find that there is a strong resonant forcing of the disk radial motion near radii $r_R$ where $2n\\Omega_{bhz}= \\Omega_r(r_R)$, with $n=1,~2,..$, where $\\Omega_r$ is the radial epicyclic frequency. Near $r_R$ there can be a strong enhancement of the density in a ring of the disk and shock formation which may trigger the formation of a ring of stars. We note that models for the formation of observed ``ring galaxies'' assume the (single) passage of one galaxy nearly through the center of another (Theys \\& Spiegel 1976, 1977; Lynds \\& Toomre 1976). The gravitational interaction of the two galaxies causes the formation of a pronounced ring(s) of enhanced density (in one or both of the galaxies) where star formation is expected to be enhanced. For the case of BH ejection in the plane of the galaxy we find that the BH leaves behind it a supersonic wake where star formation may be triggered. The shape of this wake is calculated and the slow-down time of the BH is estimated. The differential rotation of the disk stretches the wake into ring-like segments. Section 2.1 describes the gravitational potential of the equilibrium disk galaxy at the time of the BH merger. This includes an exponential disk of stars and gas, a Plummer model the bulge, and an isothermal dark matter halo. Section 2.2 derives the dynamical equations of the system as driven by the radial force due to a BH oscillating perpendicular to the disk. The frequencies and amplitudes of the BH oscillations are taken from the $N$--body simulations of KL08, Section 2.3 discusses the forced oscillations of the gas disk, and \\S 2.4 the possible parametric instability of the disk. Section 3 discusses the response of a galaxy disk to a BH ejected in the plane of the galaxy. Section 4 gives the conclusions of this work. ", "conclusions": "Here, we first analyzed the linear axisymmetric perturbations of a gas disk driven by a black hole oscillating vertically along the axis of symmetry. We find that there is a strong resonant forcing of the disk radial motion near ``resonant radii'' $r_R$ where the BH oscillation frequency is equal one-half, one-fourth, ($1/6$, etc.) of the radial epicyclic frequency in the plane of the disk. Near the resonant radii there can be a strong enhancement of the radial flow velocity and disk density which can lead to shock formation. The shock formation occurs during one period of oscillation of the BH which is assumed longer than the BH damping time due to dynamical friction. This shock may trigger the formation of a ring of stars near $r_R$. As an example, for a BH mass of $10^8~M_\\odot$ and a kick velocity of $150$ km s$^{-1}$, we find that the resonant radii lie between $0.2$ and $1$ kpc. The magnitude of the disk response is proportional to the BH mass, inversely proportional to the Toomre~$Q$ of the gas disk, and decreases rapidly as $r_R$ increases. The resonant radii increase as the initial BH kick velocity increases (KL08). For BH motion parallel to the plane of the galaxy we find that the BH leaves behind it a supersonic wake which over time gets contorted into a complicated shape by the galaxy's differential rotation. The shape of the wake is calculated for an illustrative case as well as the slow-down time of the BH. The differential rotation of the disk stretches the wake into ring-like segments of radii approximately equal to the maximum excursion of the BH, $x_m$, where the BH most slowly. Many other processes may be involved in the formation and evolution of nuclear rings in galaxies (e.g., van de Ven \\& Chang 2009)." }, "0911/0911.1092_arXiv.txt": { "abstract": "Using the nuclear symmetry energy that has been recently constrained by the isospin diffusion data in intermediate-energy heavy ion collisions, we have studied the transition density and pressure at the inner edge of neutron star crusts, and they are found to be $0.040$ fm$^{-3}$ $\\leq \\rho _{t}\\leq 0.065$ fm$^{-3}$ and $0.01$ MeV/fm$^{3}$ $\\leq P_{t}\\leq 0.26$ MeV/fm$^{3}$, respectively, in both the dynamical and thermodynamical approaches. We have also found that the widely used parabolic approximation to the equation of state of asymmetric nuclear matter gives significantly higher values of core-crust transition density and pressure, especially for stiff symmetry energies. With these newly determined transition density and pressure, we have obtained an improved relation between the mass and radius of neutron stars. ", "introduction": "\\label{introduction} Studying the properties of neutron stars, which are among the most mysterious objects in the universe, allows us to test our knowledge of matter under extreme conditions. Theoretical studies have shown that neutron stars are expected to have a liquid core surrounded by a solid inner crust~\\cite{Cha08}, which extends outward to the neutron drip-out region. While the neutron drip-out density $\\rho _{\\rm out}$ is relatively well determined to be about $4\\times 10^{11}$ g/cm$^{3}$ or $0.00024~{\\rm fm}^{-3}$~\\cite{Rus06}, the transition density $\\rho _{t}$ at the inner edge is still largely uncertain mainly because of our very limited knowledge on the nuclear equation of state (EOS), especially the density dependence of the symmetry energy ($E_{\\rm sym}(\\rho)$) of neutron-rich nuclear matter~\\cite{Lat00,Lat07}. These uncertainties have hampered our understanding of many important properties of neutron stars~\\cite{Lat00,Lat07,Lat04} and related astrophysical observations~\\cite{Lat00,Lat07,BPS71,BBP71,Pet95a,Pet95b,Ste05,Lin99,Hor04,Bur06,Owe05}. Recently, significant progress has been made in constraining the EOS of neutron-rich nuclear matter using terrestrial laboratory experiments (See Ref.~\\cite{LCK08} for the most recent review). In particular, the analysis of the isospin-diffusion data \\cite{Tsa04,Che05a,LiBA05c} in heavy-ion collisions has constrained tightly the $E_{\\rm sym}(\\rho)$ in exactly the same sub-saturation density region around the expected inner edge of neutron star crust. The extracted slope parameter $L=3\\rho_0\\frac{\\partial E_{\\rm sym}(\\rho)}{\\partial\\rho}|_{\\rho=\\rho_0}$ in the density dependence of the nuclear symmetry energy was found to be $86\\pm25$ MeV, which has further been confirmed by a more recent analysis~\\cite{Tsa09}. With this constrained nuclear symmetry energy, we have obtained an improved determination of the values for the transition density and pressure at the inner edge of neutron star crusts. This has led us to obtain significantly different values for the radius of the Vela pulsar from those estimated previously. Also, we have found that the widely used parabolic approximation (PA) to the EOS of asymmetric nuclear matter gives much larger values for the transition density and pressure, especially for stiff symmetry energies. ", "conclusions": "" }, "0911/0911.4354_arXiv.txt": { "abstract": "Using the results of extensive Monte Carlo simulations we discuss corrections to the linear mixing rule in strongly coupled binary ionic mixtures. We analyze the plasma screening function at zero separation, $H_\\mathit{jk}(0)$, for two ions (of types $j=1,2$ and $k$=1,2) in a strongly coupled binary mixture. The function $H_\\mathit{jk}(0)$ is estimated by two methods: (1) from the difference of Helmholtz Coulomb free energies at large and zero separations; (2) by fitting the Widom expansion of $H_\\mathit{jk}(x)$ in powers of interionic distance $x$ to Monte Carlo data on the radial pair distribution function $g_{jk}(x)$. These methods are shown to be in good agreement. For illustration, we analyze the plasma screening enhancement of nuclear burning rates in dense stellar matter. ", "introduction": "More than 30 years ago \\cite{HV76} the linear mixing rule for multicomponent strongly coupled mixtures was shown to be highly accurate. However, only recent studies \\cite{PCR09,Mixt_New} have achieved enough accuracy to describe the corrections to the linear mixing rule for a wide range of plasma parameters; previous attempts, e.g.\\ \\cite{DWSC96,DWS03}, were restricted at least by a limited number of data points. We discuss the corrections to the linear mixing rule in application to the plasma screening of nuclear reactions in strongly coupled mixtures. Following Ref.\\ \\cite{DWS99} we apply two approaches to calculate the screening enhancement: one is based on the thermodynamic relations and the other on fitting the mean-field potentials. The main advance of the present work is in using a much wider set of numerical data and most precise thermodynamic results. ", "conclusions": "\\label{Sec:approx} We suggest to use the following approximation for the enhancement factor for all $\\Gamma$ and mixture composition \\begin{equation} h^0_{jk}=h^\\mathrm{lin}_{jk}/ \\left[1 +C_{jk}\\left(1-C_{jk}\\right)\\, \\left(h^\\mathrm{lin}_{jk}/h^\\mathrm{DH}_{jk} \\right)^2 \\right]. \\label{appr} \\end{equation} Here, $h^\\mathrm{lin}_{jk}$ is given by (\\ref{h0lin}), $ h^\\mathrm{DH}_{jk}=3^{1/2} Z_j\\,Z_k\\ZZmid^{1/2}\\Gamma_\\mathrm{e}^{3/2}/\\Zmid^{1/2}$ is the well known Debye-H\\\"{u}ckel enhancement parameter, and % $C_{jk}= 3Z_j\\,Z_k\\ZZmid^{1/2}\\Zmid^{-1/2}/\\left[\\left(Z_j+Z_k\\right)^{5/2}-Z_j^{5/2}-Z_k^{5/2}\\right] $. Eq.\\ (\\ref{appr}) reproduces the Debye-H\\\"{u}ckel asymptote at low $\\Gamma$ and the linear mixing at strong coupling. \\begin{figure} \\begin{center} \\leavevmode \\epsfxsize=150mm \\epsfbox{ChugunovAI_fig2.eps} \\end{center} \\caption{(Color online) Enhancement factors $h_{jk}^0/\\Gamma_{11}^{3/2}$ vs $\\Gamma_{11}$ for three binary ionic mixtures. } \\label{Fig_hvsG} \\end{figure} In Fig.\\ \\ref{Fig_hvsG} we show the dependence of the approximated enhancement factors $h_{jk}^0/\\Gamma_{11}^{3/2}$ on $\\Gamma_{11}$. The figure contains three panels; each for a specific binary ionic mixture. Each panel shows three groups of four lines. They are (from top to bottom) $h^0_{22}/\\Gamma_1^{3/2}$, $h^0_{12}/\\Gamma_1^{3/2}$ and $h_{11}^0/\\Gamma_1^{3/2}$. Two of any four lines (solid and thick dashed lines) are almost the same in the majority of cases. This couple represents the approximation (\\ref{appr}) and the thermodynamic enhancement factor (\\ref{h0}), respectively. The dotted horizontal lines refer to the Debye-H\\\"{u}ckel model and the dash-dot lines are the linear mixing results. One can see that our approximation is in a good agreement with thermodynamic results for most of cases, especially in panel (a) (for all mixtures with not too large $Z_2/Z_1$). If $Z_2/Z_1$ becomes too large [panel (c)], the thermodynamic model of $h_{11}^0/\\Gamma_1$, calculated in accordance with \\cite{Mixt_New}, has a specific feature ($h_{11}/\\Gamma_{11}^{3/2}$ increases at $\\Gamma_{11}\\sim10^{-2}$), while our approximation has not. We expect that this feature is not real, but results from not too accurate extractions of the enhancement factors from thermodynamic data. The free energy is almost fully determined by larger charges $Z_2$ which also dominate by number ($99\\%$) in panel (c). Using Eq.\\ \\ref{h0} to get $h_{11}^0$, one should differentiate the free energy with respect to $N_1$, which provides vanishing contribution to the free energy. Hence this procedure is very delicate and can strongly amplify the errors of original thermodynamic approximation. We expect that our approximation can be more accurate than the original thermodynamic result. Another, less probable option is that we still have not enough data to prove the presence of the feature of $h^0_{11}/\\Gamma_{11}^{3/2}$. To conclude, we have calculated the enhancement factors of nuclear reactions in binary ionic mixtures by two methods and showed good agreement of the results. We have proposed a simple approximation of the enhancement factors valid for any Coulomb coupling. This approximation is almost the same as thermodynamic ones for not too specific mixtures. It does not confirm some questionable features of the enhancement factors for mixtures with large $Z_2/Z_1$ and small $x_1$. \\begin{acknowledgement} We are grateful to D.G.~Yakovlev and A.Y.~Potekhin for useful remarks. Work of AIC was partly supported by the Russian Foundation for Basic Research (grant 08-02-00837), and by the State Program ``Leading Scientific Schools of Russian Federation'' (grant NSh 2600.2008.2). Work of HED was performed under the auspices of the US Department of Energy by the Lawrence Livermore National Laboratory under contract number W-7405-ENG-48. \\end{acknowledgement}" }, "0911/0911.0161_arXiv.txt": { "abstract": "Assuming that density waves trigger star formation, and that young stars preserve the velocity components of the molecular gas where they are born, we analyze the effects that non-circular gas orbits have on color gradients across spiral arms. We try two approaches, one involving semi-analytical solutions for spiral shocks, and another with magnetohydrodynamic (MHD) numerical simulation data. We find that, if non-circular motions are ignored, the comparison between observed color gradients and stellar population synthesis models would in principle yield pattern speed values that are systematically too high for regions inside corotation, with the difference between the real and the measured pattern speeds increasing with decreasing radius. On the other hand, image processing and pixel averaging result in systematically lower measured spiral pattern speed values, regardless of the kinematics of stellar orbits. The net effect is that roughly the correct pattern speeds are recovered, although the trend of higher measured $\\Omega_p$ at lower radii (as expected when non-circular motions exist but are neglected) should still be observed. We examine the \\citet{mar09} photometric data and confirm that this is indeed the case. The comparison of the size of the systematic pattern speed offset in the data with the predictions of the semi-analytical and MHD models corroborates that spirals are more likely to end at Outer Lindblad Resonance, as these authors had already found. ", "introduction": "Until the detection of an azimuthal color gradient across one of the arms of the SAc galaxy M~99 \\citep[][GG96 hereafter]{gon96}, only sparse evidence of star formation triggered by spiral density waves had been found in stellar counts in the Milky Way\\citep{sit89,sit91,ave89} and M~31\\citep{efr80a,efr80b,efr85}. More recently, applying the same method defined by GG96 and described below, \\citet[][MG09 hereafter]{mar09} examined a sample of 13 spiral galaxies of types A and AB, and found color gradients consistent with theoretical expectations in 10 of their objects.\\footnote{See also \\citet{gro09} for an infrared study of young stellar complexes in NGC~2997.} Although they did not compute stellar orbits, both GG96 and MG09 implicitly assumed that all stars, including those recently born near spiral arms, move in circular orbits. MG09 did investigate the effects of variable circular speeds and variable densities on observed color gradients and found them to be negligible. However, complicated non-circular motions have been reported in studies about the migration of young stars following star formation triggered by spiral shocks \\citep{yua69,wie79,fer08}. \\citet{bash81} had also noticed that the ballistic orbits calculated from galactic HII region complexes follow non-circular trajectories that initially move along (and not across) the spiral arms. In the present work, we examine how the presence of non-circular motions would modify the pattern speeds derived from color gradients under the assumption of circular orbits. We will explore two different approaches. The first one (see \\S~\\ref{appr1}) involves the semi-anaylitical solutions obtained for spiral shocks \\citep{rob69,shu73,gitt04}; we assume that newly born stars preserve the orbital motion of the shocked gas where they form. The second approach (see \\S~\\ref{appr2}) is based on data from MHD simulations; we follow the gas flow vectors near spiral arms. We always assume that stars are triggered all along the studied regions of spiral arms in the respective model. Other methods, not discussed here, may involve orbit calculations for young stellar groups. ", "conclusions": "Under the assumption that the orbits of young stars preserve the velocity components of the parent molecular clouds where they form, we have analyzed the effects that non-circular motions would have on azimuthal color gradients. Semi-analytical calculations and MHD simulations show that the spiral pattern speeds derived from the comparison between color gradients and stellar population synthesis models, assuming purely circular motions, would have values systematically higher than the real ones for regions within corotation. Also inside corotation, the effect would decrease with galactocentric radius. Non-circular motions, however, would not prevent the detection of legitimate azimuthal color gradients in real galaxies. Using a synthetic image, we have also analyzed the effects of image processing and pixel averaging on the method applied in GG06 and MG09 to detect color gradients and derive pattern speeds. We have found that pixel averaging (due to image processing) systematically decreases the derived $\\Omega_p$, such that it nearly compensates for the systematic effect introduced when neglecting non-circular motions in the analysis. The net result is that the correct spiral pattern speeds and resonance locations can be obtained. Nevertheless, a residual trend of slightly higher pattern speeds at lower radii can still be discerned ({\\it solid} triangles in Fig.~\\ref{myOMEGAS} and Fig.~\\ref{diffOM_OLR}), so that it is possible to detect the presence of non-circular motions and confirm the link between star formation and disk dynamics. We have re-examined the results obtained by MG09. When normalizing the mean radii where gradients were found by the end point radii, in order to treat the whole sample as a single galaxy, we were able to reproduce the trend of $\\delta \\Omega_{p}$ with radius expected if non-circular motions are ignored (as these authors did). The size of the observed $\\delta \\Omega_{p}$ in the data only matches the theoretical expectations if spiral patterns end at the OLR, and not at the 4:1 resonance (cf.\\ Fig.~\\ref{myOMEGAS} with both Fig.~\\ref{diffOM_4_1} and Fig.~\\ref{diffOM_OLR}). Future studies of azimuthal color gradients in disk galaxies by orbit calculations should include the effects of disk heating \\citep{asi99}, and a correct modeling of the IMF as a discrete distribution. These factors may be important, since high mass stars may take different trajectories when compared to low mass stars, due to the effects of dynamical friction~\\citep{chandra43}." }, "0911/0911.5091_arXiv.txt": { "abstract": "{The dynamics of prominence fine structures is a challenge to understand the formation of cool plasma prominence embedded in the hot corona.} {Recent observations from the high resolution \\emph{Hinode}/SOT telescope allow us to compute velocities perpendicularly to the line-of-sight or transverse velocities. Combining simultaneous observations obtained in H$\\alpha$ with \\emph{Hinode}/SOT and the MSDP spectrograph operating in the Meudon solar tower we derive the velocity vectors of a quiescent prominence.} {The velocities perpendicular to the line-of-sight are measured by time slice technique, the Dopplershifts by the bisector method.} {The Dopplershifts of bright threads derived from the MSDP reach 15 km s$^{-1}$ at the edges of the prominence and are between $\\pm$ 5 km s$^{-1}$ in the center of the prominence. Even though they are minimum values due to seeing effect, they are of the same order as the transverse velocities.} {These measurements are very important because they suggest that the vertical structures shown in SOT may not be real vertical magnetic structures in the sky plane. The vertical structures could be a pile up of dips in more or less horizontal magnetic field lines in a 3D perspective, as it was proposed by many MHD modelers. In our analysis we also calibrate the \\emph{Hinode} H$\\alpha$ data using MSDP observations obtained simultaneously.} ", "introduction": "The existence of cool structures in so called prominences and filaments during a few solar rotations embedded in the hot corona has been a mystery since the beginning of their spectrographic observations \\citep{Azambuja48}. Many reviews concerned the study of quiescent prominences \\citep{Schmieder89,Tandberg94,Labrosse09,Mackay09}. Since that time period, it is a challenge to derive what is the best mechanism to succeed to maintain cool plasma in the corona. A very popular idea is that the plasma is frozen in magnetic field lines and stays cool due to the low transverse thermal conduction \\citep{Demoulin89}. Many magnetic and static models have been developed on this idea \\citep{Kuperus74,Kippenhahn57,Aulanier98,Dudik08}. But a big question remains: how to get cool plasma inside the field lines into the corona and how to keep it there? It is recognized that this material should come from the chromosphere by levitation or by injection \\citep{Saito73}. Sufficient mass should be extracted from the chromosphere by magnetic forces which inject or lift the plasma or by pressure forces which evaporate the plasma and then cool it to prominence temperatures. Many models have been developed in this sense, i.e. thermal non-equilibrium models \\citep{Mariska85,Karpen03,Karpen05}. Levitation models are proposed through possible magnetic reconnection \\citep{Ballegooijen89}. Injection models could be due to injection through the reconnection of magnetic field during canceling flux. These models indicate that the plasma in prominences should have a large dynamics and static models may be obsolete. Recent observations at the Swedish Solar Telescope (SST) show highly dynamic plasma in filament threads \\citep{Lin03,Lin05}. It was also a first attempt to compute the velocity vectors of the filament threads. They conclude that the threads inclination from the horizontal were around 16 degrees with a net flow in both directions of 8 km s$^{-1}$. Fine counterstreaming flow is often observed along horizontal threads or in the barbs \\citep{Zirker98,Schmieder91,Schmieder08}. \\emph{Hinode}/SOT movies (available with the electronic version of the paper) reveal strong dynamics in the prominence fine structures. The spicules close to the barbs could allow to inject plasma inside the fine threads. Is that sufficient to feed continuously the main core of the prominence? A mass budget should be done. Using \\emph{Hinode} observations at the limb \\citet{Berger08} and \\citet{Chae08} tried to answer these questions. They reported different velocity measurements. \\citet{Berger08} found upflows of dark bubbles around 20 km s$^{-1}$ and down flows of bright knots less than 10 km s$^{-1}$ looking at Ca II H images in the line center. \\citet{Chae08} analysed an hedgerow prominence and found horizontal displacements before observing downflows of bright knots suggesting the existence of vortex motions. \\begin{figure} \\centering \\hspace*{-1.5cm} \\includegraphics[width=0.60\\textwidth,clip=]{fig1.ps} \\caption{(a) EIT 304 \\AA\\ image observed on April 20, 2007 at 01:00 UT, the field of view is 1100$\\times$750 arcsec, (b) Filament observed in H$\\alpha$ at Meudon on April 22, 2007 at 15:28 UT, (c) MDI longitudinal magnetic field on April 21 at 08:00 UT. The magnetic inversion line is represented by the dashed line in the middle of the EUV filament channel and in MDI image. The letters F1 and F2 indicate approximately the location of the two filament fragments visible in H$\\alpha$. }\\label{meudon} \\end{figure} Prominence dynamics at the limb look quite different from filament dynamics on the disk. The integration along the line of sight complicates the interpretation. \\citet{Mein91} show that more than 15 threads may be integrated along the line of sight and the resulting velocities should depend on two different Gaussian distributions. At the edges of prominences the velocity values are higher because fewer threads are integrated and velocity cells are larger than intensity knots revealing that {\\bf bunches} of threads may move with the same velocity plasma (Dopplershifts). Similar results have been found from completely different approaches. Therefore to reproduce Lyman line profiles observed by SOHO/SUMER, \\citet{Gunar07} introduce 10 threads perpendicularly to the line-of-sight with a random velocity distribution in a 2D non LTE radiative transfer code. We proposed in this study to study both the velocity perpendicular to the line of sight using one hour of observations of \\emph{Hinode}/SOT in H$\\alpha$ combined with Dopplershifts observed also in H$\\alpha$ with the Multichannel Subtractive Double Pass spectrograph (MSDP) operating in the Meudon solar tower. These observations were obtained simultaneously during a coordinated observing program (JOP178). It is the first time that using \\emph{Hinode} data such fine structures are resolved in prominences and that oscillations and transverse velocities can be derived \\citep{Okamoto07,Berger08,Chae08}. \\emph{Hinode} has a much better spatial resolution than MSDP by a factor of 5, but the MSDP observations are very useful to calculate the Dopplershifts and to calibrate the intensity of \\emph{Hinode}/SOT observations. ", "conclusions": "For the first time an H$\\alpha$ hedgerow prominence has been observed simultaneously by a high spatial resolution telescope (\\emph{Hinode}/SOT) and by a spectrograph (MSDP) operating in the Meudon solar tower on April 25 2007. \\emph{Hinode}/SOT allows us to get the velocities perpendicular to the line-of-sight V(x,z). The second -- the line-of-sight velocities V(y) derived from Dopplershifts. The prominence shows strong dynamics in the SOT movie with dark cavities rising from the limb with an upward velocity reaching 24 km s$^{-1}$ and downflowing vertical-like bright threads. These threads are moving horizontally to avoid the dark cavities. During the rise of cavities, ahead of them, are observed bright curved fine structures from time to time with high velocities similar to the speed rise of the bubble. The \\emph{Hinode}/SOT observations have been calibrated by using the MSDP data. The integrated H$\\alpha$ intensity of the threads reaches 1.5 $\\times 10^5$ erg/s/sr/cm$^2$. The contrast of the dark cavities is between 70 and 90$\\%$. \\begin{figure} \\vspace*{-1cm} \\centering \\hspace*{-2cm} \\includegraphics[width=0.60\\textwidth,clip=]{fig7.ps} \\vspace*{-1cm} \\caption{Transverse velocities in SOT bright structures using time slice technique (axis x unit is time, axis y unit is arc sec along the slide). Top/medium/bottom frame corresponds to slice A/ D/ E, drawn in Fig. \\ref{doppler}. They show the velocities measured respectively in points A1, A2, D1, D2, E1, E2 (Table 1). Positive/negative velocities correspond to up/down flows. The large value +24 km s$^{-1}$ corresponds to the speed of a rising bubble from the limb or the flow speed of its bright edge. Fine threads close to the limb are spicules. Wave pattern corresponds to oscillations with 15 to 20 min of period. Adjacent pixels in a slice have coherent velocities. } \\label{sot_velocity} \\end{figure} The transverse velocities V(x,z) of the bright threads are computed by time slice technique and these values are of the order of a few km s$^{-1}$ to 6 km s$^{-1}$ reaching 11 km s$^{-1}$ for individual fine threads. The pattern of the Dopplershift map show elongated cells nearly perpendicular to the limb. They are {\\bf wider than the MSDP spatial resolution}. The time slice maps exhibit several pixels close to each other along the slice having a similar velocity trend. This means that fine threads close to each other have a coherent displacement. According to the observations of the prominence three days before when it is still on the disk as a filament, it appears that only the feet or barbs are enough dense to be observed. The prominence would represent the barb threads integrated along the line-of-sight as the filament is crossing the limb. The structures are not vertical in the sky plane (x,z) as suggested by the movie. Dopplershifts and transverse velocities are of same order of magnitude (less than 6 km s$^{-1}$). In Table 1 we have selected individual threads with the largest transverse velocities in region with higher Dopplershifts. The other parts of the prominence exhibit coherent velocity much smaller (1 to 2 km s$^{-1}$) difficult to measure. The narrow profiles of H$\\alpha$ lines of the prominence indicate that the different threads integrated along the line of sight have similar velocities. The dispersion of the velocities along the line-of-sight is small. The longitudinal magnetic field observed (by the SOHO/MDI instrument) in the filament channel on the disk and on both edges of the inversion line is weak. The strength of the small polarities are less than 10 Gauss. The prominence lies in a quiet region and corresponds to a quiescent filament. The small polarities can change rapidly and this would explain the fast dynamics of the structures. In a flux tube model \\citep{Aulanier98,Dudik08} the H$\\alpha$ filament is considered as cool material trapped in shallow dips along long magnetic field lines. The feet are extension of the flux tube disturbed laterally by parasitic polarities. The barbs are piled up dips touching the photosphere. When a parasitic polarity cancels or moves, the feet move and can even disappear \\citep{Aulanier98b,Schmieder06,Gosain09}. {\\bf In this 3D perspective the prominence material would be trapped in inclined field lines and the downflow motion would occur along the shallow dips. The brightness would result of the integration of the threads along the line-of-sight \\citep[see Fig. 3e in ][]{Dudik08}.} \\citet{Aulanier98b} explain very well through their magnetic extrapolation the relationship between parasitic polarities and the flux tube itself and their evolution. The flux of parasitic polarity overcomes that of the twisted flux tube and destroys the twisted configuration. The bubbles would be structures more magnetized that the surrounding and represented by the separatrices. A small increase of magnetic pressure in the bubble would lead to the rise of it in the atmosphere. Strong currents can be created in the quasi separatrice layers (QSL) around the separatrices by photospheric displacements of the parasitic polarities \\citep{Demoulin96}. Energy release is expected. This could correspond to the brightening {\\bf rims} associated with filaments \\citep{Heinzel95}. They are not systematically visible due to the dense plasma of filaments. In SOT observations it could correspond to the bright top edge of the cavities where reconnection could occur and expel plasma. This would explain the fast velocity material in brighter threads surrounding the dark cavities. On the next days, such bubbles are not observed in the prominence because the feet and parasitic polarities related to it would be on the back side of the disk. {\\bf In the arcade model with dips proposed by \\citet{Heinzel01} the prominence observed on April 25 may would consist of vertical threads trapped in dips and piled up giving the impression of vertical continuous threads. The downflows of 1 to 5 km s$^{-1}$ would be due to shrinkage or successive reconnections of field lines.} Another explanation for the buoyancy of the dark cavities or bubbles could be adiabatic expansion of a heated volume of plasma \\citep{Berger08}. This is not exclusive of the magnetic pressure increase scenario and both magnetic and thermal buoyancy may play a role in the formation of these structures. We would like to measure the magnetic field in the bubbles and in the prominence. An interesting aspect would be also to analyse the EIS and SUMER data to see if the dark bubbles are filled with hot material. Such measurements are needed to know which mechanism is valid for the formation of these dark low cavities." }, "0911/0911.0999_arXiv.txt": { "abstract": "We have used archival RXTE PCA data to investigate timing and spectral characteristics of the transient XTE J1817-330. The data pertains to 160 PCA pointed observations made during the outburst period 2006, January 27 to August 2. A detailed analysis of Quasi-Periodic Oscillations (QPOs) in this black hole X-ray binary is carried out. Power density spectra were obtained using the light curves of the source. QPOs have been detected in the 2-8 keV band in 10 of the observations. In 8 of these observations, QPOs are present in the 8-14 keV and in 5 observations in the 15-25 keV band. XTE J1817-330 is the third black hole source from which the low frequency QPOs are clearly detected in hard X-rays. The QPO frequency lies in $\\approx$ 4-9 Hz and the rms amplitude in 1.7-13.3\\% range, the amplitude being higher at higher energies. We have fitted the PDS of the observations with Lorentzian and power law models. Energy spectra are derived for those observations in which the QPOs are detected to investigate any dependence of the QPO characteristic on the spectral parameters. These spectra are well fitted with a two component model that includes the disk black body component and a power law component. The QPO characteristics and their variations are discussed and its implication on the origin of the QPOs are examined. ", "introduction": "\\label{sec:intro} Accretion powered X-ray binaries are the brightest X-ray sources in our Galaxy. These binaries contain either an accreting neutron star or an accreting black hole as the X-ray source. Based on estimates of the mass of the accreting object and its X-ray characteristics, about 40 X-ray sources have been classified as black hole binaries (Remillard and McClintock 2006a). Of these 20 have reliable mass estimates and are, therefore, regarded as confirmed black holes while the remaining 20 are considered to be black hole candidates (Remillard and McClintock 2006a). A majority of the black hole binaries (BHBs) are transients and most of them have a low mass optical companion.\\\\ The black hole X-ray binaries have several distinctive X-ray characteristics. During the outburst there is a strong soft component that originates in the inner region of the hot accretion disk. The presence of a hard X-ray component in energy spectra is another common feature of black hole binaries. The hard X-rays arise through Compton scattering of low energy photons from the accretion disk in a hot optically thin plasma. The resulting hard X-rays are referred to as the thermal Compton component (McClintock and Remillard 2006).\\\\ Low frequency quasi-periodic oscillations (QPOs) in $\\approx$ 0.1-40 Hz range and high frequency QPOs $\\approx$ 50-450 Hz also occur in many BHBs. Presence of several distinct spectral states and transition from one state to another at irregular intervals, is another distinguishing feature of BHBs. Spectral and temporal characteristics of the sources vary from one state to another. Remillard and McClintok (2006a) have broadly classified the spectral state for BHB as (a) High or Soft state (HS) dominated by the thermal component, (b) Low or Hard State (LH) marked by the hard X-ray power law and (c) Steep power law (SPL) or very high state characterized by steep slope power law component. For a more complete description of the spectral states McClintock and Remillard (2006) included two more states namely an Intermediate state (IM) that occurs when the source moves from LH to HS state and an extreme LH state which they called as quiescent state. Gierlinski, Done and Page (2008) have included an additional state termed as Ultra soft state (US) which is an extreme case of high/soft state found in this source. The US state is characterized by a very weak high energy tail with a low disk temperature and very low hardness ratio $\\leq$ 0.1 (Gierlinski, Done and Page 2008). The low frequency QPOs are usually detected in the SPL and LH states and have been observed in 14 BHBs so far (Remillard and McClintock 2006a). Detection of LFQPO in XTE J1817-330 indicates that LFQPOs can also be found sometime in the HS state.\\\\ An X-ray transient known as XTE J1817-330 was discovered by Remillard et al. (2006b) on 2006 January 26, with the All Sky Monitor (ASM) on Rossi Timing X-ray Explorer (RXTE) (Levine et al. 1996). At the time of its detection the 2-12 keV flux was 0.93 Crab. The intensity then rose to a peak value of 1.9 Crab on January 28, and then declined to 1.2 Crab by January 30. Subsequently it decayed exponentially with a decay time of 27 days (Sala et al. 2007). From its high/soft state at the time of the outburst, the source declined to a low/hard state characterized by kT = 0.2 keV. A change in the intensity state of the source occurred around $\\sim$ February 9 (Shaw et al. 2006). Hard X-rays (20-60 keV) flux increased from 46 $\\pm$ 2 mCrab on February 9 to 79 $\\pm$ 2 mCrab on February 14 accompanied by a decrease of soft X-ray (2-10 keV) flux from 840 $\\pm$ 4 mCrab to 670 $\\pm$ 4 mCrab during this period indicating the onset of the hard state (Kuulkers et al. 2006).\\\\ Its energy spectrum was studied with the RXTE, XMM-Newton, Integral and Swift instruments. Sala et al. (2007) measured the spectra using data from the XMM-Newton and Integral instrument when the source was in a high/soft state during 2006 February-March. Their spectral results indicated that its energy spectrum was typical of a BHB source with a dominant thermal disk component well described by kT $\\sim$ 0.7-0.9 keV and a thermal Compton power law component with a photon index of $\\sim$ 2-3 (Sala et al. 2007). Its spectral characteristics were observed with X-ray telescope (XRT) on the Swift satellite covering different stages of the outburst over 160 days. During this period the source made transition from the high/soft state in the initial outburst to a low hard state near the end of the outburst. The XRT spectra in the high soft state in 0.6-10 keV are well described by a two component model consisting of a thermal component from a optically thick and geometrically thin accretion disk and a hard power law component. The temperature of the inner disk producing the soft component declined from $\\sim$ 0.8 keV during the initial outburst phase to $\\sim$ 0.2 keV near the end of the outburst (Rykoff et al. 2007). A detailed study of the spectral evolution of XTE J1817-330 at different phases of the outburst was carried out by Gierlinski, Done and Page (2008) using the RXTE and the Swift data. The spectra of 150 PCU2 (RXTE) observations were modeled with the two component spectral model consisting of a disk component and a hard thermal Compton power law component. Using the same XRT data as used by Rykoff et al. (2007), they also found that the inner disc temperature declined from $\\sim$ 0.9 keV to $\\sim$ 0.2 keV as the source intensity declined. Based on this they claimed that the accretion disk recedes when the source transits from the high/soft state to the low/hard state. It may also be noted that apart from XTE J1118+480, this black hole binary has the lowest absorption along the line of sight among all the bright black hole candidates as obtained from the Chandra and Swift spectral data (Miller et al. 2006a;b). \\\\ Power density spectra of XTE J1817-330 obtained from the first two X-ray observations during 06:03-17:04 UTC on 2006 February 24, revealed strong QPOs at $\\sim$8.5 Hz (Homan, Miller and Wijnands 2006). Third observation (13:20-13:43 UTC) on the same day showed only a weak QPO at around 6.4 Hz. The last 2 observations (14:55-17:04 UTC) again indicated the presence of strong QPOs at 5.0 Hz.\\\\ Rupen, Dhawan and Mioduszewski (2006a), detected with VLA a radio object in the error box of the X-ray source having a flux density of 2.1 mJy at 1.4 GHz on 2006 January 31. The radio source faded away by 2006 February 2 (Rupen, Dhawan and Mioduszewski 2006b). A bright optical counterpart of the X-ray source was found at the time of the outburst with V = 11.3 magnitude and its brightness decreased to V = 15.5 by February 10 (Torres et al. 2006). The optical star was also detected in the near-infrared with a K magnitude of 15.0 on 2006 February 7 (D'Avanzo et al. 2006). Near-UV observations with the Optical Monitor on the XMM-Newton showed variations in the UV flux correlated to the hard X-ray flux variations (Sala et al. 2007). \\\\ All these characteristics strongly suggest that XTE J1817-330 is most likely a black hole binary. This transient was repeatedly observed with the Proportional Counter Array (PCA) on RXTE in the pointed mode during the period 2006 January 27 - August 2. We have carried out detailed timing analysis of the PCA data to study the properties of the QPOs and in this paper we present the results of this analysis. ", "conclusions": "\\label{Discussion section} The LFQPOs occur most frequently when the power law flux is the dominating component in the energy spectrum. Some times they are also present in the high luminosity state with the presence of a hard component. From table \\ref{table1.tab} and \\ref{table2.tab} it will be noticed that in all the cases of the detection of LFQPOs from XTE J1817-330, except the observations of MJD 53764 and 53775 in which no QPOs are detected, the ratio of the power law flux to the thermal disk flux lies in 0.20 to 1.13 range consistent with its occurrence only in the states with a significant power law component. Also note that the LFQPOs have significant coherence (Q $=$ $\\nu$/$\\delta\\nu $) with the Q in range of 3-11 and their rms amplitude vary from a few percent to as high as 13\\%.\\\\ Variation of the QPO frequency with the source intensity is another feature detected in some BHBs. The fundamental QPO frequency in XTE J1817-330 varies in a narrow band of 4.4-5.9 Hz. We have investigated the variation of the QPO frequency in the 2-8 keV band with the thermal disk component flux ( d$_{bb}$ ). This is shown in Fig \\ref{fig:fig9} and a trend similar to that of Fig \\ref{fig:fig7}(a) is seen here indicating that the frequency is correlated with the thermal disk component. As expected a clear 1:2 relationship of the QPO fundamental frequency and that of the first harmonic is seen. All the characteristics of the LFQPOs reported by us in this paper from XTE J1817-330 are similar to those seen in the other black hole binaries and further strengthen the black hole nature of this source (Remillard et al. 2003).\\\\ Correlation of the properties of LFQPOs with the spectral parameters of the BHBs has been studied in detail for several sources (Muno, Remillard and Morgan 2001; Tomsick and Kaaret 2001; Remillard et al. 2003; Belloni, Psaltis and van der Klis 2002; Vignarca et al. 2003; Rossi, Homan and Belloni 2004). These studies show that the LFQPO characteristics are generally well correlated with the thermal disk and the power law components of the energy spectra. While the QPO frequency is closely correlated with the disk flux, the amplitude of the QPOs for the fundamental frequency is found to track the flux of the power law component (Remillard et al. 2003). In general the QPO amplitude is higher for the higher energy X-rays up to about 20 keV and then it tends to decrease at the higher energy. Most of the LFQPOs detected in the black hole binaries occur below 10 keV. In two of the BHBs namely GRS 1915+105 and XTE J1550-564 the QPOs have been reported above 20 keV (Trudolyubov, Churazov \\& Gilfanov 1999; Remillard et al. 2003). In some cases the QPO detection is claimed in a broad spectral band of 2-60 keV but since no breakdown of QPO properties is given in the different energy intervals say below 10 keV and above 10 keV, it is not obvious whether the QPOs have indeed been detected in hard X-rays. The enigmatic source GRS 1915+105 is the only black hole binary in which the QPOs in 0.8-3.0 Hz have been reported at an energy up to 124 keV (Tomsick and Kaaret 2001). It is found that in this object, the amplitude of the fundamental frequency QPO increases with energy up to 29 keV and then decreases in 30-60 keV and 60-124 keV bands. Remillard et al. (2003) investigated the QPO characteristics in the transient XTE J1550-564 from the RXTE - PCA observations and detected LFQPOs in 2-13 keV and 13-30 keV energy channels making it only the second black hole binary in which the LFQPOs have been detected up to 30 keV. We have detected the QPOs from XTE J1817-330 in five of the observations up to an energy of $\\sim$ 25 keV, making it only the third BHB showing unambiguous presence of the LFQPOs at higher energy. From Table \\ref{table1.tab} it may be noticed that the amplitude of the fundamental QPOs is always higher in the 8-14 keV channel, being in 3.2-13.3 \\% range, compared to the values of 1.7-7.0 \\% in the 2-8 keV channel. At still higher energy 15-25 keV, the QPO amplitude is comparable to or slightly higher than that in the 8-14 keV indicating that it has reached a plateau level.\\\\ A detailed study of the QPO centroid frequency, its coherence, amplitude, phase lag and their dependence on the photon energy, is of vital importance to understand the origin of the QPOs and pin point the emission process. The QPOs are believed to originate in the innermost region of the accretion disk and the most common models explain their generation to the modulation of the disk flux by the Keplerian motion of the localized hot regions termed as 'blobs'. Lehr, Wagoner and Wilms (2000) have developed a model to compute by Monte Carlo simulations, the energy dependence of the QPO amplitude to probe the site of their origin in the accretion disk. They use two components with repeated Compton scattering to produce the high energy X-rays and assume a radial dependence of the disk temperature. They computed the energy dependence of the QPO amplitude for GRS 1915 + 105 and found it to be in agreement with the observation of Morgan and Remillard (1997). Thus they are able to localize the QPO origin in the inner disk. The amplitude of the QPOs will either increase or decrease with energy depending on the region of the disk in which the QPOs are produced and the temperature of the corona and its gradient. It is reasonable to assume that the QPO frequency is related to the dynamical time scale of the blobs and therefore, the LFQPOs observed by us in the 4.4-5.9 Hz from XTE J1817-330 will originate farther out in the accretion disk. In the outer region, the Compton scattering corona will be relatively cooler and the QPO amplitude will decrease with increasing energy. In our case we have detected a marginal increase in the amplitude of the QPOs at the higher energy in the two observations, a decrease in the amplitude in the other two observations and no change in amplitude in one case. This suggests that the site of QPO generation is itself dynamically varying in XTE J1817-330 due to variation in the X-ray luminosity which in turn depends on the accretion rate.\\\\ We have detected the QPOs in the 8-14 keV band but not in the 2-8 keV for the observations of MJD 53790.2 and 53790.3. This is similar to the detection of the QPOs at the high energy and its absence at the low energy in some observations from GRS 1915+105 (Chakraborti and Manickam 2000). This behavior of GRS 1915+105 was explained by Chakrabarti and Manickam (2000) on the basis of \"on\" and \"off\" (burst and quiescent) state of the source with the shock oscillation model. Further detailed studies of the LFQPOs at the higher energy are required to test the validity of the model. \\\\" }, "0911/0911.3448_arXiv.txt": { "abstract": "Weak gravitational lensing provides a sensitive probe of cosmology by measuring the mass distribution and the geometry of the low redshift universe. We show how an all-sky weak lensing tomographic survey can jointly constrain different sets of cosmological parameters describing dark energy, massive neutrinos (hot dark matter), and the primordial power spectrum. In order to put all sectors on an equal footing, we introduce a new parameter $\\beta$, the second order running spectral index. Using the Fisher matrix formalism with and without CMB priors, we examine how the constraints vary as the parameter set is enlarged. We find that weak lensing with CMB priors provides robust constraints on dark energy parameters and can simultaneously provide strong constraints on all three sectors. We find that the dark energy sector is largely insensitive to the inclusion of the other cosmological sectors. Implications for the planning of future surveys are discussed. ", "introduction": "In the last few decades, a wealth of cosmological data (from large scale structure \\citep[][2dF]{Peacock:2005}; the cosmic microwave background \\citep[][WMAP5]{Komatsu:2009}; supernovae (\\citealt[][SNLS]{Astier:2006}; \\citealt[][ESSENCE]{Miknaitis:2007}; \\citealt{Kowalski:2008a}); weak lensing \\citep{Schrabback:2009}) has revolutionised our vision of the Universe. In this concordance cosmology, initial quantum fluctuations are believed to have seeded dark matter perturbations in which the Large Scale Structure we observe today has formed. Within this concordance model the Universe is composed only of a small proportion of baryons (4\\%), the rest being dark matter (25\\%, which can be hot or cold) and dark energy. One of the main challenges today is to understand the nature of the mysterious dark energy which causes cosmic acceleration and constitutes 75\\% of the Universe's energy density \\citep{DETF, Peacock:2006}. There exists a wealth of potential models for dark energy. To distinguish these models the determination of the dark energy equation of state $w$ has gained importance since some models can result in very different expansion histories. Current data can constrain the dark energy equation of state $w$ to 10\\%, with the assumption of flatness, but a percentage level sensitivity as well as redshift evolution information are required in order to understand the nature of dark energy. Future cosmic shear surveys show exceptional potential for constraining the dark energy equation of state $w(z)$ \\citep{DETF, Peacock:2006} and have the advantage of directly tracing the dark matter distribution (see \\citealt{Hoekstra:2008} for a review). In fact, cosmic shear surveys have the potential to constrain all sectors of our cosmological model. As shear measurements depend on the initial seeds of structure, it can be used to probe the slope and running of the initial power spectrum \\citep[see e.g.][]{Liu:2009} which is central to our understanding of the inflationary model \\citep[see e.g.][]{Hamann:2007}. Shear measurements have also been used to complement neutrino constraints from particle physics (\\citealt{Tereno:2008}; \\citealt*{Ichiki:2008}) and galaxy surveys \\citep*{Takada:2006}, and future weak lensing surveys will provide bounds on the sum of the neutrino masses, the number of massive neutrinos and the hierachy (\\citealt*{Hannestad:2006, Kit2008b}; \\citealt{debernardisdraft2009}). The parameters that describe each sector are degenerate in the cosmic shear power spectrum, which means fixing parameters in one sector may results in anomalies being detected in another. We think all sectors should therefore be considered simultaneously. As each sector provides information about a different area of physics, we also argue they should also be considered on equal footing, i.e. have roughly the same number of parameters describing them, so that one sector is not favoured. Indeed, evidence for a departure from $\\Lambda$CDM may come from any sector. In section 2, we describe the cosmological parameter set we consider, which includes the three sectors of dark energy, initial conditions and neutrinos. We introduce a new parameter to include the second order running of the initial power spectrum. We also present the weak lensing tomography and the Fisher matrix forecast methods, and briefly discuss systematic effects. In section 3, we present the weak lensing constraints we expect from future space based surveys with and without \\textit{Planck} priors and investigate the stability of the results as new parameters are added to the analysis. In section 4, we optimise such a survey using the Figure of Merit as well as constraints on all sectors. In section 5, we present our conclusions. \\section[]{Method} \\subsection{Cosmology} \\label{Cosmology} We start by describing the 12-parameter cosmological model which includes dark energy, dark matter (hot and cold) and initial conditions sectors. Throughout this paper, we work within a Friedmann-Robertson-Walker cosmology. Our cosmological model contains baryonic matter, cold dark matter (CDM) and dark energy, to which we add massive neutrinos (i.e. hot dark matter -- HDM). We also consider different parametrizations of the primordial power spectrum (defined in section \\ref{Power_spectrum}. We allow for a non-flat geometry by including a dark energy density parameter $\\Omega_\\mr{DE}$ together with the total matter density $\\Omega_m$, such that in general $\\Omega_m+\\Omega_\\mr{DE}\\neq 1$. The dynamical dark energy equation of state parameter, $w=p/\\rho$, is expressed as function of redshift and is parametrized by a first-order Taylor expansion in the scale factor $a$ \\citep{ChevPol2001, Linder:2003}: \\bq w(a)=w_n + (a_n -a)w_a,\\eq where $a=(1+z)^{-1}$. The pivot redshift corresponding to $a_n$ is the point at which $w_a$ and $w_n$ are uncorrelated. Our most general parameter space consists of: \\begin{enumerate} \\item Total matter density -- $\\Omega_m$ (which includes baryonic matter, HDM and CDM) \\item Baryonic matter density -- $\\Omega_b$ \\item Neutrinos (HDM) -- $m_\\nu$ (total mass), $N_\\nu$ (number of massive neutrino species) \\item Dark energy parameters -- $\\Omega_{\\mathrm{DE}},\\; w_0,\\; w_a$ \\item Hubble parameter -- $h$ \\item Primordial power spectrum parameters -- $\\sigma_8$ (amplitude), $n_s$ (scalar spectral index), $\\alpha$ (running scalar spectral index), $\\beta$ (defined in section \\ref{Power_spectrum}) \\end{enumerate} We shall refer to this fiducial cosmology as `$\\nu\\mr{QCDM+\\alpha+\\beta}$'. We choose fiducial parameter values based on the five-year WMAP results \\citep{WMAP5} similar to those used in \\citet{Kit2008b}. The values are given in Table \\ref{Cosmo_models}. \\setcounter{table}{0} \\begin{table*} \\caption{Cosmological parameter sets used in our calculations. For each parameter set, the ticks ($\\checkmark$) and crosses ($\\times$) indicate whether a parameter is allowed to vary or not, respectively.} \\label{Cosmo_models} \\begin{tabular}{lcccccccccccc} Parameters & $w_0$\t&$w_a$&$\\Omega_\\mr{DE}$&$\\Omega_m$&$ \\Omega_b$&$h$&$\\sigma_8$&$n_s$&$\\alpha$&$\\beta$&$m_\\nu$&$N_\\nu$ \\\\ \\hline Fiducial values & $-0.95$&$0$&$0.7$&$0.3$&$ 0.045$&$0.7$&$0.8$&$1$&$0$\\quad&$0$\\quad&$0.66$&$3$\\\\ \\hline QCDM & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\times$ & $\\times$ & $\\times$ & $\\times$ \\\\ $\\mr{QCDM}+\\alpha$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\times$ & $\\times$ & $\\times$ \\\\ $\\mr{QCDM}+\\alpha+\\beta$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\times$ & $\\times$ \\\\ $\\nu$QCDM & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\times$ & $\\times$ & $\\checkmark$ & $\\checkmark$ \\\\ $\\nu\\mr{QCDM}+\\alpha$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\times$ & $\\checkmark$ & $\\checkmark$ \\\\ $\\nu\\mr{QCDM}+\\alpha+\\beta$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ & $\\checkmark$ \\\\ \\end{tabular} \\end{table*} \\subsection{Matter power spectrum} \\label{Power_spectrum} The matter power spectrum is defined as: \\bq \\langle \\delta(\\mathbf{k})\\delta^\\ast(\\mathbf{k}')\\rangle= {(2\\pi)}^3\\delta_D^3(\\mathbf{k}-\\mathbf{k}') P(k) \\eq and can be modelled by: \\bq P(k,z)=\\frac{2\\pi^2}{k^3}A_sk^{n_s(k)+3}{T^2(k,z)}\\left(\\frac{D(z)}{D(0)}\\right)^2,\\eq where $A_s$ is the normalisation parameter, $T(k,z)$ is the transfer function and $D(z)$ is the growth function. The primordial spectral index is denoted by $n_s(k)$, and can depend on the scale $k$. In our cosmological model, the shape of the primordial power spectrum is of particular interest, since it may mimic some of the small-scale power damping effect of massive neutrinos. In the concordance model, the primordial power spectrum is generally parametrized by a power-law \\citep[see e.g.][]{Kosowsky:1995, Bridle:2003} \\bq\\mathcal{P}_\\chi(k)=A_s\\left(\\frac{k}{k_{s0}}\\right)^{n_s-1}.\\eq We parametrize the running of the spectral index by using a second-order Taylor expansion of $\\mathcal{P}_\\chi$ in log-log space, defining the running as $\\alpha = \\mr{d} n_s/\\mr{d}\\ln k |_{k_0}$, so that the primordial power spectrum is now scale-dependent, with the scalar spectral index defined by \\citep{Spergel:2003, Hannestad:2002} \\bq n_s(k)=n_s(k_0)+\\frac{1}{2}\\frac{ \\mr{d} n_s}{ \\mr{d} \\ln k}\\bigg|_{k_0} \\ln\\left(\\frac{k}{k_0}\\right),\\eq where $k_0$ is the pivot scale. We use a fiducial value of ${k_0=0.05 \\mr{Mpc^{-1}}}$ for the primordial power spectrum pivot scale. Although it is motivated by simplicity and standard slow-roll inflation theory, the second-order truncated Taylor expansion is limited and may lead to incorrect parameter estimation \\citep*[see][]{Abazajian:2005, Leach:2003}. In order to test this, we allow an extra degree of freedom in the primordial power spectrum by adding a third-order term in the Taylor expansion, which we call $\\beta:$ \\bq n_s(k)=n_s(k_0)+\\frac{1}{2!}\\alpha\\ln\\left(\\frac{k}{k_0}\\right)+\\frac{1}{3!}\\beta \\ln\\left(\\frac{k}{k_0}\\right)^2,\\eq where $\\beta=\\mr{d}^2 n_s/\\mr{d} \\ln k^2 |_{k_0}$. We use the \\cite{EH97} analytical fitting formula for the time-dependent transfer function to calculate the linear power spectrum, which includes the contribution of baryonic matter, cold dark matter, dark energy and massive neutrinos, with the modification in the transfer function suggested by \\citet*{Kiakotou:2008jk}. We use the \\citet{Smithetal} correction to calculate the non-linear power spectrum. The matter power spectrum is normalised using $\\sigma_8$, the root mean square amplitude of the density contrast inside an $8\\,h^{-1}\\mr{Mpc}$ sphere. Following \\citet{EH97}, we assume $N_\\nu$, the number of massive (non-relativistic) neutrino species, to be a continuous variable, as opposed to an integer. Neutrino oscillation experiments do not, at present, determine absolute neutrino mass scales, since they only measure the difference in the squares of the masses between neutrino mass eigenstates \\citep{Quigg:2008}. Cosmological observations, on the other hand, can constrain the neutrino mass fraction, and can distinguish between different mass hierarchies (see \\citealt*{Elgaroy:2005} for a review of the methods). The \\citeauthor{EH97} transfer function assumes a total of three neutrino species (i.e. $N_\\mr{massless}+N_\\nu=3$), with degenerate masses for the most massive eigenstates, i.e. if $m_\\nu$ is the total neutrino mass, then \\bq m_\\nu=\\sum^{N_\\nu}_{i=0}m_i=N_\\nu m_i, \\eq where $m_i$ is the same for all eigenstates. Thus, $N_\\nu=2$ for the normal mass hierarchy, and $N_\\nu=1$ for the inverted mass hierarchy, while $N_\\nu=3$ corresponds to the case where all three neutrino species have the same mass \\citep[see][]{Quigg:2008}. The temperature of the relativistic neutrinos is assumed to be equal to $(4/11)^{1/3}$ of the photon temperature \\citep{KolbTurner1990}. Dark energy affects the matter power spectrum in three ways. Its density $\\Omega_\\mr{DE}$ changes the normalisation and $k_{eq}$, the point at which the power spectrum turns over. $\\Omega_\\mr{DE}$ and the dark energy equation of state parameter $w$ change the growth factor at late times by changing the Hubble rate. In addition to this, for departures from a cosmological constant the shape of the matter power spectrum on large scales is affected through dark energy perturbations. In all our calculations, we only consider small deviations from $w_0=-1$, and so neglect dark energy perturbations, only considering the first two mechanisms. In comparing parameter constraints in different parameter spaces, we shall use six parameter sets. We start with the simplest set (QCDM) to which we add neutrino and additional primordial power spectrum parameters. In all cases the central fiducial model (given in Table \\ref{Cosmo_models}) is the same, but the number of parameters marginalized over varies. \\subsection{Weak lensing tomography} The cosmological probes considered in this paper are tomographic cosmic shear and the CMB. In weak lensing surveys, the observable is the convergence power spectrum. In our analysis, we calculate this quantity from the matter power spectrum via the lensing efficiency function. Our convergence power spectrum therefore depends on the survey geometry and on the matter power spectrum. We use the power spectrum tomography formalism by \\citet{Hu:2004}, with the background lensed galaxies divided into 10 redshift bins. Cosmological models are then constrained by the power spectrum corresponding to the cross-correlations of shears within and between bins. The 3D power spectrum is projected onto a 2D lensing correlation function using the \\citet{Limber:1953} equation: \\bq C_\\ell^{ij}=\\int \\mr{d}z \\frac{H}{D^2_A} W_i(z)W_j(z)P(k=\\ell/D_A,z),\\eq where $i$, $j$ denote different redshift bins. The weighting function $W_i(z)$ is defined by the lensing efficiency: \\bq W_i(z)=\\frac{3}{2}\\Omega_m \\frac{H_0}{H}\\frac{H_0 D_{OL}}{a}\\int_z^\\infty \\mr{d} z^{\\prime}\\frac{D_{LS}}{D_{OS}}P(z^{\\prime}), \\eq where the angular diameter distance to the lens is $D_{OL}$, the distance to the source is $D_{OS}$, and the distance between the source and the lens is $D_{LS}$ (see \\citealt{Hu:2004} for details). Our multipole range is $10<\\ell<5000$. The galaxies are assumed to be distributed according to the following probability distribution function \\citep*{Smail1994}: \\bq P(z)=z^a \\exp\\left[-\\left(\\frac{z}{z_0}\\right)^b\\right],\\eq where $a=2$ and $b=1.5$, and $z_0$ is determined by the median redshift of the survey $z_m$ \\citep[see e.g][]{AR2007}. Our survey geometry follows the parameters for a `wide' all-sky survey with $A_s=20 000\\,\\mr{sq\\; degrees}$. The survey parameters are shown in Table \\ref{survey_params}. The median redshift of the density distribution of galaxies is $z_\\mr{median}$ and the observed number density of galaxies is $n_g$. We include photometric redshift errors $\\sigma_z(z)$ and intrinsic noise in the observed ellipticity of galaxies $\\sigma_\\epsilon$. We follow the definition $\\sigma_\\gamma^2=\\sigma_\\epsilon^2$, where $\\sigma_\\gamma$ is the variance in the shear per galaxy \\citep[see][]{Bartelmann:2001}. \\setcounter{table}{1} \\begin{table} \\begin{center} \\caption{Fiducial parameters for the all-sky weak lensing survey considered.} \\label{survey_params} \\begin{tabular}{@{}lr@{}} \\hline $A_s$/sq degree &20 000\\\\ $z_\\mr{median}$&0.9\\\\ $n_g/\\mr{arcmin}^{2}$&35\\\\ $\\sigma_z(z)/(1+z)$&0.025\\\\ $\\sigma_\\epsilon$&0.25\\\\ \\hline \\end{tabular} \\end{center} \\end{table} \\subsection{Error forecast} The predictions for cosmological parameter errors presented in this paper use the Fisher matrix formalism. The Fisher matrix gives us the lower bound on the accuracy with which we can estimate model parameters from a given data set \\citep*{Fisher1935, Tegmark:1997, Kitching:2009}. In calculating forecast survey errors, we are implicitly making assumptions about the parameter set \\citep*[see the discussion of nested models in][]{Heavens:2007}. We want to know whether our constraints are robust against variations in the parameterisation of the cosmological model. This is of particular importance when dark energy constraints are considered, because of the degeneracies with other parameters (see Fig. \\ref{pk_deg}). The forecast parameter precision is improved by combining independent experiments. Using this technique, joint constraints or error forecasts can be obtained by combining weak lensing with other observational techniques. The Fisher matrix for the shear power spectrum is given by \\citep{Hu:2004}: \\bq\\label{eq_10} F_{\\alpha\\beta} = f_\\mr{sky}\\sum_\\ell{\\frac{(2\\ell+1)\\Delta \\ell}{2}}\\mr{Tr}\\left[D_{\\ell\\alpha}\\widetilde{C}_\\ell^{-1}D_{\\ell\\beta}\\widetilde{C}_\\ell^{-1}\\right],\\eq where the sum is over bands of multipole $\\ell$ of width $\\Delta \\ell$, $\\mr{Tr}$ is the trace, and $f_\\mr{sky}$ is the fraction of sky covered by the survey. Equation \\ref{eq_10} assumes the likelihod obeys a Gaussian distribution with zero mean. The observed power spectra for each pair $i,j$ of redshift bins are written as the sum of the lensing and noise spectra: \\bq \\widetilde{C}_\\ell^{ij}=C_\\ell^{ij}+N_\\ell^{ij}.\\eq The derivative matrices are given by \\bq [D_{\\alpha}]^{ij}=\\frac{\\partial C_\\ell^{ij}}{\\partial p_\\alpha} ,\\eq where $p_\\alpha$ is the vector of parameters in the theoretical model. In order to quantify the potential for a survey to constrain dark energy parameters, we use the Figure of Merit, as defined by the Dark Energy Task Force \\citep{DETF}: \\bq \\mr{FoM}=\\frac{1}{\\Delta w_n\\Delta w_a}.\\eq \\subsection{\\textit{Planck} priors} In this article, together with lensing-only constraints, we also include joint lensing and CMB constraints. To combine constraints from different probes we add the respective Fisher matrices: $\\bf{F}_\\mr{joint}=\\bf{F}_\\mr{lensing}+\\bf{F}_\\mr{CMB}$. For our CMB priors, we use the forthcoming \\textit{Planck} mission as our survey. The \\textit{Planck} Fisher matrix is calculated following \\citet{Rassat:2008}, which estimates errors using information from the temperature and E mode polarisation (i.e. TT, EE, and TE), with the help of the publicly available \\textsc{camb} code \\citep*{Lewis:2000}. We conservatively do not use information from B modes, and only use the 143~GHz channel, assuming other frequencies will be used for the foreground removal. This is conservative compared to other \\textit{Planck} priors in the literature. More details are given in \\citet[][Appendix B]{Rassat:2008}. The full parameter set for the \\textit{Planck} calculation is: $\\{\\Omega_\\mr{DE},\\,w_0,\\,w_a,\\,\\Omega_m,\\,\\Omega_b,\\,m_\\nu,\\, N_\\nu,\\,h,\\,\\sigma_8,\\,n_s,\\,\\alpha, \\,\\tau\\}$. We use the same central values as for our weak lensing calculations, as described above, with a fiducial value for the reionisation optical depth $\\tau=0.09$ and subsequently marginalize over $\\tau$. We consider neutrino parameters as $\\Omega_\\nu h^2$ and $N_\\nu$ and use a Jacobian to translate this into constraints on $m_\\nu$ and $N_\\nu$. \\subsection{Systematic effects} Weak lensing measurements are affected by systematic effects which reduce the precision on cosmological parameters and introduce bias (see \\citealt{Refregier:2003a} and \\citealt{Schneider:2006} for a review). In this section we discuss some of these effects. Intrinsic correlations can contaminate the lensing signal. Solutions include using tomography or 3D lensing, which decouple the long-distance line-of-sight effects from the the physical proximity of the galaxies \\citep[see e.g.][]{King:2003, Bridle:2007, Kit2008b}. Using this approach, a nulling technique for shear-intrinsic ellipticity has been proposed by \\citet{Joachimi:2008, Joachimi:2009}. Measurement systematics are due to PSF effects \\citep[see][]{KSB1995}. The results presented in \\citet{Bridle:2009} indicate that the required accuracy will be reached for the next generation of all-sky weak lensing surveys. Redshift distribution systematics, which lead to an uncertainty in the median galaxy redshift, cause an uncertainty in the amplitude of the matter power spectrum \\citep{Hu:1999b}. The problem of photometric redshift systematics can be met given a number of galaxies in the spectroscopic calibration sample of $10^4-10^5$ \\citep{Ma:2006, AR2007}. Finally there are theoretical uncertainties on the shape of the matter power spectrum. The existing non-linear corrections to the matter power spectrum \\citep{Peacock:1996ys, Ma:1998} are only accurate to about $10\\%$ and disagree with one another to this level in the non-linear r\\'egime \\citep[see][]{Huterer:2001}. Newer prescriptions such as those by \\citet{Smithetal} offer more accurate predictions particularly for non-$\\Lambda\\mr{CDM}$ cosmological models. The error in the non-linear part may still be significant if effect of massive neutrinos is included \\citep*[see e.g.][]{Saito:2008}. Current semi-analytical models need to be improved to match the degree of statistical accuracy expected for future weak lensing surveys. The solution is to run a suite of $N$-body ray-tracing simulations \\citep[see e.g.][]{White:2004, Huterer:2005, Hilbert:2009, Sato:2009, Teyssier:2009}. ", "conclusions": "The main aim of this paper was to put three different sectors of the cosmological model: dark energy, dark matter (cold and neutrinos) and initial conditions on equal footing while forecasting constraints for a future weak lensing survey. To do this we introduce a new parameter $\\beta$, which models the second order running spectral index. We have forecast errors for an all-sky tomographic weak lensing survey, and for weak lensing+\\textit{Planck}, using different cosmological parameter sets, studying the effect of the addition of parameters in the model. We have shown that error forecasts for some parameters are stable against changes in the parameter set (Table \\ref{table_lensing}), and that degeneracies between the dark energy parameters $w_0$ and $w_a$ are not significantly affected by the addition of parameters (Fig. \\ref{ellipses_article_joint}). We have investigated the shift in the optimal pivot scale for primordial power spectrum constraints with the addition of extra parameters (Fig. \\ref{Prim_v_k0}). We have shown that parameter constraints are nevertheless dependent on the parameterisation of the primordial power spectrum. With the addition of CMB priors, we have shown that we can obtain improved constraints in this sector, which reduces the dependency on constraints on the parameter set. We have obtained predicted constraints for both the total neutrino mass and the number of massive neutrino species. This article has a resonance with \\citet{debernardisdraft2009}, in which it is shown that the neutrino mass hierarchy can be constrained using cosmic shear, using a more general parameterisation of the neutrino mass splitting. We find that for the parameters that are common between the two articles there is an agreement between the predicted errors, despite the slightly different parameter sets and assumptions. The results obtained here also mirror those in \\citet{Zunckel:2007}, in which it is found that the neutrino mass constraints are degraded when the hypothesis space is enlarged. This is due to degeneracies between the neutrino mass and other parameters. Adding priors only tightens the constraints when the additional information comes from independent experiments, which reduces the freedom in the degenerate parameters. We have used information from the CMB, which constrains the primordial power spectrum particularly well, resulting in improved constraints on our neutrino parameters. In the neutrino sector, parameter constraints would be improved by a hierarchical parameterisation in which different neutrino species have non-degenerate masses. This would model more accurately the process whereby each massive species becomes non-relativistic at a different redshift. While the different transition redshifts have only a very small effect on the CMB anisotropy power spectrum, the effect is non-negligible in future cosmic shear experiments which measure the matter power spectrum to a sufficient accuracy to discriminate between different mass hierarchies. This article concludes that a future all-sky weak lensing survey with CMB priors provides robust constraints on dark energy parameters and can simultaneously provide strong constraints on all parameters. The results presented here show that error forecasts from our weak lensing survey are stable against the addition of parameters to the fiducial model, and that this stability is improved by adding CMB priors." }, "0911/0911.3397_arXiv.txt": { "abstract": "We report on a new class of fast-roll inflationary models. In a huge part of its parameter space, inflationary perturbations exhibit quite unusual phenomena such as scalar and tensor modes freezing out at widely different times, as well as scalar modes reentering the horizon during inflation. One specific point in parameter space is characterized by extraordinary behavior of the scalar perturbations. Freeze-out of scalar perturbations as well as particle production at horizon crossing are absent. Also the behavior of the perturbations around this quasi-de Sitter background is dual to a quantum field theory in flat space-time. Finally, the form of the primordial power spectrum is determined by the interaction between different modes of scalar perturbations. ", "introduction": "Inflation has become the standard paradigm for the description of the early universe, solving the monopole, horizon and flatness problems as well as providing for a mechanism to seed structure in the universe. Various explicit models for this period of exponential expansion, most relying on the dynamics of real scalar ``inflaton'' fields, mostly agree with current observations of perturbations of an isotropic background, such as cosmic microwave background (CMB) fluctuations and large scale structure (LSS) (although see \\cite{Copi:2008hw} and references therein). When analyzing these models, we mostly rely on the assumption of ``slow-roll'' \\cite{Stewart:1993bc, Sasaki:1995aw, Martin:2002vn} -- the slow evolution of the inflaton field -- although fast-roll models have also been considered \\cite{Linde:2001ae}. Slow roll is an approximation made to solve the perturbation equations, but not an absolute requirement for an extended period of accelerated expansion. Even though the scalar spectral index has arguably been measured by the WMAP satellite\\cite{Dunkley:2008ie} to be smaller than unity, $n_s\\approx0.96$, this constrains the slow-roll parameters \\begin{eqnarray} \\epsilon\\equiv-\\frac{\\partial_t{H}}{H^2}\\,, \\eta\\equiv2\\frac{\\partial_\\phi^2H}{H}\\,, \\zeta^2=\\frac{1}{2}\\frac{\\partial_\\phi H \\partial_\\phi^3H}{H^2}\\,, \\end{eqnarray} all in reduced Planck units $M_p^2=\\frac{1}{8\\pi G}=1$, only if they are small. This was exactly the assumption made when solving the perturbation equations, be it using Hankel functions \\cite{Stewart:1993bc}, the $\\Delta N$ formalism \\cite{Sasaki:1995aw, Starobinsky:1986fxa}, or the WKB approximation \\cite{Starobinsky:1982ee, Martin:2002vn}. If the slow-roll parameters are large, then it is impossible to make any general statement about their values from the measurement of $n_s$. While it is true that $0<\\epsilon<1$ for successful inflation, higher order slow-roll parameters may be large, contrary to intuition. Besides exploring unknown corners of inflationary model space, there are many additional motivations to study models with large $\\eta$. For example, string theory motivated models of inflation are typically characterized by large values of $\\eta$ \\cite{Kachru:2003sx}. Also, inflationary non-Gaussianities are suppressed by powers of the slow-roll parameters \\cite{Maldacena:2002vr}, so that one might expect large non-Gaussianities in fast-roll inflationary models. As the problem of analyzing the behavior of fast-roll inflationary models is technically very challenging, so far those models received only limited attention. Essentially, there two ways to study this largely unexplored realm. One of them is finding exact solutions to a coupled system of non-linear differential equations. Another approach is to solve these equations numerically. In our paper, we adopt both approaches, providing an exact solution for a subset of initial conditions as well as numerically charting the phase space of a new class of inflationary models with large acceleration, i.e. large $\\eta$. This allows us to study the behavior of both background and perturbations in new and interesting regimes, where we find various unexpected and even downright bizarre phenomena. While our model can be made to agree with observations for a small range of parameters, it possesses several most unusual features for parameter ranges that are not compatible with observations. Some of these features might seem counter-intuitive. For example, according to general lore, perturbation modes that exit the Hubble horizon, freeze out and remain classical until the present epoch. We find that is not true for a particular model of this family. The modes remain quantum mechanical at all times and wavelengths, even long after they cross the Hubble scale. Also, it is generally believed that the interactions among the different modes of scalar perturbations are suppressed by powers of the slow roll parameters, so that only insignificant amounts of non-Gaussianity are created during single-field inflation. Instead we show that it is the interaction between different modes that defines the very form of the scalar power spectrum. Other unusual features include the fact that tensor and scalar modes can freeze out at different times and even unfreeze during inflation. Finally, it turns out that for a specific member of this family, the theory of scalar perturbations is described by a quantum field theory in flat space time. \\begin{figure*}[t!] (a)\\includegraphics[width=0.45\\linewidth]{V_of_phi_p0_04} (b)\\includegraphics[width=0.45\\linewidth]{V_of_phi_p-2} \\caption{(a)$V(\\phi)$ for $p=0.04$. Inflation takes place close to the top of the potential, rolling down from $\\phi\\approx0$ towards larger values of $|\\phi|$. The potential is very shallow and turns negative only for $\\phi>\\frac{\\sqrt{24}}{0.04}\\approx122$, i.~e.~about $30$ e-folds after the end of inflation. For this model, the power spectrum of scalar perturbations has a spectral index that is compatible with observations, $n_s=1-p=0.96$. (b)$V(\\phi)$ for $p=-2$. For negative $p$, the potential goes to $-\\infty$ for $\\phi\\rightarrow\\infty$. Inflation takes place rolling from small values of $|\\phi|$ towards $\\phi=0$ and ends through tunneling. The scalar power spectrum possess some interesting features, see text.} \\label{fig:potentialp2} \\end{figure*} Our paper is organized as follows. In section \\ref{sec:most_curious_model} we present an explicit one-parameter family of fast-roll models whose background equation is exactly solvable for certain initial conditions. In section \\ref{sec:linear_perturbations} we discuss the spectrum of scalar and tensor perturbations, which can be reliably computed despite the fact that the slow-roll parameter $\\eta$ may be larger than unity. In section \\ref{sec:p-2} we present an in-depth analysis of the perturbations for a special member of the family of models, where the scalar perturbations do not follow the usual behavior of oscillating, being stretched to the horizon, and freeze-out, but instead obey an exactly massless harmonic oscillator type equation. In section \\ref{sec:conclusions} we conclude and offer some ideas for future work on this and similar classes of models. ", "conclusions": "\\label{sec:conclusions} In this paper we discuss the behavior of the family of inflationary models \\begin{eqnarray} \\label{eq:Vconclusion} V&=&H_0^2M_P^2 e^{-\\frac{p}{4}\\frac{\\phi^2}{M_P^2}}\\left(3-\\frac{p^2}{8}\\frac{\\phi^2}{M_P^2}\\right)\\,, \\end{eqnarray} with arbitrary $p$ and $H_01$ \\\\ \\hline $-61$\\\\ & $\\infty$ long classical inflation\\\\ \\hline $-41$\\\\ & $\\infty$ long classical inflation\\\\ \\hline $p=-2$ & no freeze out\\\\ & no gravitational particle production\\\\ & no stochastic eternal inflation\\\\ & non-suppressed mode interactions\\\\ & (pre)thermalization\\\\ & dual to a QFT in flat space\\\\ & $|\\eta|\\approx 1$\\\\ & $\\infty$ long classical inflation\\\\ \\hline $-2