{ "0609/astro-ph0609025_arXiv.txt": { "abstract": "{We quantitatively investigate how collisional avalanches may develop in debris discs as the result of the initial breakup of a planetesimal or comet-like object, triggering a collisional chain reaction due to outward escaping small dust grains. We use a specifically developed numerical code that follows both the spatial distribution of the dust grains and the evolution of their size-frequency distribution due to collisions. We investigate how strongly avalanche propagation depends on different parameters (e.g., amount of dust released in the initial breakup, collisional properties of dust grains, and their distribution in the disc). Our simulations show that avalanches evolve on timescales of $\\sim\\!1000$\\,years, propagating outwards following a spiral-like pattern, and that their amplitude exponentially depends on the number density of dust grains in the system. { We estimate a probability for witnessing an avalanche event as a function of disc densities, for a gas-free case around an A-type star, and find that features created by avalanche propagation can lead to observable asymmetries for dusty systems with a $\\beta$\\,Pictoris-like dust content or higher. Characteristic observable features include: (i) a brightness asymmetry of the two sides for a disc viewed edge-on, and (ii) a one-armed open spiral or a lumpy structure in the case of face-on orientation. A possible system in which avalanche-induced structures might have been observed is the edge-on seen debris disc around HD\\,32297, which displays a strong luminosity difference between its two sides. } } ", "introduction": "Direct imaging of circumstellar discs \\citep[e.g.,][]{Heap00, ClampinKrist03, Liu04, Schneider05} have provided resolved disc morphologies for several systems (e.g., \\bpic, HD\\,141569A, HD\\,100546, HD\\,32297) and have shown that dust distribution is not always smooth and axisymmetric. Warps, spirals, and other types of asymmetries are commonly observed \\citep[e.g.,][ for the \\bpic~system]{Kalas95}. These morphological features can provide hints on important ongoing processes in the discs and improve our understanding of the evolution of circumstellar discs and of planetary formation. The usual explanation proposed for most of these asymmetries is the perturbing influence of an embedded planet. As an example, the warp in the \\bpic~disc has been interpreted as induced by a jovian planet on an inclined orbit \\citep{Mouillet97, Au01}. Likewise, for annulus-like discs with sharp inner or outer edges, the most commonly proposed explanation is truncation or gap opening due to planets or bound stellar companions \\citep[e.g.,][]{Au04}, although alternative mechanisms such as gas drag on dust grains within a gas disc of limited extent have also been proposed \\citep{TA01}. For spiral structures, authors have also been speculating on gravitational instabilities \\citep{Fuka04}, as well as on a bound stellar companion \\citep{Au04}. The catastrophic breakup of one single large object releasing a substantial amount of dust fragments could be an alternative explanation for some observed asymmetries. \\cite{WyattDent02} have examined how such collisionally produced bright dust clumps could be observed in Fomalhaut's debris disc. Likewise, such clumps have been proposed by \\citet{Telesco05} as a possible explanation for mid-infrared brightness asymmetries in the central \\bp~disc, but only based on preliminary order of magnitude estimates. More recently, the detailed study of \\citet{KenyonBromley05} investigated the possibility of detecting catastrophic two-body collisions in debris discs and found that such a detection would require the breakup of 100-1000\\,km objects. The common point between these different studies is that they focus on global luminosity changes due to the debris cloud directly produced by the shattering events themselves. In the present paper, we re-examine the consequences of isolated shattering impacts from a different perspective, i.e., by considering the collisional evolution of the produced dust cloud $after$ its release by the shattering event. The main goal here is to study one possibly very efficient process, first proposed by \\citet{Art97}, but never quantitatively studied so far, i.e., the so-called collisional avalanche mechanism. The basic principle of this process is simple. After a localized disruptive event, such as the collisional breakup of a large cometary or planetesimal-like object, a fraction of the dust then produced is driven out by radiation pressure on highly eccentric or even unbound orbits. These grains moving away from the star with significant radial velocities can breakup or microcrater other particles farther out in the disc, creating in turn even more small particles propagating outwards and colliding with other grains. Should this collisional chain reaction be efficient enough, then a significant increase in the number of dust grains could be achieved. In this case, the consequences of a single shattering event, in terms of induced dust production, could strongly exceed that of the sole initially released dust population. The outward propagation of the dusty grains could then induce observable asymmetric features in the disc, even if the initially released dust cloud is undetectable. The goal of this work is to perform the first quantitative study of the avalanche process and investigate the morphology of avalanches in debris discs, under the assumption that dust dynamics is not controlled by gas \\citep{Lagrange00}. For this purpose we have created a numerical code, described in Sect.~\\ref{sec:model}, that enables us to simulate the coupled evolution of dynamics and size-frequency distribution of dusty grains. The results of our simulations, which explore the effect of several parameters (total mass and radial distribution of dust in the disc, mass and size distribution of the planetesimal debris, physical properties of the grains and the prescription for collisional outcome for grain-grain collisions) are presented in Sect.~\\ref{sec:res}. In Sect.~\\ref{sec:obs} we examine under which conditions avalanche-induced features might become observable. We end with a discussion of the probability of witnessing an avalanche (Sect.~\\ref{sec:discussion}) and finally a summary (Sect.~\\ref{sec:sum}). ", "conclusions": "\\label{sec:discussion} \\subsection{Probability of witnessing an avalanche event} \\label{sec:Pobs} The numerical investigation of the previous sections has shown that collisional avalanches are a powerful and efficient mechanism that naturally develops in debris discs after the breakup of a large planetesimal. However, in our nominal case of a $\\beta$\\,Pic-like system and $M_0=10^{20}$\\,g of dust initially released, the asymmetric features produced by the avalanche probably remain too weak to be observable in scattered light (Sect.~\\ref{sec:obs_nom}). This result should, however, be taken with great care since our parameter exploration has shown that avalanche strength strongly depends on several critical and often poorly constrained parameters. The first set of parameters is those linked to the initial breakup event. Here we obtain the intuitive result that higher amounts of initially released dust leads to more powerful avalanches (see Sect.~\\ref{sec:mo}), with the avalanche strength scaling linearly with $M_0$. This is not unique to the avalanche mechanism: \\citet{KenyonBromley05} find a similar dependence when only considering the signature of the cloud of primary debris produced immediately after the planetesimal breakup. What distinguishes our results from studies in which only dust released at impact is considered is that avalanches strongly depend on the number density of dust in the disc. Section~\\ref{sec:fp} has indeed shown that the global optical depth of the dust disc $\\tau_{\\|}$ is the parameter avalanche development depends most on, the dependence being close to an exponential. We have seen that other parameters, mostly related to the way the physical response of grains to collisions is modeled, might also lead to observable events when stretched to the extreme values that were numerically explored here. This is in particular the case for $Q_*$, for which very low $\\simeq 10^{6}$\\,erg/g values might lead to powerful avalanches. We shall however leave these ``technical'' parameters aside to focus on the 2 parameters directly related to the system's properties themselves, i.e., the optical depth, both $\\tau_{\\bot}$ and $\\tau_{\\|}$, and the initial amount of dust released $M_0$, and derive an order-of-magnitude estimate for the probability of witnessing avalanche events as a function of these parameters. From the results of Sect.~\\ref{sec:res}, the $L_\\mathrm{av}/L_\\mathrm{d}$ criterion for observability might be written \\begin{equation} \\left( \\frac{\\tau_{\\bot}}{\\tau_\\mathrm{\\bot,nom}} \\right)^{-1} \\frac{F_\\mathrm{max}}{F_\\mathrm{max(nom)}}\\, \\frac{M_0}{10^{20}\\,\\mathrm{g}} \\ga 100, \\label{visi0} \\end{equation} which is equivalent to saying that the luminosity ratio between avalanche and field grains should be at least 100 times higher than in the nominal case (for which $L_\\mathrm{av}/L_\\mathrm{d} \\sim 10^{-2}$). Section~\\ref{sec:mc} has shown that $F_\\mathrm{max}$ is independent of $M_0$, so that in our approximation $F_\\mathrm{max}$ is only a function of $\\tau_{\\|}$, and this $\\tau_{\\|}$ dependence is given by Eq.~\\ref{eq:Ftau}. Thus, Eq.~\\ref{visi0} reduces to \\begin{equation} \\exp \\left[ 5.3 \\left( \\frac{\\tau_{\\|}}{\\tau_\\mathrm{\\|,nom}} \\right)^{0.55}\\right] \\, \\frac{M_0}{10^{20}\\,\\mathrm{g}} ~ \\frac{\\tau_{\\bot,nom}}{\\tau_\\mathrm{\\bot}}\\, \\ga 2\\times10^{4}, \\label{visi1} \\end{equation} which gives a direct link between a given disc density ($\\tau_{\\bot}$ and $\\tau_{\\|}$) and the { minimum mass of released dust} able to produce a visible avalanche in such a disc (the denser the disc, the smaller the corresponding $M_0$ value). The other important issue affecting witnessing probabilities is the duration of an avalanche. Our simulations show that the typical lifetime of an avalanche-induced pattern is $t_\\mathrm{av} \\sim 10^3$\\,yrs. With this value and Eq.~\\ref{visi1}, one can estimate the probability $P_\\mathrm{obs}$ of witnessing an observable avalanche event in a given disc: \\begin{equation} P_\\mathrm{obs} = \\frac{t_\\mathrm{av}}{t_\\mathrm{shatt(M_0,\\tau_{\\bot})}}, \\label{probv0} \\end{equation} where $t_\\mathrm{shatt(M_0,\\tau)}$ is the average time between 2 shatterings producing $M_0$ of dust in a disc of average optical depth $\\tau_\\bot$, with $M_0$, $\\tau_{\\|}$ and $\\tau_\\bot$ satisfying Eq.~\\ref{visi1}. { As suggested in Sect.~\\ref{sec:plb}, we consider that the object releasing $M_0$ of dust has a mass $M_\\mathrm{PB}\\simeq 10M_0$.} { From unpublished results of the \\citet{Th03} simulations of collisional rates and outcomes in the inner \\bpic~disc, we determine that the approximate timescale for the shattering of a $M_\\mathrm{PB}=10M_0$ object to occur in the innermost $<50$\\,AU (the typical location for the initial shattering events considered in our simulations) of a \\bpic~like system is $t_\\mathrm{shatt}\\simeq 150[(10M_0)/10^{21}$\\,g$]^{1.25}$\\,yrs. Since, for systems with similar spatial distributions, the frequency of collisional events is proportional to the square of a system's total mass, we get the empirical relation: \\begin{equation} t_\\mathrm{shatt(M_0,\\tau_\\bot)} \\simeq \\,150\\, \\left( \\frac{\\tau_\\bot}{\\tau_\\mathrm{\\bot,nom}} \\right)^{-2} \\left( \\frac{M_0}{10^{20}\\, \\mbox{g}} \\right)^{1.25} \\,\\mbox{yrs,} \\label{eq:tshat} \\end{equation} where we implicitly assume that the system's spatial distribution is the same as in the nominal case, so that the ratio between two systems' total masses is equal to the ratio $\\tau_\\bot / \\tau_\\mathrm{\\bot,nom}$ anywhere in the disc. This equation should of course be regarded as giving a $very$ rough estimate, since $t_\\mathrm{shatt(M_0,\\tau_\\bot)}$ depends on many poorly constrained parameters, like the number density of planetesimals and their average kinetic energy at impact. Equation~\\ref{eq:tshat}, however, gives the global trend of the way $t_\\mathrm{shatt(M_0,\\tau_\\bot)}$ increases with $M_0$. Taking the lowest $M_0$ value satisfying Eq.~\\ref{visi1} and plugging it into Eq.~\\ref{eq:tshat}, we get, from Eq.~\\ref{probv0}: \\begin{equation} P_\\mathrm{obs} \\approx 3 \\times 10^{-5} \\left( \\frac{\\tau_\\bot}{\\tau_\\mathrm{\\bot,nom}} \\right)^{0.75} \\exp \\left(6.6 \\left( \\frac{\\tau_\\|}{\\tau_\\mathrm{\\|,nom}} \\right)^{0.55}\\right). \\label{probv2} \\end{equation} Equation \\ref{probv2} indicates that $P_\\mathrm{obs}\\simeq 0.03$ for the nominal case field particle disc, which means that we have about a $3\\%$ chance of witnessing the avalanche caused by the breakup of a $M_\\mathrm{PB}=10M_0 \\sim\\!\\! 10^{23}$\\,g object ($M_0=10^{22}$\\,g being the smallest released dust mass able to trigger a visible avalanche for such a disc, as given by Eq.~\\ref{visi1}). } This makes it a rather unlikely event, although it cannot be completely ruled out. Nevertheless, slightly denser discs (i.e., higher $\\tau_{\\bot}$ and $\\tau_{\\|}$) can easily raise $P_\\mathrm{obs}$ up to 1. As a matter of fact, the dependence on $\\tau_{\\|}$ is so sharp that $P_\\mathrm{obs}=1$ is obtained for $\\tau_{\\|}\\simeq 2.1\\tau_{\\|,nom} \\simeq 0.046$. We thus see that a $\\beta$\\,Pic-like system is below, but not too far from the limit for which chances of witnessing an avalanche are high, especially when considering the uncertainties regarding avalanche strength due to its dependence on several poorly constrained parameters related to the collision-outcome prescription (also keeping in mind that higher $\\tau_{\\|}$ values could alternatively be achieved for a thinner disc of the same dust mass (see Sect.~\\ref{sec:dd})). { Moreover, our $L_\\mathrm{av}/L_\\mathrm{d}>1$ criterion for observability is probably too conservative, and avalanche-induced patterns might be detectable for lower luminosity excess values. Taking, for example, $L_\\mathrm{av}/L_\\mathrm{d}>0.1$ would raise the detection probability to $\\simeq 45\\%$ for a $\\beta$\\,Pic-like system}. { \\subsection{Avalanches in observed systems, perspectives} We defer a detailed application of our model to specific circumstellar discs to a future study. However, the present results can already give a good idea of the typical profile for a ``good'' avalanche-system candidate Our numerical exploration has shown that structures that are the most likely to be associated with avalanche-events have two-sided asymmetry for discs viewed edge-on and open spiral patterns for discs viewed pole-on or at intermediate inclinations. An additional requirement is that these discs should be dust-rich systems, with a dustiness at least equal to, and preferably higher than that of $\\beta$-Pic. Note also that our model makes an additional prediction, i.e., that avalanche affected regions should consist of grains significantly smaller than the \"field\" particles in the rest of the disc. If the blow-out radius of grains is of the order of the wavelength of the observed light, then this should translate into color differences between avalanche (bluer) and non-avalanche (redder) regions. In this respect, one good edge-on candidate might be the recently discovered HD\\,32297 system, which exhibits a strong two-sided asymmetry. As reported by \\citet{Schneider05} and \\citet{Kalas05}, this system is a $\\beta$\\,Pic analog with its SW side significantly brighter than the NE one within $\\simeq$ 100\\,AU \\citep{Schneider05} and possibly outside 500\\,AU \\citep{Kalas05}. Such a two-sided asymmetry would be compatible with the ones obtained in our simulations (as shown for example in the bottom panel of Fig.~\\ref{fig:ass_const}). Furthermore, \\citet{Kalas05} also reported a color asymmetry between the two sides, with the brighter one (SW) being significantly bluer. This seems to indicate that this side is made of smaller, possibly submicron grains \\citep{Kalas05}. As previously discussed, this is what should be expected for an avalanche-affected region. However, an alternative scenario, like the collision with a clump of interstellar medium proposed by \\citep{Kalas05}, might also explain the HD\\,32287 disc structure. Future imaging and spectroscopic observations are probably needed before reaching any definitive conclusions. Among all head-on observed systems, the one displaying the most avalanche-like structure is without doubt HD\\,141569 \\citep{ClampinKrist03}, with its pronounced spiral pattern. Furthermore, the disc's mass, significantly higher than $\\beta$-Pictoris, makes it a perfect candidate in terms of witnessing probabilities. Of course, avalanche is not the only possible scenario here, and several alternative explanations, like an eccentric bound planet or stellar companion, or a stellar flyby have already been proposed \\citep[e.g.,][]{Au04, Wyatt05, Ardila05}. One should, however, be aware that this system is strictly speaking not a \"standard\" debris disc as defined by \\citet{Lagrange00} and as considered in the present study. Indeed, several studies seem to suggest the presence of large amounts of primordial gas \\citep{Zuckerman95,Ardila05}. Gas drag effects have been left out of the present study on purpose, mainly because, in the strict sense of the term, debris discs are systems where dust dynamics is not dominated by gas friction \\citep{Lagrange00}. Moreover, the correct description of dust-gas coupling adds several additional free parameters (gas density and temperature distributions, etc.) and requires a full 2-D or 3-D treatment of gas by far exceeding the scope of the present paper. However, the issue of avalanches in a gaseous medium might be a crucial one for those systems that are most favorable for avalanches, i.e., discs more dusty than $\\beta$-Pic, a system which is already at the upper end of debris-discs in terms of dustiness \\citep[e.g.,][]{Spangler01}. Such more massive systems should fall into a loosely defined category of \"transition\" discs between T-Tauri or Herbig Ae protoplanetary systems and \"proper\" debris discs \\citep[see for example Sect.~4 of][]{Dutrey04}. For such systems (of which HD141569 is a typical example), which are younger than the more evolved debris discs, risks (or chances) of encountering large amounts of remaining gas are high. This crucial issue will be the subject of a forthcoming paper.}" }, "0609/astro-ph0609814_arXiv.txt": { "abstract": "The first three years of observation of the {\\it Wilkinson Microwave Anisotropy Probe} (WMAP) have provided the most precise data on the anisotropies of the cosmic microwave background (CMB) to date. We investigate the impact of these results and their combination with data from other astrophysical probes on cosmological models with a dynamical dark energy component. By considering a wide range of such models, we find that the constraints on dynamical dark energy are significantly improved compared to the first year data. ", "introduction": "Observations of type Ia supernovae (SNe Ia) \\cite{Astier:2005qq,Riess:2004nr}, structure formation (LSS) \\cite{Percival:2001hw,Tegmark:2003ud} and the cosmic microwave background (CMB) \\cite{Spergel:2006hy,Readhead:2004gy,Goldstein:2002gf} all agree on an accelerated expansion of our Universe. This rather unexpected phenomenon can be explained by modifying 4-D gravity \\cite{Dvali:2000hr,Bekenstein:2004ne} or adding a new component to the total energy momentum tensor. The simplest such component is a cosmological constant. It fits all current observations flawlessly and has a simple interpretation in terms of a vacuum energy. Yet, its observed value is 120 orders of magnitude off from the naive estimate $\\Lambda \\sim \\mplank^4$, where $\\mplank$ is the reduced Plank mass. The coincidence between this minute dark energy contribution and the observed energy density of matter is rather puzzling. If not given by chance via some sort of anthropic principle, it necessitates a mechanism that explains this coincidence. An immediate possibility is a coupling \\cite{Wetterich:1994bg,Amendola:1999er,Amendola:2006qi} between (dark) matter and dark energy (though there might be problems due to quantum effects \\cite{Doran:2002bc}). Another solution is an attractor behavior \\cite{Wetterich:fm,Ratra:1987rm,Caldwell:1997ii} of dark energy that leads to an almost constant ratio between the fractional energy density $\\od(z)$ of dark energy and the species otherwise dominating the expansion, i.e. photons and neutrinos during radiation domination and matter during matter domination. Coincidentally, such an attractor behavior corresponds to a scalar field with exponential potential that arises in string theories and when solving the cosmological constant problem from the point of view of dilatation symmetry \\cite{Wetterich:fm}. The non-vanishing $\\od(z)$ at higher redshifts alleviates the problem of explaining the coincidence of matter and dark energy today $\\om^0 \\approx \\od^0$. Instead of fine tuning $\\Lambda$ to many orders of magnitude, the tuning needed is of the order of $10^{-3}$. However, the tuning needed for such \\emph{early dark energy} cosmologies increases the less dark energy there is at earlier times. A detection of early dark energy, on the other hand, would give crucial hints to fundamental laws of nature. The aim of this study therefore is to investigate the implications of the three year data of WMAP on dynamical dark energy models in general and their respective fractions of early dark energy. In view of the theoretical uncertainties many different techniques have been employed in the analysis of the dark energy, ranging from atttempts to reconstruct the potential of a scalar field dark energy (e.g. Ref. \\cite{Sahlen:2006dn}) to the principal component approach of Ref. \\cite{Huterer:2002hy}. We consider the redshift dependence of the fractional dark energy $\\od(z)$ as a free function to be ``measured'' by observation. We investigate in this note various parameterizations and an interpolated model. The possible coupling between dark energy and dark matter is neglected in this study. ", "conclusions": "The three year data of WMAP was used to estimate cosmological parameters for a wide range of dynamical dark energy models. We have shown that this new data in combination with large scale structure data constrain the average amount of dark energy during the time of structure formation to $\\osf \\lesssim 4\\%$ for a fair sample of dark energy models from the literature. We have also constructed a parameterization of $\\od(a)$ which linearly interpolates between the dark energy fraction at several redshift bins. This allows for a considerably higher fraction of $\\osf$. The analysis also shows that the values of the standard cosmological parameters for the dynamical dark energy models are well compatible with the values found by WMAP for \\lcdm. The effect of dynamical dark energy on $\\sigma_8$ and on the formation of nonlinear structure offer promising routes for further constraints on the time evolution of dark energy or a possible falsification of the \\lcdm\\ model." }, "0609/astro-ph0609739_arXiv.txt": { "abstract": "Traditional thermal evolution models of giant planets employ arbitrary initial conditions selected more for computational expediency than physical accuracy. Since the initial conditions are eventually forgotten by the evolving planet, this approach is valid for mature planets, if not young ones. To explore the evolution at young ages of jovian mass planets we have employed model planets created by one implementation of the core accretion mechanism as initial conditions for evolutionary calculations. The luminosities and early cooling rates of young planets are highly sensitive to their internal entropies, which depend on the formation mechanism and are highly model dependent. As a result of the accretion shock through which most of the planetary mass is processed, we find lower initial internal entropies than commonly assumed in published evolution tracks. Consequently young jovian planets are smaller, cooler, and several to 100 times less luminous than predicted by earlier models. Furthermore the time interval during which the young jupiters are fainter than expected depends on the mass of planet. Jupiter mass planets ($1\\,\\rm M_J$) align with the conventional model luminosity in as little at 20 million years, but $10\\,\\rm M_J$ planets can take up to 1 billion years to match commonly cited luminosities, given our implementation of the core accretion mechanism. If our assumptions, especially including our treatment of the accretion shock, are correct and if extrasolar jovian planets indeed form with low entropy, then young jovian planets are substantially fainter at young ages than currently believed. Furthermore early evolution tracks should be regarded as uncertain for much longer than the commonly quoted $10^6$ years. These results have important consequences both for detection strategies and for assigning masses to young jovian planets based on observed luminosities. ", "introduction": "In the past decade, a number of nearby star associations have been recognized as being quite young, less than $10\\,\\rm Myr$ old (e.g. IC 348, TW Hydrae, MBM12, $\\eta$ Cha \\citep{Lada95, Webb99, Luhman04}). Such associations are likely well stocked with recently formed, presumably bright giant planets that should in principle be easy prey for a variety of planet detection technologies. Planning for the hunt, however, requires knowledge of the expected luminosity of young giant planets as a function of time since their formation, particularly at young ages when they are presumably easy game. While models of the luminosity evolution of giant planets have a long pedigree \\citep[e.g.,][]{Grossman72, Graboske75}, the early work focused on the evolution of the solar system giants, attempting to explain their current luminosity at an age of 4.5 Gyr. Since planets lose memory of their initial conditions over time, initial conditions were selected more for computational convenience than for accuracy. Many improvements have subsequently been made to the models, particularly in the characterization of jovian atmospheres at various effective temperatures, although essentially the same initial conditions are still employed \\citep{Hubbard80a, Burrows97, Chabrier00b}. The standard evolution model begins with a hydrogen-helium sphere having a large radius, high internal entropy, and large effective temperature. Such an object is not necessarily one that would be the result of any particular planet formation model. This model planet is allowed to radiate and cool over time. Since there have been no detections yet of young planets with measured masses, the applicability of this initial condition is untested, although there are data from more massive objects. A pair of eclipsing, young ($\\sim 1\\,\\rm Myr$) brown dwarfs with known dynamical masses (57 and 36 Jupiter masses ($\\rm M_J$)) indeed have radii exceeding five times that of Jupiter \\citep{Stassun06}, confirming that young massive brown dwarfs, at least, are in fact large and hot. But giant planets that formed in a disk around a primary star, or even isolated planet-mass objects, may have experienced very different initial conditions. Evolution tracks computed in the usual way--even though they do not necessarily reflect any particular planet formation theory--have been used to evaluate detection strategies for true giant planets orbiting solar type stars \\citep[e.g.,][]{Burrows05}, and they have been used to characterize isolated, very low mass brown dwarfs. \\citet{Chauvin04}, for example, reported on the detection of a faint companion to the M8 brown dwarf \\object{2MASSWJ1207334-393254} (hereafter 2MASS1207) in TW Hydrae with an estimated age of 8 Myr. Using its observed luminosity and applying published evolution tracks, they estimated a mass of just $5\\,\\rm M_J$ for the mass of the companion. The early modelers certainly did not foresee that direct detections of putative young planets would be compared against the models at exceptionally young ages, at times when the model planet may not yet have forgotten its hot start. \\citet{stevenson82} wrote that evolution calculations ``...cannot be expected to provide accurate information on the first $10^5$--$10^8$ years of evolution because of the artificiality of an initially adiabatic, homologously contracting{\\footnote{See \\citet{Stahler88} for a discussion of homologous contraction.}} state.'' More recently, \\citet{Baraffe02} examined the uncertainties in evolution tracks of brown dwarfs at young ages and cautioned about the applicability of evolution models at ages less than a few million years, on the lower end of Stevenson's uncertainty range. \\citet{Wuchterl05} has also expressed concern that standard evolution models do not capture the early evolution correctly. Given the clear imperatives to interpret observations of young, low mass objects and to plan for future direct detections of giant planets formed in orbit about solar type stars, there is a need to connect models of giant planet formation to giant planet evolution. Our goals here are both to help fill the void in physically plausible models of extrasolar giant planets (EGPs) at young ages and to better quantify the age beyond which the evolution models are robust and applicable. We aim to understand whether or not the current generation of evolution models can reliably predict the luminosity of giant planets at young ages and, if not, then define the age beyond which current models are reliable. Instead of using an arbitrary starting condition, we employ planets formed by one implementation of the core accretion model. In this scenario gas giant planets form by rapid accretion of gas onto a solid core that grew by accretion of planetesimals in the nebula. This mechanism is one of two competing scenarios of gas giant formation, the other being the gas instability model by which giants form from a local disk instability \\citep{Boss98} that creates a self-gravitating clump of gas. The planet resulting from such a clump could also be used as the starting point of an evolutionary calculation (see \\citet{Bodenheimer74, Bodenheimer76, Bodenheimer80}), but we choose here to focus solely on the core accretion mechanism as it currently seems the more promising mechanism for explaining the formation of the giant planets (see \\citet{Lissauer06} for a review). For specificity, we rely upon the implementation by \\citet{Hubickyj05} of the core accretion mechanism. By necessity, their work makes a host of assumptions that ultimately affect the properties of newly born giant planets. As we will demonstrate, following the end of accretion this model predicts that giant planets are substantially fainter than standard evolution models. While we believe that this conclusion is secure, we stress that the precise numerical value of the post-accretion luminosity depends upon the particular assumptions employed by \\citet{Hubickyj05}. Therefore, we first briefly review this model and highlight the assumptions upon which the work rests in Section 2. We describe our method of evolving these model planets over time in Section 3 and compare our results with standard giant planet evolution models. We find in Section 4 that the initial conditions influence subsequent planetary evolution for longer than generally appreciated and that planets formed by the \\citet{Hubickyj05} recipe for core accretion are substantially fainter than standard models have previously predicted. We end by cautioning those who wish to rely upon evolution models to characterize detected young giant planets that may have grown by the core accretion mechanism would be wise to be judicious in their estimation of the model-dependent uncertainties. ", "conclusions": "We have computed the first giant planet evolution models that couple planetary thermal evolution to the predicted core mass and thermal structure of a core accretion planet formation model. \\citet{Baraffe06} investigated the evolution of planets with core sizes and heavy element abundances derived from the core accretion models of \\citet{Alibert05}. However, \\citet{Baraffe06} did not attempt to match the thermal structure (and hence, temperature, entropy, and density) at the interface between planetary formation and subsequent evolution. Our implementation of the core accretion model processes most of the planetary mass through an accretion shock in which the accreting gas loses most of its internal entropy. As a result, our young giant planets are cooler, smaller, fainter, and take longer to evolve than the standard hot-start model giant planets. We note, however, that our accretion model does not resolve the radiative transfer within the shock, but rather uses the shock boundary conditions of \\citet{stahler80}. A more complete or detailed treatment of accretion at the surface of the planet could very well result in different initial conditions, including possibly a warmer, larger, brighter and more conventional young planets. Specifically, a three-dimensional hydrodynamic simulation of gas accretion by giant planets, allowing for material accreted through a circumplanetary disk and shock radiation, would provide more rigorous post-formation models for subsequent evolution calculations. Until such models are available, our approach--which likely provides a lower limit to the post accretion luminosity--demonstrates that plausible initial conditions can lead to very different early evolution tracks for giant planets than the `hot start' models that are commonly relied upon. For example, at $10^7$ years--a time greater than the age of the TW Hydrae association--our $10\\,\\rm M_J$ core accreted planet is more than a factor of 100, or 5 magnitudes, fainter than the equivalent hot start planet. The luminosity difference falls with decreasing mass, so that our model luminosity for a $1\\,\\rm M_J$ planet is comparable to the standard case. Thus the thermal luminosity of young, massive giant planets, which have been assumed to be easy targets for coronagraphy, may be substantially less than previously assumed. If this result is correct, then searches for the thermal emission from young, several Jupiter mass planets must be far more sensitive than previously anticipated in order to detect these relatively faint, young planets. Thermal infrared planet searches by the Large Binocular Telescope, the James Webb Space Telescope, and other planned telescopes would all be impacted, although efforts to detect planets in reflected light would not. This conclusion holds true even to ages as great as that of the Pleiades for the most massive planets considered here. Ironically the least massive, most intrinsically faint planets (1 to $2\\,\\rm M_J$) match their hot-start luminosity tracks by just a few tens of millions of years, or less and are much less underluminous before that time. Direct detections of young giant planets with dynamically measured masses will test this conclusion. Since the numerical values for luminosity that we derive depend primarily upon our treatment of energy radiated from the accretion shock, these results should not be viewed as specific predictions of the core accretion model. Rather our point is that core accretion naturally leads to gas accretion through a shock, which may result in low entropy planets. The viability of giant planet formation via core accretion depends on physical processes happening earlier in the accretionary process (at smaller masses) than those processes that we have shown to be crucial for the luminosity of young planets of Jupiter's mass and larger. We note in passing that the faintness predicted for young Jupiter-mass planets compounds with non-equilibrium chemistry to make detection of young giant planets at M band particularly challenging. \\citet{Marley96}, after the discovery of \\object{Gl229B}, suggested that a substantial M-band 4 - 5 $\\mu$m flux peak should be a universal feature of giant planets and brown dwarfs. In addition to the intrinsic emergent flux, this spectral range has looked promising for planet detection due to the favorable planet/star flux ratio \\citep[e.g.][]{Burrows05}. However, it has been known since the 1970s \\citep[see][]{Prinn77} that Jupiter's 5 $\\mu$m flux is suppressed by absorption by CO present in amounts exceeding that predicted by equilibrium chemistry. This same effect has now been observed in brown dwarf M-band photometry \\citep{Golim04, Leggett06}, as anticipated by \\citet{Fegley96}. Excess CO leads to strong absorption at 4.5 $\\mu$m, leading to diminished flux in M-band \\citep{Saumon03}. This effect further suppresses the M band fluxes of young planets below the existing models. Taken together, fainter young planets and reduced M-band flux may well reduce the catch from what had seemed a promising fishing hole for direct planet detection. We also conclude that the predicted evolution of giant planets objects, regardless of formation mechanism, is far more sensitive to the precise conditions at the termination of accretion than has been previously recognized. Most workers have assumed that evolving model planets `forget' their initial conditions within $10^6$ \\citep{Baraffe03b} to $10^8$ years \\citep{stevenson82} of the first time step. While 1 to 2 Jupiter mass planets do have a short ($\\sim10^7\\,\\rm yr$) memories, we have shown that more massive planets remember their initial thermal state far longer. The evolution time scale for young, hot planets depends exponentially on their initial entropy. Until the initial thermal state of young, low mass objects--even isolated brown dwarfs--is known with more certainty, the early evolution tracks must be regarded with some skepticism. Any effort to assign a mass to a very young putative giant planet must consider the uncertainties in these early evolutionary tracks." }, "0609/astro-ph0609213_arXiv.txt": { "abstract": "Roughly $80\\%$ of Ultraluminous Infrared Galaxies (ULIRGs) show blue shifted absorption in the resonance lines of neutral sodium, indicating that cool winds are common in such objects, as shown by Rupke et al and by Martin. The neutral sodium (NaI) columns indicated by these absorption lines are $\\sim10^{13}-3\\times10^{14}\\cm^{-2}$, while the bolometric luminosity varies by a factor of only four. We show that the gas in ULIRG outflows is likely to be in photoionization equilibrium. The very small ULIRG sample of Goldader et al. demonstrates that the ratio of ultraviolet flux to far infrared flux varies by a factor $\\sim100$ from object to object. While the Goldader sample does not overlap with those of Rupke et al. and Martin, we show that such a large variation in ultraviolet flux will produce a similar variation in the column of neutral sodium for a fixed mass flux and density. However, if the cold gas is in pressure equilibrium with a hot outflow with a mass loss rate similar to the star formation rate, the range of ionization state is significantly smaller. Measurements of the UV flux for objects in the Martin and Rupke et al. catalogs will definitively determine if photoionization effects are responsible for the wide variation seen in the sodium columns. If they are, a determination of the gas density and mass loss rate in the cool winds will follow, with attendant improvements in our understanding of wind driving mechanisms and of the effects of galaxies on their surroundings. ", "introduction": "Recent observations of the sodium resonance line (NaD) doublet in Ultra-luminous Infra-red Galaxies (ULIRGS) by Rupke and coworkers \\cite{2002ApJ...570..588R,2005ApJ...631L..37R,2005ApJS..160...87R,2005ApJS..160..115R} and by Martin (2005; 2006) show that cool outflows from ULIRGS are common. Martin (2005) finds that 15 of 18 ULIRGS ($83\\%$) possess such flows, while Rupke et al. (2005b) find a detection rate of $80\\pm7\\%$ for 30 local ULIRGs. These results indicate that the outflow emerges in most directions. The absorption troughs, which typically extend over $\\sim300\\kms$, are not black, which at first blush might suggest that the outflows are optically thin in the NaD line. However, the doublet line ratio, which is equal to 2:1 in optically thin gas, rarely has that ratio in Martin's sample, while Rupke et al. (2005a) find optical depths for the weaker of the doublet lines that range from $0.06$ to $7$, with an average of $1.5$. This finding shows that some sight lines to ULIRG galaxies are optically thick in the NaD line, that the optically thick outflow covers only about $25-30\\%$ of the (optically emitting) galaxy, but that this optically thick component is seen toward $\\sim80\\%$ of ULIRGS. Given the patchy nature of the interstellar medium in most galaxies, the last finding is not entirely surprising. However, another implication of the observations is very surprising. As Figure \\ref{Fig:column} shows, the mean column averaged over both samples (Rupke et al. 2005a; Martin 2006) is $13.7$ in the log, with a standard deviation of $0.66$ dex, or from $10^{13}\\cm^{-2}$ to $3\\times10^{14}\\cm^{-2}$. The actual Na I columns vary from $7\\times10^{12}\\cm^{-2}$ to $5\\times10^{14}\\cm^{-2}$ in the sample of \\cite{2005ApJS..160...87R}. Since the luminosity, and star formation rate of the galaxies in both samples varies by only a factor of about 4 (the standard deviation is a factor of 2), one might expect the $N_H$ column and hence the $N_{NaI}$ column toward the galaxies to vary by a similar factor; in simple galactic wind models the mass outflow rate increases with increasing star formation rate. The Na I column can be used to find the mass loss rate from each galaxy, in principle. For example, in a smooth flow, the density is related to the mass loss rate by \\be \\label{eq:mdot}% \\dot M_w = \\Omega r^2\\mu(r) n(r) v(r), \\ee % where $\\mu(r)$ is the mean molecular weight, $n(r)$ is the number density of hydrogen, $v(r)$ is the velocity of the flow at a distance $r$ from the center of the galaxy, and $\\Omega\\le4\\pi$ is the global wind cover factor in steradians. For simplicity we employ a spherical model, although ULIRGs appear to form a substantial fraction of their stars in a kiloparsec (or smaller) scale disk \\cite{1991ApJ...378...65C}. The hydrogen column through such a wind is \\be % N_H = {\\dot M_w\\over \\Omega}\\int_{r_0}^\\infty {dr\\over r^2 \\mu(r) v(r)}. \\ee % A very rough estimate is found by taking $\\mu(r)=m_p$ and $v(r)=v_\\infty$, where the latter is the terminal velocity of the outflow, which in the spirit of approximation adopted here is taken to be the maximum observed blue shift: \\be \\label{eq:column}% N_H\\approx{\\dot M_w\\over\\Omega m_p v_\\infty r_0}, \\ee % where $r_0$ is the radius at which the absorption line forms. Solving for the mass loss rate in terms (as far as possible) of observed quantities, \\be % \\dot M_w \\approx \\Omega m_p r_0 v_\\infty N_{NaI} \\left({N_{Na}\\over N_{NaI}}\\right) \\left({N_H\\over N_{Na}}\\right)d_{Na}, \\ee % where $N_{NaI}$ is the (observed) column of gas phase neutral sodium, $N_{Na}$ the total gas phase sodium column, $N_H$ is the total hydrogen column, and $d_{Na}$ is the ratio of total to gas phase Na; it is larger than unity since much of the Na is locked up in dust grains. Neither of the two terms in parentheses nor $d_{Na}$ are measured, nor is $r_0$ well constrained. For solar abundances $N_H/N_{Na}=4.9\\times10^5$ \\cite{AG}, while \\cite{1996ARA&A..34..279S} find that the gas phase abundance of Na in the cool gas of the diffuse cloud toward $\\zeta$ Ophiuci is a factor of $d_{Na}=8.9$ smaller than the solar abundance of Na. There is evidence of dust in ULIRG outflows; first, NaD is a resonance line, so that it scatters continuum photons rather than destroying them. Since the winds are nearly spherical (they are seen toward 80\\% of ULIRGS), any photons removed from our line of sight by the NaD transition should appear along some other line of sight as redshifted emission; since we see no such emission, some other mechanism is removing the scattered photons, absorption by dust being the number one suspect. Second, there is a correlation seen between reddening and NaD absorption toward luminous infrared galaxies, again suggesting dust in the outflow \\cite{1995ApJS...98..171V}; \\cite{2000ApJS..129..493H}. In the rest of this paper we will adopt solar metalicities and the Milky Way ISM $d_{Na}$ just quoted. Using these values, the typical observed column $NaI=10^{14}\\cm^{-2}$ would correspond to a hydrogen column $N_H\\approx10^{21}\\cm^{-2}$ if somewhat less than half the gas-phase sodium were neutral. Any local heating, such as might be produced by shocks, will lead to lower depletion levels, but by less than the factor of ten corresponding to all the grains being destroyed, given the evidence for dust cited above. The force responsible for expelling the wind is uncertain. One candidate is ram pressure from outflowing supernova-heated gas, which will entrain cold gas from the interstellar medium of the ULIRG. A variant on this is ram pressure from hot gas produced by an accreting massive black hole (an active galactic nucleus, or AGN). The Compton temperature of both Seyfert and Quasar nuclei is around $10^7$K, similar to that from supernova heating. Momentum driving from supernovae can be significant, if most supernovae explode while surrounded by cold gas. Another candidate for momentum driving is radiation pressure on dust grains embedded in the cold gas. The radiation could arise either from stars or from a central black hole. The radiation pressure from a black hole can exceed that from a starburst. Whether such a centrally driven outflow can produce an outflow that will be detected from $80\\%$ of the sky is far from certain. In fact it seems likely that both energy and momentum driving operate. ULIRGS are enveloped in x-ray emitting gas, Arp 220 being an excellent example \\cite{2003ApJ...591..154M,2002ApJ...581..974C}; the presence of hot gas is a necessary (but not sufficient) condition for energy driving of a cool outflow. As we have just noted, the winds are most likely dusty. The wind optical depth to optical or UV emission is likely substantial, while the galaxies themselves are optically thick even to the far infrared. The outflow may well be multiphase; the questions to be addressed revolve around which phase carries more mass, momentum, energy, and metals into the surrounding intergalactic medium. To answer such questions, we would like to measure the mass loss rate in cold gas. If the NaI columns are taken at face value, and the ratio $N_H/N_{NaI}$ is assumed to be constant, the mass loss rates in cold gas vary from ULIRG to ULIRG by a factor of $\\sim50$. Is this really true? There are numerous possible reasons for the NaI column seen in galactic outflows to vary from object to object. The mass loss rate will vary from object to object, so that the total column of hydrogen, and hence Na, will vary, even at a fixed metallicity. The metallicity of the outflowing gas may vary from object to object. Both these properties are likely to depend on the star formation rate (and history); the star formation rate, at least, varies over only a small range for the samples we are discussing. A more likely cause of variation in NaI column is a variation in either or both the depletion on to dust grains and the ionization state of Na. Neutral sodium is a rather delicate atom, easily ionized by ultraviolet radiation. The wind may be irradiated either by starlight from the host galaxy, or by shocks in the wind. In this paper we argue that the outflow is in photoionization equilibrium, except possibly for the most UV dim systems, and then show that if Martin and Rupke et al.'s galaxies have spectral energy distributions like those of other nearby ULIRGs that have been observed in the ultraviolet, then it may be that the mass loss rates are more nearly equal than the observed NaD columns imply; unless the neutral gas is very dense, $n\\sim 10^5\\cm^{-3}$, the ionization fraction, or the ratio of neutral to total sodium, $(NaI/Na)$, will vary by a large factor from object to object. This follows from the observed variation of $\\sim100$ in the ratio of UV to FIR flux seen in nearby ULIRGs (Goldader et al. 2002), and simple photoionization calculations, as we show below. ", "conclusions": "The mass loss rates of ultraluminous infrared galaxies, as determined by observations of the NaD lines, are believed to be substantial---of order one fifth of the star formation rate. This estimate assumes that the ionization fraction of sodium is $\\sim0.1$. The gross physical properties of ULIRGs, including their stellar masses, gas masses, sizes, and star formation rates vary by factors of only a few, so at first glance the assumption of fixed NaI ionization fraction appears reasonable. However, one property of ULIRGS that appears, based only on a few galaxies, to vary by a large amount is the ultraviolet flux. We have shown that cool ULIRG winds are likely to be in photoionization equilibrium. We used a simple argument to show that the large object-to-object UV flux variations seen in the small number of galaxies that have been observed in the ultraviolet would be enough to produce similarly large variations in NaI column density, if the density of the cool gas is $n\\lesssim10^5(500\\pc/r)^2\\cm^{-3}$. More complete photoionization calculations using CLOUDY produce the same magnitude of variation in NaI/Na with variation in UV flux. They also show, for these moderate densities, that the MgII column should not vary appreciably with changes in UV flux. If, however, the cold gas is in pressure equilibrium with a hot wind having a mass loss rate similar to the star formation rate of a typical ULIRG, the cool gas will have a density $n\\gtrsim 10^6(500\\pc/r)^2)\\cm^{-3}$. In that case the ionization parameter is $U_{NaI}<10^{-3}$, and there will be little variation in NaI column from object to object. As a check, the (so far unobserved) MgII columns are predicted to show variations that are anti-correlated with the UV/FIR ratio. We conclude that the observed factor of $\\sim100$ variation in the UV/FIR flux density ratios in ULIRGs with similar bolometric luminosities is consistent with the factor of $\\sim40$ variations in the column density of Na I outflows seen by Martin (2006) and \\cite{2005ApJS..160...87R}; the different UV fluxes may well be the cause of the variation seen in the NaI columns, if the cool gas has a density $n\\lesssim 10^5\\,\\cm^{-3}$. A tight correlation between Na I column and UV flux would indicate that the cool gas is not accelerated by a hot wind. The origin of the object-to-object variation in the UV/FIR ratio is not clear. The known correlation between reddening and NaI column in lower luminosity galaxies suggests that dust in the outflow may play a role. However, the NaI optical depth toward luminous infrared galaxies in the sample of Rupke et al. (2005b) actually exceeds that toward their (more luminous) ULIRGs. It is unlikely that the UV/FIR ratio is smaller in the less luminous objects. Thus dust absorption in the outflow is probably not the sole reason for the low UV fluxes in ULIRGS. Whether the galaxies studied by Martin (2005) and by Rupke et al. (2005a,2005b) actually have large object-to-object variations in their near-UV fluxes is not known. We note that the results of Goldader et al. (2002) show that near IR or even optical flux measurements are not a good proxy for the UV flux; see the data for VV114, IRAS 15250+3609, and Mrk 273 in Figure \\ref{Fig:UV}. This demonstrates that optical observations will not be sufficient to answer this question. Observations by GALEX or by the ACS on the Hubble Space Telescope could answer this question definitively." }, "0609/astro-ph0609449_arXiv.txt": { "abstract": "We report on the detection in Sloan Digital Sky Survey data of at least three, roughly parallel components in a $65\\arcdeg$-long, stellar stream complex previously identified with the Anticenter or Monoceros Ring. The three-stream complex varies in width from $4\\arcdeg$ to $6\\arcdeg$ along its length and appears to be made up of two or more narrow substreams as well as a broader, diffuse component. The width and complexity of the stream indicate that the progenitor was likely a dwarf galaxy of significant size and mass. The stream is 8.9 kpc distant and is oriented almost perpendicularly to our line of sight. The visible portion of the stream does not pass near any known dwarf galaxies and a preliminary orbit does not point to any viable progenitor candidates. Orbits for the narrower substreams can be modeled with velocity offsets from the broad component of $\\approx 8$ km s$^{-1}$. We suggest that the broad component is likely to be the remains of a dwarf galaxy, while the narrower streams constitute the remnants of dynamically distinct components which may have included a native population of globular clusters. While the color of the main sequence turn-off is not unlike that for the Monoceros Ring, neither the visible stream nor any reasonable projection of its orbit passes through Monoceros or Canis Major, and we conclude that this stream is probably unrelated to the overdensities found in these regions. ", "introduction": "The value of large scale digital sky surveys to studies of Galactic structure and the Local Group has become abundantly clear in recent years, and particularly in the last few months. In addition to the large scale features attributed to past galaxy accretion events \\citep{yann03,maje2003,roch04}, Sloan Digital Sky Survey (SDSS) data were used to detect the remarkably strong tidal tails of Palomar 5 \\citep{oden2001,rock2002, oden2003, grill2006b}, NGC 5466 \\citep{belo2006a, grill2006a}, and streams due to extant or extinct globular clusters \\citep{grill2006c} and dwarf galaxies \\citep{belo2006b, grill2006d}. \\citet{will2005}, \\citet{zuck2006}, and \\citet{belo2006c} recently used SDSS data to discover several new dwarf satellites of the Milky Way. In this paper we continue our analysis of the SDSS database to search for extended structures in the Galactic halo. We briefly describe our analysis in Section \\ref{analysis}. We discuss the structure of a stream complex, attributed to the Monoceros Ring by \\citet{belo2006b} in Section \\ref{discussion}, estimate distances in Section \\ref{distance}, and put constraints on the orbit in Section \\ref{orbit}. ", "conclusions": "} Apparent in Figure 1 is a broad and complex stream running from north to south across the field. Portions of this stream are visible in \\citet{newberg02} and \\citet{belo2006b}. The narrow, curved stream running towards the northeast is part of the 63$\\arcdeg$-long globular cluster stream found by \\citet{grill2006c}, hereafter referred to as GD-1. The broad, east-west stream just above the main gap in the data is the Sagittarius stream discussed by \\citet{belo2006b}, though somewhat muted by our filtering due to its greater distance. The stream complex of interest extends from (R.A., decl.) = ($126.4\\arcdeg, -0.7\\arcdeg$) to (R.A., decl.) = ($133.9\\arcdeg, 64.2\\arcdeg$), and runs in a $65\\arcdeg$, nearly great circle path from Ursa Major in the north to Hydra in the south. The stream is truncated at both the southern and northern ends by the limits of the available data. We note that there is also an apparent concentration of stars at (R.A., decl.) = ($134\\arcdeg, 3.4\\arcdeg$), surrounded by faint, banded substructure roughly parallel to that of the main western stream. From the reddening map of \\citet{schleg98}, the maximum values of $E(B-V)$ are $\\approx 0.2$ (near the northern tip), with typical values near 0.03 along the remainder of the stream. There are diminutions here and there in the stream that could be attributed to regions of higher reddening, but there are no long features with a north-south orientation which could be held to account for either the appearance of the stream as a whole, or for the different components within it. Sampling at several representative points, the main stream complex appears to be about $5\\arcdeg$ wide on average. This is significantly broader than the globular cluster streams found by \\citet{oden2003, grill2006a, grill2006b} and \\citet{grill2006c}. $5\\arcdeg$ corresponds to about 800 pc at our estimated distance to the stream (see below), which is much larger than the tidal diameters of globular clusters. We conclude that the progenitor was considerably more extended than a globular cluster and was most likely a dwarf galaxy. The stream is clearly not just a broad swath of stars. Rather, it appears to be made up of a $\\sim 2\\arcdeg$ wide broad component running roughly along the center of the stream complex, and at least two narrower, $1\\arcdeg$-wide streams to the east and west of the broad component. The fine structure in the stream complex is illustrated in Figure 2, where we have made several east-west slices across the complex at various declinations. Comparing Figures 1 and 2, there are indications of additional structure within the broad component, and still other, more tenuous parallel streams to the east and west of the three major components. Integrating the background subtracted, weighted star counts along the stream over a width of $\\approx 5\\arcdeg$ we find that the total number of stars in the discernible stream down to $g = 22.5$ is $9200 \\pm 1500$. The mean surface density of stars in the southern portion of the stream ($-1\\arcdeg < \\delta < 14\\arcdeg$) is about 67 stars deg$^{-2}$, which is roughly twice the 32 stars deg$^{-2}$ (corrected for $cos(\\delta)$) found in the northern section ($39\\arcdeg < \\delta < 52\\arcdeg$). The highest local surface densities exceed 200 stars deg$^{-2}$. If we assume a globular cluster-like luminosity function in the stream, then we can use the color transformation equations of \\citet{smith2002}, the deep M 4 luminosity function of \\citet{richer2002}, and the mass-luminosity relation of \\citet{baraffe97} to extrapolate to fainter magnitudes. Integrating over $3 < M_V < 17$, we estimate a total number of stars in the visible portion of the stream of $33,000 \\pm 5400$, a total luminosity of $5500 \\pm 900 L_\\odot$, and a total estimated mass of $9300 \\pm 1500 M_\\odot$. \\subsection{Color-Magnitude Distribution and Distance to the Stream \\label{distance}} In Figure 3 we show $g, g - i$ color-magnitude distributions for the stream stars, extracted by generating a Hess diagrams of stars lying along $2\\arcdeg$-wide regions covering the western and eastern halves of the stream complex and subtracting a similar field star distribution sampled over $\\approx 500$ deg$^2$ to the east and west of the stream complex. Despite the somewhat limited statistics, a clear signature of the turn-off and main sequence is evident in the stream population. Moreover, the turn-off regions of the distributions match the dereddened, shifted main sequence locus of M 13 fairly well. The turn-offs in the eastern and western halves of the complex lie at dereddened $g - i = 0.27$ and $0.3$, respectively. A similar diagram in $g - r$ yields dereddened turn-off colors of $g - r = 0.23$ and 0.25, respectively, with the difference most likely a due to $\\sim 0.02$ mag measurement uncertainty. These estimates lie between values of 0.26 and 0.22 measured by \\citet{newberg02} for the Monoceros Ring and the Sagittarius stream, respectively. They are also much bluer than \\citet{newberg02}'s estimate of dereddened $g - r = 0.40$ for thick disk stars. Based solely on turn-off color, we can therefore rule out association of the stream complex with thick disk stars. However, within the uncertainties, the turn-off colors are consistent with those of either the Monoceros Ring or the Sagittarius stream. Varying the magnitude shift applied to M 13's main sequence locus from -1.0 to +3.0 mag, we measured the foreground-subtracted, mean surface density of stream stars in the regions $-1\\arcdeg < \\delta < 9\\arcdeg$, $17\\arcdeg < \\delta < 39\\arcdeg$, and $39\\arcdeg < \\delta < 63\\arcdeg$. To avoid potential problems related to a difference in age between M 13 and the stream stars, we used only the portion of the filter with $19.5 < g < 22.5$, where the bright cutoff is 0.8 mag below M 13's main sequence turn-off. Though this reduces the stream contrast somewhat, it provides sufficient integrated signal-to-noise to enable main sequence fitting. Fitting Gaussians to the mean surface densities as a function of magnitude shift (e.g. \\citet{grill2006c}, we find that the highest contrasts occur for shifts of +0.31, +0.29, and +0.37 mag for the eastern half of the stream complex over the declination ranges given above, respectively. For the western half, we find that the filter response peaks at shifts of +0.34, +0.25, and +0.38 mag, respectively. The magnitude shifts for the two halves are highly consistent with one another, and we conclude that there is no significant distance offset from one side of the complex to the other. Adopting a distance to M 13 of 7.7 kpc \\citep{harris96} we find an average heliocentric distance of $d = 8.9 \\pm 0.2$ kpc. The stream is roughly perpendicular to our line of sight and slightly curved about the Galactic center as expected. Our distance estimate is in excellent agreement with the $\\approx 9$ kpc found by \\citet{ibata03} in their WFS-0801 field, which is situated on the western edge of the stream complex at $\\alpha = 120.5, \\delta = 40.3$. \\subsection{Constraints on the Orbit \\label{orbit}} The visible portion of the stream complex spans the Galactic anticenter direction and, projecting a great circle path, is inclined by $35\\arcdeg$ to the Galactic plane. Though we are currently limited by a lack of velocity information, for a given model of the Galactic potential the progenitor's orbit is actually fairly well constrained by the observed distance and orientation of the stream. Using the Galactic model of \\citet{allen91} (which includes a disk, bulge, and spherical halo, and which \\citet{grill2006a, grill2006b} and \\citet{grill2006c} found to work reasonably well for NGC 5466, Pal 5, and GD-1), we use a least squares method to fit both the orientation on the sky and the distance measurements in Section \\ref{distance}. In addition to a number of normal points lying along the central component of the stream, we chose as a velocity fiducial point a position at the northern end of the stream at (R.A., decl) = (125.463\\arcdeg, 51.492\\arcdeg). If we allow the proper motions to be free ranging and uninteresting parameters, the model which best fits the data predicts $v_{LSR} = -18 \\pm 10$ km s$^{-1}$ at the fiducial point, where the uncertainty corresponds to the 95\\% confidence interval. A projection of this orbit is labeled C in Figure 4. We note that the uncertainty is primarily determined by the large lever arm over which it has been possible to measure relative distances. The 95\\% range in $v_{LSR}$ in turn predicts a range in perigalactic and apogalactic radii of $6.6 < R_p < 6.9$ kpc and $16.8 < R_a < 17.3$ kpc. Of course, these ranges do not take into account uncertainties in the absolute distance of the stream (which depends on the uncertainty in M 13's distance) or of the validity of \\citet{allen91}'s Galactic model. Given the very similar distances estimated for the eastern and western portions of the stream, it is highly unlikely that the complex could be a superposition of multiple wraps around the Galaxy. The two narrower streams (E and W in Figure 4) can be reasonably well modeled by 0.18 $mas$ yr$^{-1}$ offsets in east-west proper motion. At the distance of the stream complex this amounts to $\\approx 8$ km s$^{-1}$. The entire stream complex is therefore likely to be the remains of a dwarf galaxy which contained distinct components spanning a range of binding energies. The narrower streams might, for example, be the remnants of the parent galaxy's globular cluster population. Piecing together the original structure and evolution of the stream's progenitor will require detailed N-body modeling. Integrating orbits for parameter sets spanning the range above, we find that, with the exception of the Sagittarius dE, there are no known dwarf galaxies within $5\\arcdeg$ of the projected orbit. The Sagittarius dE lies $3.1\\arcdeg$ from the projection of the best-fit orbit, but the orbital planes of the the Sagittarius dE and the new stream are clearly distinct (Figure 4); we attribute this apparent proximity to the expected confluence of orbit projections in the direction of the Galactic center and not to any physical association between them. The orientation of the stream on the sky puts fairly strict limits on the plane of the orbit. The visible portion of the stream passes through Lynx, Cancer, and Hydra. We find no reasonable combination of parameters that would place the southern projection of the stream in either Monoceros or Canis Major. Nor does the visible extent of the stream complex fit either the prograde or retrograde models of the Monoceros stream computed by \\citet{penarrubia05}. Thus, even while the turn-off colors appear to be similar, we conclude on orbital grounds that this stream complex is unlikely to be related to either the Monoceros stream or the Canis Major overdensity." }, "0609/astro-ph0609163_arXiv.txt": { "abstract": " ", "introduction": "A large body of observational data provides evidence for the existence of dark matter, which manifests itself only through its gravitational influence on stellar systems of various scales (from individual galaxies to clusters of galaxies and superclusters), determining to a great extent their dynamics and structure. The controversy surrounding the following two main questions related to the dark matter is still continuing: What is its nature and to what extent does it exceed in mass the luminous matter? The first question is far from being resolved. As regards the second question, on the scales of individual galaxies, it is formulated as follows: Beginning from which regions (inner or outer) does the dark halo mass prevail over the luminous mass? There are several observational constraints on the mass and extent of the dark halos in galaxies; these constraints do not depend on what the nature of the dark matter is. In general, they give an estimate of the lower limit for the dark halo mass within a fixed radius. For spiral galaxies, this estimate primarily follows from the analysis of the contributions from various components of the system to its rotation curve for the so-called maximum disk model (a classical example of the separation of the contributions from the disk and the dark halo to the rotation curve of a galaxy, NGC 3198; van Albada et al. 1985). In this model, the dark halo mass within the optical radius of a galaxy generally does not exceed the stellar disk mass. When specific systems are investigated, a joint interpretation of the data on the rotation curves and radial velocity dispersion profiles, along with considerations regarding the marginal stability of stellar disks (see, e.g., Bottema and Gerritsen 1997; Khoperskov et al. 2001; Zasov et al. 2001; Khoperskov and Tyurina 2003)\\footnote{In these papers, the results of numerical $N$-body simulations with a variable disk mass were used to explain the observed stellar velocity dispersion for a number of spiral galaxies.}, often raises significantly the lower limit for the dark halo mass within a fixed radius (occasionally, interpretation of the data on the stellar velocity dispersion based on numerical simulations increases the dark halo mass by almost an order of magnitude compared to its value given by the maximum disk model\\footnote{The galaxy NGC 891 may be cited as an example (Khoperskov et al. 2001).}). The upper limit for the dark halo mass is much more difficult to constrain. Where the region within the optical radius of a galaxy is involved, stability-related considerations are invoked. For example, the presence of a regular spiral pattern related to the propagation of density waves in galaxies requires a gravitationally ``active'' stellar disk, which reduces the contribution from the spherical component to the total gravitational field of the spiral galaxy and to the total mass (Athanassoula et al. 1987). Another important test for the presence of a massive halo is the length of the tidal tails in interacting spiral galaxies. In interacting systems, we quite often observe extended tidal features stretching to distances as large as 40--100 kpc. Based on numerical simulations of interacting disk galaxies with dark halos that correspond to models obtained in cosmological calculations (Navarro et al. 1999), Dubinski et al. (1999) showed that the calculated parameters could be reconciled with the parameters of observed systems only for two types of models: models with an extended moderate-mass halo and a rotation curve that is determined almost entirely by the stellar disk within the optical radius and models with a compact low-mass halo that makes the main contribution to the rotation curve in the inner disk. Reconciling the rotation curves of specific interacting systems (e.g., NGC 4038/39 and NGC 7252) and the sizes of their tidal features with the model parameters leads one to conclude that the stellar disks in such systems dominate in mass within the optical radius, which gives a strong upper limit on the dark mass. A similar conclusion was also drawn from the simulations of the famous Mice galaxy, NGC 4676 (Sotnikova and Reshetnikov 1998). Comparison of the various estimates for the dark mass in galaxies shows that the upper and lower limits do not overlap (see, e.g., McGaugh and Block 1998). This gives reason to revise a number of tests, in particular, the test related to the thickness of the stellar disks in spiral galaxies suggested by Zasov et al. (1991, 2002). These authors argue that the thickness of a stable stellar disk at a given halo mass is limited below. If the disk thickness is smaller than a certain value, then bending instability will lead to increasing the vertical velocity dispersion and to a thickening of the system. The growth of bending perturbations is stabilized by a massive dark halo: the more massive the halo is, the thinner the stable disk can be. As we showed previously (Sotnikova and Rodionov 2005), not only a massive dark halo has a stabilizing effect on the growth of bending perturbations, leaving the stellar disk fairly thin. A low-mass compact bulge also produces such an effect. In this case, the lower limits on the dark mass can be mild and consistent with the upper limits. In this paper, we analyze the relationship between the stellar disk thickness and the dark halo mass that is derived both from theoretical considerations and from numerical simulations of the dynamical evolution of thin disks in the presence of spherical components with different density profiles and different masses. The derived relationship predicts moderate dark halo masses even for the thinnest galaxies. ", "conclusions": "" }, "0609/astro-ph0609480_arXiv.txt": { "abstract": "{A study of the structural and scaling properties of the temperature distribution of the hot, X-ray emitting intra-cluster medium of galaxy clusters, and its dependence on dynamical state, can give insights into the physical processes governing the formation and evolution of structure.} {Accurate temperature measurements are a pre-requisite for a precise knowledge of the thermodynamic properties of the intra-cluster medium.} {We analyse the X-ray temperature profiles from XMM-Newton observations of 15 nearby ($z<0.2$) clusters, drawn from a statistically representative sample. The clusters cover a temperature range from 2.5 keV to 8.5 keV, and present a variety of X-ray morphologies. We derive accurate projected temperature profiles to $\\sim 0.5\\,\\rv$, and compare structural properties (outer slope, presence of cooling core) with a quantitative measure of the X-ray morphology as expressed by power ratios. We also compare the results to recent cosmological numerical simulations.} {Once the temperature profiles are scaled by an average cluster temperature (excluding the central region) and the estimated virial radius, the profiles generally decline in the region $0.1\\,\\rv \\lesssim R \\lesssim 0.5\\,\\rv$. The central regions show the largest scatter, attributable mostly to the presence of cool core clusters. There is good agreement with numerical simulations outside the core regions. We find no obvious correlations between power ratio and outer profile slope. There may however be a weak trend with the existence of a cool core, in the sense that clusters with a central temperature decrement appear to be slightly more regular. } {The present results lend further evidence to indicate that clusters are a regular population, at least outside the core region. } ", "introduction": "The temperature and density are the key measurable characteristics of the hot, X-ray emitting intracluster medium (ICM). The determination of important derived properties such as entropy, pressure, and, under the assumption of hydrostatic equilibrium, the total mass, is dependent on accurate estimation of these quantities. Because of limited photon statistics\\footnote{Also the need for an azimuthally symmetric approximation for purposes of deprojection.} it is usual to measure the density and temperature in terms of radial profiles. However, while the density of the ICM is relatively easy to measure from the surface brightness profile of a given cluster, the temperature determination requires sufficient photon statistics to build, and fit, a spectrum. Thus ICM temperature profiles are typically determined with considerably less spatial resolution than density profiles. The measurement of radial temperature profiles is further complicated by the density squared ($n_e^2$) dependence of the X-ray emission. The steep drop of the X-ray surface brightness with distance from the centre, combined with the background from cosmic, solar and instrumental sources, makes accurate measurement of the temperature distribution at large distances from the centre a technically challenging task. The earliest temperature profiles were measured with {\\it Einstein}, {\\it EXOSAT}, {\\it Spacelab-2\\/} and {\\it GINGA\\/} only for the nearest, brightest clusters (e.g., \\citealt*{fab1,fab2,hughes,eyles,koy}). The low, stable background of {\\it ROSAT} made possible spatially resolved spectroscopy of poor clusters \\citep*[e.g.][]{david}; however, limited spectral resolution and bandwidth made such measurements difficult for hotter clusters \\citep*[e.g.][]{hbn,bh94,hb95}. {\\it ASCA\\/} and {\\it BeppoSAX\\/} had sufficient high-energy sensitivity to accurately measure the temperatures of hot clusters. However both of these satellites suffered from significant PSF blurring, which, in the case of {\\it ASCA}, was exacerbated by a significant energy dependence. As a result, at the end of the {\\it ASCA/BeppoSAX\\/} era, the exact shape of cluster temperature profiles was still under vigorous debate \\citep*{mark98,irwin99,whi00,ib00,fin01,dm02}. {\\it Chandra\\/} and \\xmm do not suffer from major PSF problems. The on-axis {\\it Chandra\\/} PSF is negligible, while the \\xmm PSF becomes an issue only for clusters with very centrally peaked core emission; in addition, neither is energy-dependent. Recent observations of moderately large samples consisting primarily of nearby cooling core clusters with \\xmm \\citep{piff} and {\\it Chandra\\/} \\citep{vikh05} have largely validated the original {\\it ASCA\\/} results of Markevitch et al., which suggested that temperature profiles declined from the centre to the outer regions. However, other {\\it Chandra} and \\xmm observations have found flatter profiles \\citep{allen01,kaa,app}. As of the time of writing, no systematic attempt has been made, with either \\xmm or {\\it Chandra}, to look at the temperature profiles of a representative sample of nearby clusters\\footnote{Some work has been done on medium-distant clusters, see \\citep{zhang,kotov}}. Although other projects on representative samples are in progress \\citep[e.g.,][]{reiprich}, they are not expected to be able to map the temperature distribution out to large radius. In this paper we deal with observations of 15 clusters from a statistically representative sample observed with \\xmm. We describe in detail the data reduction and background subtraction, and compare our results with previous work and with those from cosmological hydrodynamical simulations. We also make a preliminary investigation of correlations with quantitative morphological measures. We present only projected temperature profiles in this paper -- such profiles are direct observables and do not depend on complicated PSF and deprojection algorithms. We will deal with correction of the profiles in forthcoming papers which make use of observations of the full sample. All results are given assuming a $\\Lambda$CDM cosmology with $\\Omega_m=0.3$ and $\\Omega_\\Lambda=0.7$ and $H_0 = 70$~km~s$^{-1}$~Mpc$^{-1}$. Unless otherwise stated, errors are given at the $68$ per cent confidence level. ", "conclusions": "We have used \\xmm observations of 15 clusters drawn from a statistically representative, luminosity-selected sample to investigate the behaviour of the temperature profiles. The clusters range from morphologically relaxed looking objects with strong central surface brightness peaks (e.g., RXC\\,J1044\\,-0704), to diffuse structures with significant amounts of surrounding substructure (e.g., RXC\\,J1516\\,+0056), and constitute a representative subsample. We find that, once scaled appropriately, the temperature profiles are similar in the radial range from $0.1\\,\\rv$ to $0.5\\,\\rv$, declining steadily from the central regions to the outer boundary of the measurements with a relative dispersion of $\\sim 10$ per cent out to $0.5\\,\\rv$. The region interior to $0.1\\,\\rv$ is the region of greatest scatter in the scaled profiles: the relative scatter of $\\sim 25$ per cent is likely a lower limit. A preliminary comparison with numerical simulations shows relatively good agreement outside $\\sim 0.1\\,\\rv$, with all of the measured temperature profiles falling within the scatter of the simulated profiles. Calculating power ratios for the sample, we investigate whether there are correlations between the power ratio measured in an aperture corresponding to $R_{500}$ and the shape of the temperature profile. We characterise the temperature profile shape in two ways: with the ratio $T (0.5\\,\\rv)/T (0.2\\,\\rv)$, a measure of the outer slope, and with the ratio $T_c/\\langle T \\rangle$, a measure of the central temperature drop. There is no obvious correlation of outer slope with power ratio; neither is there a correlation of central temperature dip with $P_2/P_0$ or $P_3/P_0$. There is evidence for a weak correlation of the central temperature dip with $P_4/P_0$. The analysis thus suggests that the outer slope of the temperature profile is not particularly sensitive to the morpho-dynamical state, although there may be some correlation with the existence of a central temperature drop. Further investigation with power ratios evaluated in other apertures, for the entire sample, should be undertaken before definitive conclusions can be drawm. The overall conclusion from this work on a statistically representative sample indicates that clusters are a relatively regular population, at least outside the cool core regions, with the caveat that comparisons between samples or with simulations is limited by the available temperature profile resolution. The observed similarity in density \\citep{neuarn,croston} and temperature profiles \\citep[][this work]{mark98,dm02,piff,vikh05} indicates both similarity in the underlying gravitational mass distribution (such as has already been seen in the X-ray mass profiles of morphologically relaxed clusters, e.g., \\citealt*{point}), and similarity in the entropy of the ICM (such as has been seen by e.g., \\citealt*{psf,pap}). In this case a single integrated temperature, excluding the core region, should be a good proxy for the total mass. The observed regularity thus has important implications for the use of clusters as cosmological probes. In future papers, we will reinvestigate the trends with the full sample, make maps of quantities such as temperature, entropy and pressure, and estimate the mass, baryon fraction and entropy in the clusters. A more extensive comparison with numerical simulations will also be undertaken." }, "0609/astro-ph0609355_arXiv.txt": { "abstract": "{The BH mass (and the related Eddington ratio, $l$ = $L_{bol}/L_{edd}$) in broad line AGN is usually evaluated by combining estimates (often indirect) of the BLR radius and of the $FWHM$ of the broad lines, under the assumption that the BLR clouds are in Keplerian motion around the BH. Such an evaluation depends on the geometry of the BLR. There are two major options for the BLR configuration: spherically symmetric or ``flattened''. In the latter case the inclination to the line of sight becomes a relevant parameter. This paper is devoted to evaluate the bias on the estimate of the Eddington ratio when a spherical geometry is assumed (more generally when inclination effects are ignored), while the actual configuration is ``flattened'', as some evidence suggests. This is done as a function of luminosity and redshift, on the basis of recent results which show the existence of a correlation between the fraction of obscured AGN and these two parameters up to at least z=2.5 (date at larger redshifts being insufficient.) The assumed BLR velocity field is akin to the ``generalized thick disk'' proposed by Collin et al. (2006). Assuming an isotropic orientation in the sky, the mean value of the bias is calculated as a function of luminosity and redshift. It is demonstrated that, on average, the Eddington ratio obtained assuming a spherical geometry is underestimated for high luminosities, and overestimated for low luminosities. This bias converges for all luminosities at $z$ about 2.7, while nothing can be said on this bias at larger redshifts due to the lack of data. The effects of the bias, averaged over the luminosity function of broad line AGN, have been calculated. The results imply that the bias associated with the a-sphericity of the BLR make even worse the discrepancy between the observations and the predictions of evolutionary models. \\keywords {Galaxies: active - galaxies: nuclei - quasars: general} } ", "introduction": "The accretion rate and the black hole mass are the two fundamental parameters in our understanding of the Active Galactic Nuclei (AGN) phenomenon. Measurements of these two quantities are, unfortunately, not devoid of significant uncertainties. The accretion rate $\\dot m$ is derived from the bolometric luminosity, $L_{bol}$, under assumptions on the efficiency $\\eta$ for the conversion of gravitational energy, $L_{bol}$ = $\\eta$$\\dot m$$c^2$. For non-rotating black holes (BH), $\\eta$ = 0.057 is generally adopted, assuming that effective (for the observer of the electromagnetic radiation) conversion takes place down to the marginally stable circular orbit at three times the Schwarzschild radius, 3$R_s$; for rotating BH, $\\eta$ can reach the maximum value of 0.42. $L_{bol}$ is generally obtained from the luminosity observed in a given band, multiplied by a factor based on the Spectral Energy Distribution (SED) attributed to the specific class the AGN belongs to. This procedure is regarded to be rather safe, but in fact there are still uncertainties on the luminosity and/or redshift dependence of the bolometric correction. For the BH mass, two ``direct'' methods have been followed. The first, applicable only to AGN, is the reverberation mapping (RM) method. This method is based on the principle (Blandford \\& McKee 1982) that the delay in the response of the lines from the Broad Line Region (BLR) to variations of the continuum is a measure of the size of this region, $R_{BLR}$. Assuming that the line widths are due to motions governed by the BH, the combination of $R_{BLR}$ and a velocity derived from the line profiles yields a ``Keplerian'' estimate of the BH mass. The other method is based on the fairly strict correlation between the mass of the BH and properties of the stellar bulge of the host galaxy. This method, having proved very reliable for a rather large sample of galaxies (Gebhardt et al. 2000a, Ferrarese \\& Merritt 2000, Tremaine et al. 2002), represents a benchmark for the previous method when it can be applied to AGN (Gebhardt et al. 2000b, Ferrarese et al. 2001, Onken et al. 2004). The agreement found, although far from perfect (e. g. Collin et al. 2006), has encouraged the extension to many more AGN (especially the distant and more luminous ones), for which both the RM and the bulge methods can hardly be applied, of a ``secondary'' method. The latter is based on the estimate of $R_{BLR}$ through an empirical correlation between this quantity and the luminosity (see Sect. 2) which has emerged from the RM measurements. The BH mass can be univocally converted into the Eddington luminosity, $L_{edd}$. Thus a quantity $l$ can be defined: $l = L_{bol}/L_{edd}$, which, altough it tells us nothing precise about the accretion rate, is of high interest because $L_{edd}$ is a very significant physical limit. This quantity is often referred to as the Eddington ratio. Several papers have been devoted to explore the behaviour of $l$, in particular as a function of $L$ and the cosmological epoch, namely the redshift $z$. Among the uncertainties and the selection effects which may plague the results, the present paper is devoted to point out and evaluate a particular bias, linked to the possibility that the spatial distribution of the BLR clouds is far from spherical, a situation supported by various lines of evidence. The evaluation is based on a recent result on the fraction of AGN which are photoelectrically absorbed in the X-rays (column density $N_H$ $>$ $10^{22}$ H atoms/cm$^2$, and Compton thin), which can be summarized as follows. Calling $\\xi$ the ratio of the absorbed ones to the total, it turns out that $\\xi$ is a function of $L_X$ (hence of $L_{bol}$) (Ueda et al. 2003, La Franca et al. 2005) as well as of $z$ (La Franca et al. 2005). Qualitatively speaking, in the local Universe this fraction decreases with increasing luminosity; as the redshift grows, the anticorrelation remains but it becomes progressively shallower. If this behaviour is associated with a luminosity and redshift dependence of the opening angle of the absorbing matter, within which the BLR can be observed, it should introduce a bias on the estimate of the BH mass of broad line AGN (AGN 1 for short), when this is performed using the RM method and its ``secondary'' extrapolation. To this effect it is important to stress (see Fiore et al. 2003, Perola et al. 2004) that the value $N_H$ = $10^{22}$ H atoms/cm$^2$ works as a good (the exceptions are a minority) discriminant between AGN which are optically classified as type 1 and type 2. The plan of the paper is as follows. In Sect. 2 the results on $l$ from the literature are summarized. In Sect. 3 two lines of evidence, again from the literature, which favour a non-spherical distribution of the BLR clouds are briefly described. In Sect. 4 the bias associated with the a-sphericity of the BLR, and its dependence on $L_{bol}$ and $z$, as it can be predicted on the basis of the abovementioned finding, is quantified. A discussion follows in Sect. 6. ", "conclusions": "The curves in Fig. \\ref {qmedioZLbol} show that the mean value of the bias, $<$q$>$, is typically greater than 1 and can be as large as about 2. Furthermore, this bias increases with the luminosity, converging for all luminosities towards a value of order unity at $z$ larger than 2. As noted previously, the estimate of the opening angle $i_0$, hence of $<$q$>$, ceases to be reliable in this regime of the cosmic time, because the lack of data prevented La Franca et al. (2005) from evaluating the actual evolution of $\\xi$ further back in time. The ``saturation'' effect adopted in $\\xi$ corresponds to an angle $\\simeq 40^{\\circ}$, such that, according to eq. (\\ref {qmedio}), $<$q$>$ is $\\simeq 1$ for $a$=0.3, or $\\simeq 0.7$ for $a$=0.1. Since the weight of the solid angle within this value of $i_0$ is relatively modest, the mean value of q will result significantly lower than unity only for $i_0$ much lower than $i_{\\ast}$. If, for instance, $\\xi$ were to reach 90\\% (that is $i_0$ = 26$^{\\circ}$) at values of $z$ much greater than 2, than $<$q$>$ would be 0.32 ($a$=0.1) or 0.56 ($a$=0.3). We also calculated the rms of the $q$ distribution and found values of 0.9, 0.6 and 0.2 for $a$=0.1 and $i_0$=90$^{\\circ}$, 60$^{\\circ}$, 30$^{\\circ}$, respectively (the corresponding values of $<$q$>$ are 2.0, 1.3 and 0.4). Similar values are found for $a$=0.3. The expected scatter in masses is thus reassuringly smaller than that found in the Black Hole mass (estimated from the reverberation mapping) vs. bulge dispersion velocity relationship (Onken et al. 2004). Turning to the parameter $l$, this quantity is {\\it underestimated} when $<$q$>$ $>$ 1, if $M_{BH}$ is ``measured'' assuming a spherical distribution of the BLR, the other way round if $<$q$>$ were less than unity. A possible way to illustrate the effects of this bias consists in calculating its value averaged over the entire Luminosity Function. La Franca et al. (2005) give both the Luminosity Function (LF) and the $\\xi$ which best fit the data (fit \\#4 in their Table 2), after taking into account the selection effects (in the X-ray and optical bands) for the samples used. The LF behaviour with cosmic time follows a Luminosity Dependent Density Evolution. One can therefore combine the LF and $\\xi$ to obtain, for a given $z$, the average bias, $<$q$>$$_{L}$, over the entire luminosity range: \\begin{equation} \\label{qmedioL} _{L}=\\frac{\\int_{\\log\\!L_{X1}}^{\\log\\!L_{X2}}\\frac{\\mathrm{d}\\Phi_{1}(L_{x},z)}{\\mathrm{d}\\!\\log\\!L_{X}}\\, \\mathrm{d}\\!\\log\\!L_{X}\\int_{0}^{i_{0}(L_{X},z)}q\\, \\sin i\\, \\mathrm{d}i} {\\int_{\\log\\!L_{X1}}^{\\log\\!L_{X2}}\\frac{\\mathrm{d}\\Phi_{1}(L_{x},z)}{\\mathrm{d}\\!\\log\\!L_{X}}\\, \\mathrm{d}\\!\\log\\!L_{X}\\int_{0}^{i_{0}(L_{X},z)}\\sin i\\, \\mathrm{d}i}, \\end{equation} where $\\Phi_{1}$ is the luminosity function of unabsorbed AGN (which depends on $\\xi$) and the parameter $a$ is included in the function $q$ (eq. (\\ref {qi})). This result should, by construction, be free from selection effects. For a choice of $\\log\\!L_{X1}$ = 42 and $\\log\\!L_{X2}$ = 48, $<$q$>$$_{L}$ is illustrated in the left panel of Fig. \\ref{qmedioL} (the corresponding values of $\\log\\!L_{bol}$, obtained applying the luminosity dependent bolometric correction of Marconi et al. (2004), are: $\\log\\!L_{bol1}$ = 43.030 and $\\log\\!L_{bol2}$ = 50.937). \\begin{figure*}[ht] \\hbox{ \\includegraphics[width=7 cm]{Lamastra06fig4a.ps} \\hspace{0.02cm} \\includegraphics[width=7 cm]{Lamastra06fig4b.ps} } \\caption{Left: mean value of q, $<$q$>$$_{L}$, between 42$<\\log\\!L_{x}<$48 as a function of z, for a=0.1 (solid line) and a=0.3 (dotted line). Right: mean value of q, $<$q$>$$_{L}$, between 41$<\\log\\!L_{x}<$48 (dashed line) and 42$<\\log\\!L_{x}<$48 (solid line) as a function of z, for a=0.1.} \\label{qmedioL} \\end{figure*} To illustrate the effect arising from the inclusion of the numerous AGN in the decade around $\\log\\!L_{X}$ = 41 ($\\log\\!L_{bol}$ = 41.893), in the right panel of Fig. \\ref {qmedioL} (and only for the case $a$=0.1) $<$q$>$$_{L}$ is compared with the result shown in the left panel of the same figure. The effect is that, the higher is the luminosity, the steeper is the dependence of the mean bias $<$q$>$$_{L}$ on $z$. A direct application of our results to the samples used by the various authors, quoted in Sect. 2, is not straightforward, because this would require proper evaluation of their specific observational selection effects, a hard task which goes beyond the aims of this paper. However, some general remarks can be made. The results shown in Fig. \\ref {qmedioL} imply that, if $l$ were intrinsically constant with $z$, and if its ``mean'' value were calculated with the spherical approximation (or, more generally, with a constant value of $\\kappa$ in eq. (\\ref {VblrVfwhm})) then one should observe an increase of $l$ with $z$. Since the observational results indicate instead a constant value, the bias discussed in this paper implies that the actual value of $l$ \\textit{decreases} with $z$. In semianalytical models which link the evolution of the galaxies in the hierarchical clustering scenario with the quasar evolution (e.g. Menci et al. 2003, 2004), the black hole accretion is triggered by galaxies encounters. In this scenario, at high $z$ the protogalaxies grow rapidly by hierarchical merging, meanwhile much cold gas is imported and also destabilized, so that the black holes are fueled at their Eddington rates. At lower $z$ the accretion rate of cold gas onto the central black hole diminish due to the combined effects of the decrease of the galaxies merging and encounter rates and the decrease of the amount of galactic cold gas, which was already converted into stars or accreted onto the black hole. This model predicts an average Eddington ratio dropping from $l$ $\\simeq$ 1 at $z$ $\\simeq$ 2.5 to $l$ $\\simeq$0.01 at $z$ $\\simeq$0 (Menci et al. 2003). Our results imply that the bias associated with the a-sphericity of the BLR make even worse the discrepancy between the observations and the predictions of the models." }, "0609/astro-ph0609433_arXiv.txt": { "abstract": "We present spectra of 59 nearby stars candidates, M dwarfs and white dwarfs, previously identified using high proper motion catalogues and the DENIS database. We review the existing spectral classification schemes and spectroscopic parallax calibrations in the near-infrared $J$-band and derive spectral types and distances of the nearby candidates. 42 stars have spectroscopic distances smaller than 25~pc, three of them being white dwarfs. Two targets lie within 10~pc, one M8 star at 10.0~pc (APMPM J0103-3738), and one M4 star at 8.3~pc (LP 225-57). One star, LHS 73, is found to be among the few subdwarfs lying within 20 pc. Furthermore, together with LHS 72, it probably belongs to the closest pair of subdwarfs we know. ", "introduction": "Recent discoveries of cool objects, such as M stars or brown dwarfs, closer than 5 parsecs show that even the immediate solar neighbourhood sample is still incomplete \\citep{delfosse2001,scholz2003,teegarden2003,hambly2004}. \\citet{henry1997} estimated that about 130 systems over 359 (36\\%) are missing within 10 pc. The missing fraction is even larger within 25 pc (63\\%) with a deficit of about 3500 systems over the 5500 expected ones \\citep{henry2002}. Statistical comparisons from the local sample in the northern hemisphere led to a less pessimistic result, the current 10 pc sample being $\\sim$75\\% complete \\citep{reid2003a}. New surveys in the near infrared such as DENIS \\citep{epchtein1997} and 2MASS \\citep{cutri2003} provide unprecedented data for a systematic search for low luminosity cool dwarfs. The use of these data together with high proper motion catalogues is a powerful tool for discovering our neighbours. Hundreds of stars closer than 25 parsecs have been discovered this way \\citep[e.g.][]{phanbao2001,phanbao2003,reid2002a,reid2002b,reyle2002,reid2003b,reid2004,hambly2004,reyle2004,lodieu2005,scholz2005}. As spectroscopy provides much more information than photometry alone, spectroscopic observations were also carried out to identify and classify nearby stars. Recent studies dealing with sample sizes larger than 10 objects are cited in Table~1. All together, they revealed 490 stars in the 25 pc sample and 25 within 10 pc. However, we note that the samples investigated by the different authors do have many stars in common so that the total number of newly discovered neighbours is much smaller. \\begin{table} \\label{tab1} \\caption{Nearby stars identified by spectroscopy } \\begin{tabular}{lccc} \\hline Reference \t&Sample &\\multicolumn{2}{c}{Number of stars within} \\\\ & size &25 pc\t&10 pc \\\\ \\hline \\citet{cruz2002}\t \t&70\t\t\t&28\t\t& \\\\ \\citet{henry2002} \t&34\t\t\t&6\t\t&2 \\\\ \\citet{reid2003b} \t&357\t\t&127\t&9 \\\\ \\citet{lodieu2005}\t&71\t\t\t&25\t\t&3 \\\\ \\citet{crifo2005}\t\t&39\t\t\t&31\t\t&1 \\\\ \\citet{scholz2005}\t&322\t\t&226\t&8 \\\\ \\citet{phanbao2006}\t&45\t\t\t&5\t\t& \\\\ This study\t\t\t&59\t\t\t&42\t\t&2 \\\\ \\hline\\\\ \\end{tabular} \\end{table} In previous papers \\citep{reyle2002,reyle2004}, we have determined the photometric distances of high proper motion stars that we cross-identified with the DENIS survey. We have reported the discovery of 115 nearby candidates, probably lying within 25 pc, the limit of the Catalogue of Nearby Stars \\citep[CNS3]{gliese1991}. We selected the closest candidates for a spectroscopic follow-up performed at La Silla Observatory (Chile). Our sample of nearby candidates is described in \\S~\\ref{sample}. \\S~\\ref{spectro} describes the spectroscopic observations. In \\S~\\ref{classification} we review the spectroscopic classification schemes and compute the spectral type of the stars. Computation of distances is detailed in \\S~\\ref{distances}. In \\S~\\ref{halpha} we discuss the chromospheric activity of the stars in our sample and the conclusions are given in \\S~\\ref{conclusion}. ", "conclusions": "\\label{conclusion} We obtained spectral types and spectroscopic distances for 59 high proper motion stars. These stars are nearby candidates, as seen from their photometric distances. 42 stars have indeed spectroscopic distances below the 25~pc limit of the Catalogue of Nearby Stars. These new neighbours are few in regards to the large number of missing stellar systems within 25 pc ($\\sim$ 2000) according to \\citet{henry2002}. However, this study, joined to similar ones, allows to increase, step by step, the completeness of the solar neighbourhood sample and to cross-check the new nearby candidates in different samples (see e.g. Table 1). Nevertheless, it is clear that further efforts are needed in order to reach a much higher completeness level. The neighbours recovered from our study are mainly M dwarfs. Three are white dwarfs, cooler than 6200~K. One star, APMPM J0541-5349, appeared to be a binary consisting of two M2 stars with an angular separation of 3.33\\arcsec, at a distance of 53.4 $\\pm$ 4.1 pc from the Sun. We computed the radial velocity of the subdwarf candidate LHS 73 and found a space motion compatible with that of an object belonging to an old population. With our determined distance of 18.7 pc, LHS 73 is among the few subdwarfs within 20 pc. Furthermore, LHS 73 has a companion, LHS 72. LHS 72/LHS 73 is probably the closest known pair of subdwarfs. Two stars lie within 10~pc: APMPM J0103-3738 is a M8 star at distance d = 10.0 $\\pm$ 0.8 pc, and LP 225-57 a M4 star at d = 8.3 $\\pm$ 2.8 pc. The new sources that we identified are worthy of more detailed observations, including high-resolution imaging to search for low-mass companions, and trigonometric parallax measurements. We plan more follow-up for the nearest objects and hope to place the most promising targets on new or existing parallax programs such as that of the Cerro Tololo Inter-American Observatory Parallax Investigation \\citep[CTIOPI]{jao2005}." }, "0609/astro-ph0609119_arXiv.txt": { "abstract": "We present high sensitivity sub-arcsecond resolution images of the Herbig Ae star AB Aurigae at 11.6 and 18.5 $\\mu$m taken with Michelle on Gemini North. Bright extended dust emission close to the star is resolved at both wavelengths, with quadratically subtracted FWHM of 17$\\pm$4 AU at 11.6 $\\mu$m and 22$\\pm$5 AU at 18.5 $\\mu$m. Additional, fainter emission is detected out to a radius of 280 AU at 11.6 $\\mu$m and 350 AU at 18.5 $\\mu$m down to the sensitivity limit of the observations. The latter value is identical to the measured size of the millimeter-continuum disk, but much smaller than the CO disk. Assuming moderately absorbing material, we find that larger particles ($\\sim 1$ $\\mu$m) dominate the mid-IR emission in the inner ($<$ 100 AU) regions of the disk, and smaller particles ($<$ 0.3 $\\mu$m) dominate in the outer regions of the disk. A model of a nearly face-on passive flared disk with an inner rim accounts well for our observations. ", "introduction": "Located only 144 pc away \\citep{anc98}, the 2 Myr old A0 star AB Aurigae is the brightest (V=7.06) of the original sample of Herbig stars \\citep{her60}, which are intermediate mass (2 to 8 M$\\sun$) pre-main sequence stars. Consequently, it is not only the best-studied Herbig object, but it has become an important touchstone for our understanding of the class. The spectral energy distribution (SED) of this source shows emission in excess of the photosphere throughout the infrared region indicative of circumstellar (CS) dust. Different models, among which are a highly inclined passive flared disk with an inner rim \\citep{dul01}; a flat thick disk surrounded by a halo \\citep{vin03}; and a halo alone \\citep{eli04}, have been used to explain the spatial distribution of this dust. Despite their differences, all of these models reproduce reasonably well the observed SED, which indicates the need for high-resolution imaging to provide additional critical constraints. Spatial observations at various wavelengths imply that the CS dust in the AB Aur system lies in a disk and some type of more extended structure. An inhomogeneous envelope extending to 1300 AU is apparent in optical scattered light \\citep{gra99}, while closer to the star, in optical and near-IR scattered light, one sees what appears to be a disk with quasi-spiral structure, a radius of 580 AU, and an inclination of 30$^\\circ$ (face-on = 0$^\\circ$), assuming flat geometry \\citep{gra99, fuk04}. CO observations reveal a complex disk with an inner hole of about 70 AU extending out to 1000 AU and possibly having non-Keplerian motions, while 1.4 mm continuum observations indicate a disk with an inner radius of 110 AU and an outer radius of 350 AU \\citep [Pi\\'{e}tu et al. 2005; see also][]{man97, cor05}. Near-IR interferometric studies resolve the inner 0.7 AU region of the disk \\citep{mil01}. However, previous mid-IR studies present somewhat contradictory results. Marsh et al. (1995) report extended structure at 17.9 $\\mu$m with a semi-major axis of 80$\\pm$20 AU and an inclination of 75$^\\circ$. Chen \\& Jura (2003), using Keck, do not confirm that detection of extended structure, and at 20.5 $\\mu$m using deconvolved images with a resolution of 0.$\\arcsec$6, Pantin et al. (2005) report an elliptical ring-like structure at an average distance of 280 AU from the star. In addition, Liu et al. (2004) resolve the inner disk interferometrically at 10.3 $\\mu$m determining a size of 27$\\pm$3 AU and an inclination of 45$^\\circ$. There is also recent evidence that AB Aur could be the brighter component of a binary system, with a companion separation most likely between 1 and 3 arcseconds \\citep{bai06}. In this paper we present deep mid-infrared images of AB Aur obtained at Gemini North. We have resolved the emission close to the star at 11.6 and 18.5 $\\mu$m, and we find an additional extended component that appears to be roughly circularly symmetric. We show how these observations of the thermal emission from dust in the AB Aur system help establish a more coherent picture of the dust geometry consistent with most observations at other wavelengths. ", "conclusions": "Our mid-IR images reveal two different emission components in AB Aur. The central stronger emission is resolved, with quadratically deconvolved FWHM sizes of 17$\\pm$4 AU and 22$\\pm$5 AU at 11.6 $\\mu$m and 18.5 $\\mu$m, respectively. We also detect fainter extended emission out to a radius of 280 AU at 11.6 $\\mu$m and 350 AU at 18.5 $\\mu$m. Emission is slightly elongated at 12 $\\mu$m, indicating a disk inclination angle in the range 29$^\\circ \\pm 11^\\circ$ and a PA of 80$^\\circ \\pm 11^\\circ$. The morphology at 18 $\\mu$m is consistent with an inclination angle of 12$^\\circ \\pm 12^\\circ$. However, within the uncertainties, inclination angles at 12 and 18 $\\mu$m are the same. Assuming moderately absorbing material, we derive average radii of the mid-IR emitting dust in the system and find that larger particles ($a \\sim 1$ $\\mu$m) dominate the mid-IR emission in the inner ($<$ 100 AU) regions of the disk, and smaller particles ($a <$ 0.3 $\\mu$m) dominate in the outer regions of the disk. Our results are reasonably well accounted for by a model of a passive flared disk with an inner rim. The presence of a more spherical component fails to account for the particle size segregation derived from our observations." }, "0609/astro-ph0609605_arXiv.txt": { "abstract": "{Observations of X-Ray sources harbouring a black hole and an accretion disc show the presence of at least two spectral components. One component is black-body radiation from an optically thick standard accretion disc. The other is produced in a optically thin corona and usually shows a powerlaw behaviour. Electron-proton (ep) bremsstrahlung is one of the contributing radiation mechanisms in the corona. Soft photons from the optically thick disc can Compton cool the electrons in the corona and therefore lead to a two-temperature plasma, where electrons and ions have different temperatures.} {We qualitatively discuss effects on ep-bremsstrahlung in the presence of such a two-temperature plasma.} { We use the classical dipole approximation allowing for non-relativistic electrons and protons and apply quantum corrections through high-precision Gaunt factors. } {In the two-temperature case ($T_\\textrm{e}< T_\\textrm{p}$) the protons cause a significant fraction of the ep-bremsstrahlung if their speed is high compared to the electrons. We give accurate values for ep-bremsstrahlung including quantum-mechanical corrections in the non-relativistic limit and give some approximations in the relativistic limit.} {The formulae presented in this paper can be used in models of black hole accretion discs where an optically thin corona can comprise a two-temperature plasma. This work could be extended to include the fully relativistic case if required.} ", "introduction": "Models of black hole accretion discs are widely used to explain spectral characteristics of X-Ray observations of such objects. These models usually consist of an optically thin and hot corona and a cool, optically thick standard accretion disc. The optically thin corona consists of a two-temperature plasma. Electrons cool by bremsstrahlung, mainly by electron-proton bremsstrahlung. The electron-proton system has got a dipole moment while the electron-electron and proton-proton system can only emit radiation through the (much weaker) quadrupole moment. \\citet{1975ApJ...199L.153E} presented a model for Cyg X-1 incorporating a two-temperature plasma in the inner parts of the accretion flow. This kind of model has been subsequently developed \\citep[e.g.][]{1976ApJ...204..187S,1995MNRAS.277...70Z} and applied to different objects \\citep{1991ApJ...380L..51H,1993ApJ...413..507H} and different geometries \\citep{1991ApJ...380...84W}. The calculation of the rate of bremsstrahlung coming from a hot gas has received much attention over a long period of time \\citep[e.g.][]{1962RvMP...34..507B,1970RvMP...42..237B,1980ApJ...238.1026G,1981ApJ...243..677G}. Today the bremsstrahlung rate is known with a very high accuracy. All these calculations however assume electron and proton temperature to be equal. In the following we limit ourselves to a purely classical and non-relativistic treatment of electron-proton bremsstrahlung. We do this in order to highlight the elementary physical processes involved. Since we apply Gaunt factors to account for quantum mechanical effects, our results are strictly accurate only in the non-relativistic case. An expansion to the fully relativistic case is feasible but beyond the scope of this paper. In this contribution we first highlight the physical mechanism in Sect.~\\ref{sect:mechanism} and then present a recalculation of the electron-proton Bremsstrahlung rate for a two-temperature plasma in the non-relativistic limit in Sect.~\\ref{sect:calc}. While we limit ourselves to a electron-proton plasma, the formulae can easily be modified to consider mixtures of ions with different mass and charge. We give a discussion of the results and our conclusions in Sect.~\\ref{sect:conclusions}. ", "conclusions": "\\label{sect:conclusions} We have recalculated electron-proton bremsstrahlung for a two-temperature plasma. Owing to the increasing importance of the proton speed relative to the electron speed for $T_\\textrm{e}< T_\\textrm{p}$, the protons can contribute significantly to the electron-proton bremsstrahlung. They dominate the electron-proton Bremsstrahlung losses for electron temperatures lower than $m_\\textrm{e}/m_\\textrm{p}\\approx 1/1836$ times the proton temperature. While our results are strictly valid only in the non-relativistic regime, we give a crude extrapolation for the relativistic case. The work presented here should only be considered as exploratory work to outline the influence of a two-temperature plasma on the electron-proton bremsstrahlung. If calculated for both the non-relativistic and relativistic regime, one needs to account for the relativistic kinematics and consider the \\citet{betheheitler} cross section, adjusted for the effect presented here. While it is certainly needed and desirable, such a treatment is beyond the scope of this paper. The apparent dominance of electron-electron bremsstrahlung for electron temperatures in excess of $10^9$ K, however, does not make it that necessary. The mechanism presented here, seen in the context of Coulomb collisions between electrons and protons, is already well known in plasma physics. The NRL Plasma Formulary \\citep{NRL_FORMULARY_06}, a standard reference in plasma physics for more than 25 years, gives a formula for the Coulomb collision rate which depends only on the electron temperature for $T_\\textrm{e}>m_\\textrm{e}/m_\\textrm{p} T_\\textrm{p}$. This is the classical \\citet{1962pfig.book.....S} result. For higher proton temperatures, the electron temperature dependence weakens and the collision rate only depends on the proton temperature owing to their larger speed compared to the electrons. This effect is seen as well in the work by \\citet{1983MNRAS.202..467S} who derives the Coulomb collision rate in a fully relativistic framework. Similar to our extrapolation for the electron-proton bremsstrahlung rate, they find a change in the exponent of the temperature dependence of $+1/2$, i.e. from $T^{-3/2}$ to $T^{-1}$. The importance of the effect presented in this paper needs to be examined by comparing the characteristic timescales for Coulomb collisions to equilibrate electron and proton temperature and the Bremsstrahlung timescale. While it may not be of that strong influence for the energetics of the two-temperature plasma, it could have an observable signature due to the higher high-energy cut-off for the Bremsstrahlung created by the kinetic energy of the protons, if it is not hidden behind some more important emission mechanisms at these energies." }, "0609/astro-ph0609575_arXiv.txt": { "abstract": " ", "introduction": " ", "conclusions": "We discussed the expected neutrino flux from $\\gamma$-transparent accelerators of cosmic rays, and emphasized the case of young SNR. We showed that for RX J1713.7-3946 (the best known SNR in TeV sky, thanks to H.E.S.S.) the expectations are stable: $\\sim 5$ events per km$^2$ per year in an ideal detector. The median neutrino energy is $3$ TeV. Since the detected $\\mu$ are softer than the $\\mu$ in the production point (that in turn is softer than the impinging $\\nu$) several events will fall in an energy region where the atmospheric background and the role of imperfect detection efficiency are important: see the lectures of Lipari and Lucarelli. Thus, we believe that it would be desirable to have a detailed discussion of the characteristics of a detector that aims to see a $\\nu$ signal from RX J1713.7-3946. \\vskip2mm Sometimes soon H.E.S.S.\\ should tell us more on another intense VHE $\\gamma$-ray source, RX J0852.0-4652 (Vela Jr). This SNR has $F_\\gamma=6.5$ million $\\gamma$-rays $E^{-2.1}/(\\rm TeV\\ km^2\\ yr)$ below $10$ TeV. If the exponential cutoff is at $E_{\\rm \\gamma cut}=50$ $(150)$ TeV (and again, if these $\\gamma$ rays are of hadronic origin) we get $N_\\mu + N_{\\bar\\mu}=10$ $(14)/(\\rm km^2 yr)$ in an ideal detector, with a significantly higher energy~\\cite{vis06}. \\subsection*" }, "0609/astro-ph0609096_arXiv.txt": { "abstract": "We investigate the dynamical interaction between a galactic disk and surrounding numerous dark subhalos as expected for a galaxy-sized halo in the cold dark matter (CDM) models. Our particular interest is to what extent accretion events of subhalos into a disk are allowed in light of the observed thinness of a disk. Several models of subhalos are considered in terms of their internal density distribution, mass function, and spatial and velocity distributions. Based on a series of N-body simulations, we find that the disk thickening quantified by the change of its scale height, $\\Delta z_d$, depends strongly on the individual mass of an interacting subhalo $M_\\mathrm{sub}$. This is described by the relation, $\\Delta z_d/R_d \\simeq 8 \\sum_{j=1}^N (M_{\\mathrm{sub},j}/M_d)^2$, where $R_d$ is a disk scale length, $M_d$ is a disk mass, and $N$ is the total number of accretion events of subhalos inside a disk region ($\\le 3R_d$). Using this relation, we find that an observed thin disk has not ever interacted with subhalos with the total mass of more than 15~\\% disk mass. Also, a less massive disk with smaller circular velocity $V_c$ is more affected by subhalos than a disk with larger $V_c$, in agreement with the observation. Further implications of our results for the origin of a thick disk component are also discussed. ", "introduction": "The cold dark matter (CDM) paradigm has become a standard framework for understanding the structure formation in the Universe. According to this theoretical paradigm, the growing process of self-gravitating structures is hierarchical in the sense that small dark matter halos virialize first, and aggregate successively into larger and larger objects. This clustering process of dark matter halos is successful for explaining a wide variety of observations including the large-scale distribution of galaxies. In this CDM scenario, N-body simulations are an important tool in order to investigate the non-linear growth of cosmological structures. Early N-body simulations based on the CDM models suffered from the so-called {\\it over-merging} problem, i.e., substructures are disrupted very quickly within dense environments (\\cite{summers1995}). However, recent high-resolution N-body simulations have revealed the presence of hundreds of dark matter substructures (subhalos) which survive in not only cluster scales but also galactic scales (\\cite{moore1999}; \\cite{klypin1999}). This large number of subhalos in a galaxy-sized halo is in contrast to only about a dozen satellite galaxies in the Galaxy, which confronts so-called ``the Missing Satellite Problem''. Several authors have argued that this apparent discrepancy could be resolved by considering some suppressing process for star formation, such as gas heating by an intergalactic ionizing background or energy feedback from evolving stars. In whatever models relying on the suppression of galaxy formation, a typical galaxy-sized halo should contain numerous dark subhalos. Then, there is a possibility that a large amount of subhalos interact frequently with a stellar disk embedded in the center of a halo, so that the disk would be dynamically heated and thickened. On the other hand, an observed galactic disk is rather thin: the scale height (or half thickness) is only about $\\sim 250$ pc in the Galaxy. Likewise, recent observations of external disk galaxies (\\cite{kregel2002}) suggest that the observed scale height of a disk, $z_d$, is confined to some limiting value relative to the scale length of a disk, $R_d$, i.e, $z_d / R_d < 0.2$. This observed thinness of a disk provides important limits on the disk heating due to infalling satellites. T\\'{o}th \\& Ostriker (1992) analytically evaluated this effect and concluded that an observed disk like that of the Galaxy within the solar radius should have interacted with satellites with no more than 4~\\% of the present disk mass within the last 5 Gyr. Subsequent numerical simulations of an interaction between a disk and a single satellite (e.g. \\cite{velazquez1999}) showed that their analytical estimation for the disk heating was somewhat too high because an actual interaction process is highly non-linear and more complicated than simplified analytical representation. Interactions with many subhalos would be much more complicated and thus require a more detailed analysis. \\citet{Font01} have conducted numerical simulation of interaction between a disk and numerous subhalos based on the CDM models. They concluded that the effect of subhalos on a disk is rather small, and therefore subhalos do not conflict with the presence of a thin disk since their orbit seldom take them near the disk. However, it is worth noting that in their simulation the initial scale height of a disk (700 pc) is already thick compared with the observed one in the Galaxy ($\\sim250$ pc), thereby leading to possibly the underestimation of the disk heating effect. Their simulation is also limited to only one realization of subhalos; it is yet unclear whether the derived weak effect of subhalos on a disk is general or not. \\citet{ardi2003} have investigated more details in this disk heating by subhalos. They found that a more massive subhalo is more effective to heat the disk than a less massive one. However, in their calculation subhalos are represented by rigid bodies which never lose their mass irrespective of tidal effects of a host galaxy, so that the disk heating is overestimated. Also, the applicability of their result to an actual disk, especially, to what extent accretion events of subhalos into a disk are allowed remains unclear. Our aim in this paper is thus to set more useful limits on the dynamical interaction between numerous subhalos and a galactic disk. For this purpose, we conduct a series of numerical simulations, in which a self-gravitating disk is embedded in a dark halo containing many subhalos. In this work we set an initially thin disk with the scale height of 250 pc, in contrast to previous numerical studies starting from the scale height of $\\sim 700$ pc much larger than the observed one (\\cite{velazquez1999}; \\cite{Font01}). Several models for the system of subhalos in a host halo are taken into account in terms of their mass function, spatial distribution, and velocity distribution. We also consider two different models for the internal density distribution of subhalos: point-mass and extended-mass models. In the latter model, subhalos are affected by a tidal field of a host galaxy so that they lose their mass in the course of their orbital motions. Based on our simulations, we investigate the dependence of the disk heating on the model parameters and apply our analysis to understanding an observed thin disk in the context of the disk heating by subhalos. This paper is organized as follows. In \\S~\\ref{sec:model} we describe our galaxy model which is composed of halo, bulge, and disk components. The models of subhalos are also described in this section. In \\S~\\ref{sec:result} we present the results of our numerical simulations. In \\S~\\ref{sec:discussion} we analyze our results and present our prediction for the relation between the disk heating by subhalos and an observed thin or thick disk. Finally, in \\S~\\ref{sec:conclusion} we present our conclusions. ", "conclusions": "\\label{sec:conclusion} We summarize our conclusions as follows: \\begin{itemize} \\item The dynamical effects of subhalos on a disk are represented by the relation between the change of the disk scale height $\\Delta z_d$ (measured at the disk edge $R=3R_d$) and individual masses of subhalos $M_{\\mathrm{sub}}$, i.e., $\\Delta z_d/R_d \\simeq 8 \\sum_{j=1}^N (M_{\\mathrm{sub},j}/M_d)^2$, where $R_d$ is a disk scale length, $M_d$ is a disk mass, and $N$ is the total number of accretion events of subhalos inside a disk region ($\\le 3R_d$). \\item If subhalos with the total mass of more than 15~\\% disk mass interact with a disk, then the disk thickness is made larger than the observed range. \\item A less massive disk with smaller circular velocity $V_c$ is found to be more affected by subhalos than a disk with larger $V_c$, which is in agreement with the observed properties of a thin disk. \\item Stars in a significantly thickened disk by subhalos appear to be well mixed and show a vertical gradient in their rotation velocity, being similar to the observed properties of the thick disk in the Galaxy. \\end{itemize} We note that the relation (\\ref{eq:relation2}) we have obtained here is universal and thus useful for the applications to any relevant issues, including the dynamics of an evolving stellar disk at the center of a growing dark halo. Such detailed studies of a galactic disk in comparison with recently increasing datasets of a remote disk galaxy will be of great importance and is left to future work. \\bigskip The numerical computations reported here were carried out on GRAPE systems (project ID: g05b05) kindly made available by the Astronomical Data Analysis Center (ADAC) at the National Astronomical Observatory of Japan (NAOJ). This work has been supported in part by a Grant-in-Aid for Scientific Research (15540241, 17540210) from the Japanese Ministry of Education, Culture, Sports, Science and Technology." }, "0609/astro-ph0609743_arXiv.txt": { "abstract": "The Baikal Neutrino Telescope has been operating in its NT200 configuration since April, 1998. The telescope has been upgraded in April, 2005 to the 10\\,Mton scale detector NT200+. It's main physics goal is the detection of signals from high energy neutrino cascades. NT200+ reaches a 3-year sensitivity of $2 \\times 10^{-7}$cm$^{-2}$s$^{-1}$sr$^{-1}$GeV for an all-flavor diffuse cosmic $E^{-2}$ neutrino flux for energies 10$^2$~TeV~$\\div$~10$^5$~TeV. Desgin and sensitivity of NT200+ are described. NT200+ is forming the basic building block of a future km3-scale (Gigaton-Volume) Baikal Telescope. Research and development work on that next stage detector has started. ", "introduction": " ", "conclusions": "" }, "0609/astro-ph0609269_arXiv.txt": { "abstract": "We obtained 98 $R$-band and 18 $B$, $r'$, $i'$ images of the optical afterglow of GRB 060526 ($z=3.21$) with the MDM 1.3m, 2.4m, and the PROMPT telescopes in Cerro Tololo over the 5 nights following the burst trigger. Combining these data with other optical observations reported in GCN and the \\swift-XRT observations, we compare the optical and X-ray afterglow light curves of GRB 060526. Both the optical and X-ray afterglow light curves show rich features, such as flares and breaks. The densely sampled optical observations provide very good coverage at $T>10^4$~sec. We observed a break at $2.4\\times10^5$~sec in the optical afterglow light curve. Compared with the X-ray afterglow light curve, the break is consistent with an achromatic break supporting the beaming models of GRBs. However, the pre-break and post-break temporal decay slopes are difficult to explain in simple afterglow models. We estimated a jet angle of $\\theta_j \\sim 7^{\\circ}$ and a prompt emission size of $R_{prompt} \\sim 2\\times10^{14}$~cm. In addition, we detected several optical flares with amplitudes of $\\Delta m \\sim 0.2$, 0.6, and 0.2~mag. The X-ray afterglows detected by \\swift\\ have shown complicated decay patterns. Recently, many well-sampled optical afterglows also show decays with flares and multiple breaks. GRB~060526 provides an additional case of such a complex, well observed optical afterglow. The accumulated well-sampled afterglows indicate that most of the optical afterglows are complex. ", "introduction": "In the fireball model \\citep{meszaros97,sari98}, the afterglow emission of gamma-ray bursts (GRBs) are thought to be synchrotron emission in the external shocks. After the launch of \\swift\\ \\citep{gehrels04}, with its rapid localization of GRBs and the dedicated on-board XRT instrument, the afterglow models can be tested extensively with the regularly obtained XRT light curves. More than half of the \\swift-XRT light curves show complicated decay patterns with multiple breaks and giant X-ray flares \\citep{burrows05,nousek06,obrien06}. New ingredients were added to the models to interpret these features \\citep[e.g.,][]{zhang06,nousek06,panaitescu06a}. Compared with the large number of \\swift-XRT afterglows, only a few bursts have good optical afterglow coverage, which limits the multi-wavelength study of GRB afterglows. Moreover, a large fraction of the well-studied optical afterglows also show complicated behaviors (e.g., Guidorzi et al. 2005; Blustin et al. 2006; Rykoff et al. 2006; Stanek et al. 2006), challenging the simple, smooth-decay afterglow models. Another important aspect of the models is that GRBs are thought to be collimated in jets, based on the achromatic breaks observed in many optical GRB afterglow light curves \\citep[e.g.,][]{stanek99}. Different jet models have been proposed for GRBs either under a uniform jet model \\citep[e.g.,][]{rhoads99,frail01,granot02} or a structured jet model \\citep[e.g.,][]{lipunov01,zhang02,rossi02,lloyd-ronning04,zhang04,dai05}. Some of these jet models can also unify the closely related phenomena of X-ray flashes with GRBs \\citep{yamazaki03,zhang04,lamb05,dai05}. Since the distinct signature that a jet imposes on GRB afterglow light curves (an achromatic break) is simply a geometric effect, it is important to test the wavelength independence across the broadest possible wavelength range, for example between optical and X-ray afterglow light curves. To date, the lack of wavelength dependence has only been confirmed across different optical bands. Recently, GRB~050525A \\citep{blustin06}, GRB~050801 \\citep{rykoff06a}, and GRB~060206 \\citep{stanek06} show possible achromatic breaks across optical and X-ray light curves. However, in GRB~050801, the break is interpreted as energy injection or the onset of the afterglow, and in GRB~060206, it is debated whether the break is achromatic \\citep{monfardini06}. As many XRT light curves show multiple breaks, it is not obvious which of them should be associated with the optical break and which of them interpreted as the jet break. Recently, \\citet{panaitescu06b,fan06} showed that some of the X-ray breaks (1--4 hours after the burst trigger) are chromatic from X-rays to optical bands. However, as the X-ray light curves have several breaks, it is possible that the achromatic jet break occurs at some later time. In addition, as many optical afterglows also show rich features such as flares and multiple breaks (e.g., Stanek et al. 2006), fits to poorly sampled optical light curves may not be reliable. In this paper, we report the optical follow-up of GRB 060526 with the MDM 1.3m, 2.4m telescopes and the PROMPT at Cerro Tololo, and our detection of an achromatic break across the optical and X-ray bands. GRB 060526 was detected by the BAT on board \\swift\\ at 16:28:30 UT on May 26, 2006 (Campana et al. 2006). The XRT and UVOT rapidly localized the burst location. The burst was followed up with ground-based telescopes by several groups. In particular, \\citet{berger06} reported the burst redshift of $z=3.21$. We organize the paper as follows. First, we describe the data reduction in \\S\\ref{sec:data}. In \\S\\ref{sec:evo}, we describe the evolution of the GRB afterglow and perform a comparison between the optical and X-ray light curves. Finally, we discuss our results in \\S\\ref{sec:dis}. ", "conclusions": "} We present well-sampled optical and X-ray afterglow light curves of GRB~060526. As discussed in the previous section, the evolution of the afterglow is complicated with multiple breaks and flares both in the optical and the X-ray bands. The combination of flares and incomplete data sampling present severe challenges to measuring the temporal decay slopes of the afterglow, even for a well-sampled burst such as GRB~060526. Below, we proceed by assuming that our analysis results are not significantly affected by these factors. We detected a possible achromatic jet break in the optical and X-ray afterglow light curves. Before this late-time break ($T\\sim2.4\\times10^5$~sec), the afterglow is consistent with many \\swift\\ afterglows \\citep{nousek06}. The X-ray light curve started with a steep decay ($\\alpha_1 = 2.3\\pm0.2$), which is interpreted as the tail of the prompt emission due to the ``curvature effect'' \\citep{kumar00}. The spectral and temporal indices are constrained as $\\alpha=\\beta+2$, which is consistent with the X-ray spectral index of $\\beta_1 = 0.55\\pm0.15$. The X-ray light curve then entered into a shallow decay stage with $\\alpha_2 \\sim 0.5$, which we interpret as energy injection. Then it decays into a normal afterglow stage with $\\alpha_3 = 1.14\\pm0.02$ as constrained from the optical observations. In addition, the X-ray light curve shows two huge flares which are commonly seen in \\swift\\ X-ray afterglows and are attributed to late time central engine activities. After the achromatic break, the afterglow enters a very steep stage with $\\alpha_4=3.4\\pm0.2$. The X-ray and optical-to-X-ray spectral indices after $5000$~sec are consistent with $\\sim 1$, suggesting that the optical and X-ray bands are on the same power-law segment of the spectral energy distribution. The optical spectral index is marginally consistent with the X-ray index, although the error-bar is large. The achromatic break observed in optical afterglows is traditionally interpreted as the jet break. As mentioned in the introduction, the achromatic break is not yet confirmed across optical and X-ray light curves. Here, we present a case in GRB~060526 where such an achromatic break is observed across the optical and X-ray afterglows, supporting the beaming model of the GRBs. However, the afterglow decay slopes before and after the break are hard to reconcile with simple afterglow-jet models. The late time decay slope after the jet break should follow $\\alpha = p$ \\citep{sari99}, where $p$ is power-law index for the electrons $N(\\gamma) \\propto \\gamma^{-p}$. The post-jet slope, $3.4\\pm0.2$, is too steep for a pre-jet slope of $1.14\\pm0.02$ under any combination of either constant or wind medium and relative positions between $\\nu$, $\\nu_m$, and $\\nu_c$. Since the achromatic break is most easily explained by a jet, it is possible that more complicated afterglow models are needed \\citep[e.g.,][]{panaitescu06b} with non-standard micro-physical parameters. Another possibility is that energy injection or flares continued contributing significant flux and significantly affected the temporal decay slope. We estimated the jet angle (half opening angle for uniform jets or observer's viewing angle for structured jets) using $t_j \\simeq 6.2(E_{52}/n)^{1/3}(\\theta_j/0.1)^{8/3}$~hr \\citep{sari99} and obtained $\\theta_j \\sim 7^{\\circ}$ assuming ambient density $n=1~$cm$^{-3}$. We further estimated the size of the $\\gamma$-ray prompt emission by combining the measured jet angle and the X-ray tail emission detected before 225~sec using $t_{tail} = (1+z)(R_{prompt}/c)(\\theta_j^2/2)$ \\citep{zhang06} and obtained $R_{prompt}\\sim2\\times10^{14}$~cm. We also observed multiple optical flares in the light curve of GRB~060526. Optical flares or re-brightenings have been observed in both pre-\\swift\\ \\citep[e.g., GRB~970508, GRB~021004, and GRB~030329,][]{galama98,lazzati02,bersier03,mirabal03,lipkin04} and \\swift\\ bursts \\citep[e.g., GRB~050525A, GRB~050820A, GRB~060206, GRB~060210, GRB~060605, GRB~060607, and GRB~061007,][]{blustin06,cenko06,stanek06,schaefer06,nysewander06,bersier06}. The fraction of bursts with optical flares seemed small. However, recently many well-sampled bursts show complex optical decay behaviors. The accumulating observations argue that it is possible that most of optical afterglows are complex and the appearance of simplicity was a consequence of poor sampling. There are several interpretations for the optical flares, such as models of density fluctuations, ``patchy shell'', ``refreshed shock'', and late central engine activities \\citep[e.g.,][]{jakobsson04,ioka05,gorosabel06}. We estimated the quantities $\\Delta t/t = 0.23\\pm0.08$, $0.75\\pm0.17$, and $0.22\\pm0.11$ and $\\Delta F_{\\nu}/F_{\\nu} \\sim 0.2$, 0.7, and 0.2, respectively, for the three flares detected in GRB~060526. The $\\Delta t/t$ values are small which do not favor the models of patchy shell and refreshed shock, since the model predictions are $\\Delta t/t > 1$ and $>1/4$ for these two models \\citep{ioka05}. The properties of the flares barely satisfy Ioka et al.'s constraint, $\\Delta F_{\\nu}/F_{\\nu} < 1.6 \\Delta t/t$, under the density fluctuation model. Recently, \\citet{nakar06} also modeled the effects of density fluctuations on the afterglow light curves and found that they cannot produce the sharp features observed in many bursts. Another possibility is that the flares (or breaks) indicate the onset of the afterglow \\citep{rykoff06a,stanek06} scaled with the isotropic energy as $T_{onset} \\propto E_{iso}^{1/3}$ \\citep{sari97}. The onset time also depends on the density of the ambient medium and the initial Lorentz factor that are more difficult to measure. We might expect a correlation between $E_{iso}$ and $T_{onset}$ for a large sample of bursts, or if the densities and Lorentz factors for the bursts only spread in a narrow range. We tested this hypothesis by plotting the two properties for bursts with optical flares in Fig.~\\ref{fig:et}, and did not detect positive correlation between the two properties. However, we notice the difference between flares that occur before the optical afterglow has decayed and those that occur afterward. The flares in GRB~050820A, GRB~060210, GRB~060605, GRB~060607, and GRB~061007 possibly belong to the category of flares that occur before the afterglow has faded, and they roughly follow the scaling between isotropic energy and flare time. However, a larger sample is needed to fully test the model. In any case, the flares in GRB~060526 are unlikely to be associated with the onset of the afterglow. It is possible that these flares are from late central engine activities, which can have arbitrary variabilities. However, we are open to other theoretical models which can be tested extensively with our well-sampled light curve." }, "0609/astro-ph0609082_arXiv.txt": { "abstract": "We follow up on our (Radhakrishnan \\& Deshpande, 2001: RD01) radically different interpretation of the observed structures and morphologies in the x-ray observations of the nebulae around young pulsars (PWNe). In our general model for PWNe (RD01), originally motivated by the Chandra observations of the Vela X-ray nebula, the bright arcs, the jet-like feature and the diffuse components in such nebulae can be explained together in detail, wherein the arcs are understood as traces of the particle beams from the two magnetic poles at the shock front. We consider this as important evidence for collimated particle beams from pulsars' magnetic poles. In this paper, we discuss the variability in the features in the Vela X-ray nebula observed by Pavlov et al. (2003), and assess the relevance and implication of our model to the observations on the Crab and other remnants. Our basic picture after incorporating the signatures of free precession of the central compact object can readily account for the variability and significant asymmetries, including the bent jet-like features, in the observed morphologies. The implications of these findings are discussed. ", "introduction": "} The superb capabilities of the Chandra telescope have over the last 6 years revealed several spectacular images in X-rays of the nebulae surrounding pulsars, highlighting and resolving the various spatial structures richly loaded with information about many aspects. While most, if not all, of these images bear a remarkable commonality that is emphasized by their overall symmetric morphology about what is readily identified as the direction of the projected rotation axis, the forms and proportions of the components appear to differ significantly. An overwhelming majority of the observers and theorists interpreting these observations seem to suggest and endorse the following basic picture. The jet-like features nearly along the symmetry axis, bisecting the arcs and the diffuse glow spread about them, are identified with collimated outflows of relativistic particles along the spin axis of the central compact remnant, a pulsar. The two arc-like features lie along circular rings highlighting shocks in which the energy of an outflowing equatorial wind is dissipated to become the source of synchrotron emission for the compact nebula and the incompleteness of the rings is attributed to preferential Doppler boosting of the emission in the forward direction. The two rings, if apparent, straddle the equator symmetrically, and the deficit of emission exactly in the equatorial plane is related to the fact that this is where the direction of a toroidally wrapped magnetic field changes sign i.e. the field may vanish there. However, Radhakrishnan \\& Deshpande (2001, hereafter RD01) have suggested an alternative interpretation that differs from this mainstream model {\\it in practically every aspect}, particularly those regarding the arcs and the jet-like features. The bright arcs, the jet-like feature and the diffuse components (for example, in the Vela X-ray nebula) are explained together in detail by our ``Rotating Vector model\" in which the arcs are understood as traces of the particle beams from the two magnetic poles at the shock front. We consider this as important evidence for collimated particle beams from pulsars, a point that we find necessary to reemphasize. In this paper, we follow up on the RD01 model to address the new clues provided by the variability of various features, such as those in the Vela X-ray nebula observed by Pavlov et al. (2001, 2003), and assess the relevance and implication of our model to the observations on other remnants, such as the Crab. In the next section, we begin with a brief summary of our model (RD01) for the x-ray nebulae around pulsars. In section 3, we revisit the Vela nebula story, now incorporating signatures of free precession of the central compact object, to explain the observed variability. We try to model, in section 4, the observed morphology of the X-ray nebula surrounding the Crab pulsar, and the significant asymmetries, including the bent jet-like features. The implications of these findings are discussed in the last section. \\noindent{ ", "conclusions": "} The cavity shapes suggested by our modeling of the observed x-ray nebulae might appear to differ significantly in the two cases (namely, the Vela and the Crab) discussed here. The part of the cavity profile sampled by the particle beams from the Crab pulsar appears to be significantly shallower than that in the case of the Vela pulsar. Given the age and the proper motion of the Vela pulsar, the relative dimensions of its x-ray arcs appear to favour an hour-glass shaped cavity, rather than an ellipsoid. The visibility of the jet-like feature apparently extending well beyond the diffuse component is not inconsistent with both of these components sharing a common origin, i.e. the latitudinal spread of a small fraction of the otherwise collimated particle flow, since the apparent relative brightness of the two components could differ significantly for intrinsic reasons, as well as those dictated by the viewing geometry. In our picture, the observed extents of the jet-like features, after due accounting of the projection effect, provide lower limits for the dimension of the cavity in the latitudinal direction, which appears to exceed the equatorial dimension, at least in the Vela case. The apparent asymmetry in the `jet' and the `counter-jet' extents in the Vela nebula is consistent with our expectation of the associated radiation being symmetric with respect to the direction of the magnetic axis, modified by the viewing geometry. If the jet-like emission were to be due to any physical flow of matter at relativistic speeds along the rotation axis (e.g. as estimated by Pavlov et al., 2003), the counter-jet should have been more prominent, both in its extent and its brightness, given that it would be pointing closer to our sightline. This is definitely not what is observed, and in fact, this aspect was noted as a cause for concern by Pavlov at al. (2003). Further, such a physical jet flow cannot escape its dissipation and consequent termination at the relevant region of the cavity wall where a cap-like emission feature should have been apparent, but is not observed. In their `physical jet' picture, the jet instabilities leading to bending etc. are expected to be due to and along the proper motion of the star. The Crab case is clearly inconsistent with this expectation, implying at the least that the apparent bending has little to do with the motion of the star. These inconsistencies and the other aspects discussed by RD01 argue strongly against a physical jet along the rotation axis. Apparent bends in the jet-like features in general, and particularly those with anti-symmetry about the equatorial plane, are interpreted in our model as due to possible free-precession. Its successful application to the Crab and the Vela pulsars, differing in the viewing geometry and rotation history, is not a chance coincidence. It would not surprise us therefore, if the bent jet-like feature in the x-ray nebula around pulsar B1509-58 (Gaensler et al., 2002) would have a similar interpretation. The variability of these and other features is then a natural consequence of such a rotational history, as illustrated by the multi-epoch observations of the Vela x-ray nebula. The variability seen in the Crab nebula at optical wavelengths would be another illustration of the same effect, though on much longer time-scales. If our assumed precession period of 14 years for the Crab pulsar is correct, a significant change in the orientation of the jet-like feature should be expected in the coming years. As commented earlier, the yet unexplored interplay between the glitches and possible precession would determine how the rotation/precession history evolves with time. Hence it is not at all clear whether free-precession, if any, would be unaffected through the glitch episodes, and even if it does, whether any clear signature of free-precession of the star would be apparent from the radio pulsar timing residuals obtained after fitting for period glitches and the associated recoveries commonly observed in young pulsars. It remains to be seen if any significant variation in the shape and intensity of the radio pulses from the Crab pulsar reveals a signature consistent with the rotational history of the star suggested by the above mentioned observations at high energies. There have been suggestions of free-precession in a few pulsars, but they are based on apparent changes in pulse shapes and intensities (e.g., DM96; Shabanova et al., 2001), or in pulse arrival times (Stairs et al., 2000). We wish to point out that a direct and independent way of probing any changes in the orientation of the rotation axis of the star, as in precession, is through monitoring of the polarization position angle (PA) sweep across the radio pulses, and measuring systematic changes, if any, particularly in the PA sweep-rate (Hari Dass \\& Radhakrishnan, 1975). We have indeed begun recently such a monitoring of the Vela pulsar using the Giant Meter-wave Radio Telescope (GMRT) in India. To summarize, we find compelling evidence for collimated particle beams from pulsars, and for free-precession of the Vela and the Crab pulsars, based on the Chandra observations of the respective x-ray nebulae and their apparent systematic variability. Quantitative estimates of the intensities \\& the locations of the different elements of the nebulae and their variability, when compared in the framework of our simple interpretation, should pave the way to important clues on the properties of the cavities created by the pulsars, and the energy spectra associated with the collimated particle beams." }, "0609/astro-ph0609561_arXiv.txt": { "abstract": "Young, rapidly rotating neutron stars could accelerate protons to energies of $\\sim 1$ PeV close to the stellar surface, which scatter with x-rays from the stellar surface through the $\\Delta$ resonance and produce pions. The pions subsequently decay to produce muon neutrinos. We find that the energy spectrum of muon neutrinos consists of a sharp rise at $\\sim 50$ TeV, corresponding to the onset of the resonance, above which the flux drops as $\\epsilon_\\nu^{-2}$ up to an upper-energy cut-off that is determined by either kinematics or by the maximum energy to which protons are accelerated. We predict event rates as high as 10-50 km$^{-2}$ yr$^{-1}$ from relatively young, close neutron stars. Such fluxes would be detectable by IceCube. ", "introduction": "A new window in high-energy astronomy is opening as existing neutrino detectors are improved and new ones are developed. Since neutrinos produced in astrophysical systems are unimpeded by interstellar matter on their way to Earth, detections will provide a new way to study the highest energy phenomena in the Universe. Astrophysical neutrinos are expected to arise in many environments in which neutrinos are produced by the decay of pions created through hadronic interactions ($pp$) or photomeson production ($p\\gamma$). Neutrinos may be produced by cosmic accelerators, like those in supernova remnants \\cite{pro98}, active galactic nuclei \\cite{lm00}, micro-quasars \\cite{dist02} and gamma-ray bursts \\cite{wb97,dai01}. To detect these neutrinos, several projects are underway to develop large-scale neutrino detectors under water or ice. AMANDA-II \\cite{ah04}, in the South Pole, and Baikal \\cite{ay06} are the two neutrino telescopes currently running, whereas ANTARES \\cite{ca03} and NESTOR \\cite{tz03} are under construction in the Mediterranean Sea. Those telescopes belong to a first generation with instrumented volume smaller than 0.02 $\\rm km^3$. The IceCube detector \\cite{ha06}, with a volume of about 1 $\\rm km^3$, is under construction on the same site of AMANDA. The NEMO project \\cite{mig06} is in its starting phase, and will be a cubic kilometer size detector located at Capo Passero, Southern Italy. As neutrino astronomy comes of age, it is important to get some idea of what the sources might look like to aid in their detection. Recently, we proposed \\cite{lb05,lb06} that young ($t_{age}\\lesssim 10^5 \\rm yr$) and rapidly-rotating neutron stars could be intense neutrino sources. Here we summarize this work. ", "conclusions": "Large-area neutrino detectors use the Earth as a medium for conversion of a muon neutrino to a muon, which then produces \\v{C}erenkov light in the detector. The conversion probability in the Earth is $P_{\\nu\\mu \\rightarrow \\mu} \\simeq 1.3 \\times 10^{-6} (\\epsilon_{\\nu}/\\rm 1\\,\\, TeV)$, where $\\epsilon_{\\nu}$ is the energy of the incident muon neutrino \\cite{ghs95}. The muon event rate is \\begin{equation} \\frac{dN}{dAdt} = \\int d\\epsilon_\\nu \\frac{d\\phi_\\nu}{d\\epsilon_\\nu} P_{\\nu_\\mu \\rightarrow \\mu}. \\label{rate} \\end{equation} Estimated conversion probabilities for the Crab, Vela and 6 other pulsars are given in Table 1 (column 8) for a characteristic acceleration length $L=0.1$. The final column gives the estimated event rates. Complete consideration of the kinematics gives event rates are a factor of $\\sim 10-30$ lower than estimated in our previous paper \\cite{lb05}. Neutrinos are produced at relatively high rates only if the protons are accelerated through the resonance close to the star. In this case, we obtain integrated count rates of several to $\\sim 50$ km$^{-2}$ yr$^{-1}$. Such count rates should be easily detected by IceCube, and possibly by AMANDA-II or ANTARES with integration times of about a decade (IceCube is planned to have replaced AMANDA-II by then) for depletion factors of $f_d\\simeq 1/2$. While the characteristics of the spectrum presented here are robust, we caution that the events rates we obtain are very rough upper limits, subject to many uncertainties. For example, we have assumed that the neutrinos are beamed into the same solid angle as the radio beam, which might not be a correct assumption. The radio beam is thought to be produced at about $10R$ \\cite{cor78}. In our model, the pions are produced much closer to the star. They then propagate to $\\sim 1000R$ before decaying to neutrinos. At this distance from the star, the field is unlikely to be dipolar, and it is difficult to say anything definite about the distribution of pion velocities in this region. If the neutrinos form a beam, the beam may be more or less collimated than the radio beam. If it is more collimated, the neutrino event rates could be higher than estimated here. Results of 807 d of data from AMANDA-II are now available \\cite{gr05}. AMANDA-II has detected 10 events (over a background of 5.4) from the direction of the Crab pulsar, with energies higher than 10 GeV. This result, though intriguing, is not statistically significant; IceCube will be able to confirm or refute this result. While it would be more exciting to see neutrinos from pulsars, the accumulation of null results over the next decade would be interesting as well; it would probably mean that photomeson production is ineffective or non-existent in the neutron star magnetosphere, thus providing a bound on the accelerating potential that exists near the neutron star surface." }, "0609/astro-ph0609611_arXiv.txt": { "abstract": "The high-latitude Galactic H~I cloud toward the extragalactic radio source 3C~225 is characterized by very narrow 21~cm emission and absorption indicative of a very low H~I spin temperature of about 20~K. Through high-resolution optical spectroscopy, we report the detection of strong, very narrow Na~I absorption corresponding to this cloud toward a number of nearby stars. Assuming that the turbulent H~I and Na~I motions are similar, we derive a cloud temperature of 20$^{+6}_{-8}$~K (in complete agreement with the 21~cm results) and a line-of-sight turbulent velocity of 0.37~$\\pm$~0.08~~km~s$^{-1}$ from a comparison of the H~I and Na~I absorption linewidths. We also place a firm upper limit of 45~pc on the distance of the cloud, which situates it well inside the Local Bubble in this direction and makes it the nearest-known cold diffuse cloud discovered to date. ", "introduction": "Recent H~I 21~cm absorption surveys have revealed a significant population of cold (T~$<$~40~K) diffuse clouds in the Galactic interstellar medium (ISM) \\citep{hei03,gib05,kav05}. Two of the most remarkable examples of such clouds were actually discovered long ago by \\citet{ver69} through their very narrow 21~cm emission. He found two closely-spaced, degree-sized patches of H~I gas at $l\\sim226\\arcdeg,b\\sim+44\\arcdeg$ (``Cloud A'') and $l\\sim236\\arcdeg,b\\sim+46\\arcdeg$ (``Cloud B'') each with line widths indicative of kinetic temperatures below 30~K. \\citet{kna72} subsequently mapped the 21~cm emission of these clouds at higher resolution and derived respective H~I spin temperatures of $\\approx$24~K and $\\approx$17~K for Cloud~A and Cloud~B. In another approach, \\citet{cro85} measured spin temperatures of $\\approx$20~K amd $\\approx$14~K from the Cloud~A and Cloud~B 21~cm absorption toward the extragalactic radio sources 3C~225 and 3C~237. As part of their Millenium Arecibo 21~cm Absorption-Line Survey, \\citet{hei03} revisited these two cold clouds in the context of new observations of 3C~225 and 3C~237. Utilizing the Leiden-Dwingeloo 21~cm sky survey data \\citep{har97}, they found that Cloud~A and Cloud~B are the predominant components of a narrow ($\\approx$2$\\arcdeg$), broken ribbon of cold H~I gas stretching over 20$\\arcdeg$ across the constellation Leo. Their observations of 3C~225a,b (6.3$\\arcmin$~separation) yield spin temperatures of 22~K and 17~K and H~I column densities of $1.7\\times10^{19}$~cm$^{-2}$ and $3.2\\times10^{19}$~cm$^{-2}$ for the intervening Cloud~A gas. Assuming thermal pressure equilibrium under these conditions leads to an extremely thin, sheetlike geometry for the cloud \\citep{hei03}. A more detailed physical assessment including the specific aspect ratio, actual size, and mass of the cloud is limited by the unknown cloud distance. In this {\\it Letter}, we present the first optical absorption-line study of the cold Leo clouds and conclusively show that Cloud~A is located well inside the Local Bubble. ", "conclusions": "Figure~3 illustrates the strength of the Na~I absorption at v$_{LSR}$~$\\approx$~3~km~s$^{-1}$ toward our program stars as a function of sky position relative to the 21~cm emission contour of the cold H~I gas \\citep{hei03} and the location of 3C~225 and 3C~237. It is clear from Figure~3 that the Na~I absorption is an excellent tracer of the H~I emission for this cold gas. In the case of comparisons with the 21~cm absorption observed toward 3C~225a,b, four stars with matching-velocity Na~I absorption (HD~84194, HD~84182, HD~83683, and HD~83509) are located within 1$\\arcdeg$ of this radio source. The mean linewidth of this Na~I absorption (and the standard error of the mean) is b~$=$~0.54~$\\pm$~0.10~km~s$^{-1}$ (0.90~$\\pm$~0.17~km~s$^{-1}$~FWHM) as compared with the 1.3~km~s$^{-1}$~FWHM H~I linewidth measured toward both 3C~225a and 3C~225b by \\citet{hei03}. Assuming that the turbulent H~I and Na~I motions are similar, a comparison of the H~I and Na~I linewidths through the expression $b^2=(2kT/m)+2v_t^2$ yields a cloud kinetic temperature of 20$^{+6}_{-8}$~K and a one-dimensional rms turbulent velocity (v$_t$) of 0.37~$\\pm$~0.08~km~s$^{-1}$. Thus, the Na~I measurements are completely consistent with the remarkably cold temperatures previously derived for Cloud~A from the H~I 21~cm spin excitation toward 3C~225. The Na~I observations also provide a stringent constraint on the distance of Cloud~A. The two closest program stars within the Cloud~A H~I contour are at {\\it Hipparcos} distances \\citep{per97} of 42.1$^{+1.8}_{-1.6}$~pc (HD~83683) and 42.8$^{+1.9}_{-1.7}$~pc (HD~85091) and they both exhibit strong Cloud~A Na~I absorption. Another nearby star at a distance of 41.5$^{+1.7}_{-1.6}$~pc (HD~83808) just outside the Cloud~A contour exhibits very weak Na~I absorption at the appropriate velocity. The only program star closer than these objects is at a distance of 39.2$^{+1.5}_{-1.4}$~pc (HD~80218) and it lies just outside the H~I contour of the lowest-longitude cold Leo cloud. Consequently, the lack of Na~I absorption toward HD~80218 is not conclusive with respect to the distance of this cloud. Although it is likely that the three cold Leo clouds illustrated in Figure~3 are at the same distance based on their similar velocities, narrow linewidths, and close position on the sky, the Na~I measurements are not definitive in this regard. Nevertheless, the Na~I absorption observed toward HD~83683, HD~85091, and HD~83808 does allow us to place a firm upper limit of 45~pc on the distance of Cloud~A. With this limit, Cloud~A is not only one of the coldest diffuse clouds discovered to date, it also becomes the nearest-known cold interstellar cloud to the Sun. The proximity of Cloud~A is especially interesting in that it places this cold cloud well inside the Local Bubble of hot ($\\sim$10$^6$~K), tenuous (n$_H$~$\\sim$~0.01~cm$^{-3}$) gas surrounding the Sun out to distances of $\\sim$100~pc or more \\citep{cox87}. The Local Bubble is believed to have originated through multiple supernovae explosions over the past $\\sim$5 to $\\sim$15~million years \\citep{smi01,bre06}. Utilizing Na~I absorption measurements toward $\\approx$1000 early-type (earlier than A5) stars with {\\it Hipparcos} distances, \\citet{lal03} have mapped the neutral gas ``edge'' of the Local Bubble across the sky. In the direction of the cold Leo clouds, they place this boundary at a distance of about 100~pc. Since our star sample in the Leo field of Figure~3 is much larger (33) than that (2) of Lallement et al.\\ (due in part to our inclusion of objects with later spectral types), we can explore this boundary in greater detail. Among the 17~program stars with d~$>$~157~pc, nine exhibit interstellar Na~I absorption at LSR velocities (ranging from $-$11.5 to $-$1.6~km~s$^{-1}$) distinct from those of the cold Leo clouds whereas none of the 16 stars with d~$<$~157~pc display such absorption. As illustrated in Figure~1, such distance distinctions are also apparent when the Na~I spectra are compared in localized sky regions. Allowing for patchiness in the more distant Na~I absorption and the larger stellar distance uncertainties past $\\approx$100~pc, the nearest neutral gas clouds beyond Cloud~A are at a distance between 100 and 150~pc. Thus, Cloud~A is indeed closer to the Sun than it is to the neutral gas boundary of the Local Bubble. Within the Local Bubble, there are known to be a number of warm, partially-ionized clouds which collectively have an average temperature of 6700~K and a mean thermal pressure (P/k) of 2300~K~cm$^{-3}$ \\citep{red04}. Although the thermal pressures of these clouds appear to be below that of the surrounding hot gas in the Local Bubble, they provide a benchmark in considering the physical characteristics of Cloud~A. Assuming that Cloud~A is in thermal pressure equilibrium at 2300~K~cm$^{-3}$, the 20~K temperature and mean N(H~I) of $2.5\\times10^{19}$~cm$^{-2}$ toward 3C~225 yield a cloud thickness of $2.2\\times10^{17}$~cm or 0.07~pc (this equilibrium thickness would be $\\approx$4~times less at the hot gas thermal pressure). If Cloud~A is at a distance of 40~pc, its $2\\arcdeg\\times7\\arcdeg$ H~I angular extent would then correspond to a physical size of $1.4\\times4.9$~pc, a very thin length-to-width-to-thickness aspect ratio of 70:20:1, and an H~I mass of about 1.4~solar masses. At the other extreme, if we assume that Cloud~A is at a distance ($\\approx$2~pc) where its projected width is equal to its equilibrium thickness of 0.07~pc, the cloud size would be $0.07\\times0.25$~pc with an H~I mass of about 3.6~Jupiter masses. In any case, given our measured v$_t$ of 0.37~km~s$^{-1}$, the line-of-sight turbulent crossing time over the equilibrium thickness of 0.07~pc is $1.8\\times10^5$~years. In other words, it appears that Cloud~A is a much more transient phenomenon than the Local Bubble and thus, is likely to have originated somehow in this seemingly hostile environment to cold clouds. Recently, several theoretical studies have proposed the formation of cold clouds with filamentary or sheet-like geometry through converging flows of warm interstellar gas \\citep{aud05,vaz06}. In particular, Vazquez-Semadeni et al.\\ find that they can come close to reproducing the \\citet{hei03} description of Cloud~A through transonic compression in colliding warm gas streams. They acknowledge that their timescale to produce this cold structure is rather long ($\\sim$1~Myr) but may be consistent with the observations if Cloud~A is overpressured and v$_t$ reflects the gas flow rather than turbulent motions. Such models are promising but will need to consider other factors such as magnetic fields and certainly thermal conduction in order to provide a full explanation of the nature and evolution of Cloud~A inside the hot Local Bubble. For example, \\citet{sla06} has shown that the evaporative and condensation time scales for such cold clouds are measured in millions of years even if immersed in a hot surrounding medium. It is also possible that the hot gas in the Local Bubble may be less pervasive than generally thought, especially given recent results suggesting significant foreground contamination by heliospheric soft X-ray emission \\citep{lal04}. In any case, the size of Cloud~A on the sky and its isolated nature make it ideal for further optical/UV absorption-line and H~I 21~cm emission-line studies of its distance, small-scale structure, velocity fields, and other physical characteristics such as dust depletion, electron density, and radiation field. A better understanding of the origin and survival of this nearby 20~K cloud inside the Local Bubble will almost certainly have a bearing on interpreting the continuing discoveries of such cold diffuse clouds elsewhere in the Galactic ISM." }, "0609/astro-ph0609427_arXiv.txt": { "abstract": "In the last few years, radio detection of cosmic ray air showers has experienced a true renaissance, becoming manifest in a number of new experiments and simulation efforts. In particular, the LOPES project has successfully implemented modern interferometric methods to measure the radio emission from extensive air showers. LOPES has confirmed that the emission is coherent and of geomagnetic origin, as expected by the geosynchrotron mechanism, and has demonstrated that a large scale application of the radio technique has great potential to complement current measurements of ultra-high energy cosmic rays. We describe the current status, most recent results and open questions regarding radio detection of cosmic rays and give an overview of ongoing research and development for an application of the radio technique in the framework of the Pierre Auger Observatory. \\vspace{1pc} ", "introduction": "About 40 years ago, Jelley et al. \\cite{Jelley1965} measured pulsed radio emission originating from extensive air showers (EAS) for the first time. As the radio technique proved to be too difficult to handle with the technical limitations of the 1960s and 1970s, however, the interest in radio detection of cosmic rays diminished completely within the following decade. It was only recently, now having powerful digital technology at our disposal, that the concept of radio detection of cosmic rays experienced its renaissance \\cite{FalckeGorham2003}. By now a number of new projects dedicated to the measurement of radio emission from EAS, most prominently the LOPES project \\cite{HornefferArena2005,Horneffer2006} and the CODALEMA project \\cite{Codalema}, have been established. The radio technique for measuring cosmic rays has a number of merits in its own right: it can watch for EAS with nearly 100\\% duty cycle, even in populated areas, and it measures a bolometric signal that is only very slightly attenuated in the atmosphere, thus allowing the observation of highly inclined showers. But naturally, the most interesting application is to combine the technique with other detection methods, in particular ground-based particle detector arrays and air fluorescence measurements. Each of these techniques yields different observables, and a combination of the methods allows so-called ``hybrid detection'' of cosmic rays, yielding much more information than the individual techniques alone. In this article, we review the goals, the status and the results so far gathered within the LOPES project followed by an outlook on the application of the radio technique on large scales for the measurement of ultra-high energy cosmic rays. ", "conclusions": "In the last few years, radio detection of cosmic ray air showers has once again become a very active field of research. LOPES, as one of the projects studying radio emission from EAS with modern digital technology, has made important contributions. Radio emission from EAS is proven to be coherent at 40--80~MHz, and the field strength correlation with geomagnetic angle strongly indicates that the radiation is of dominantly geomagnetic origin. Aspects such as the absolute field strength, the detailed angular correlations and the scale radius of the lateral dependence will soon be analysed with the new LOPES30 data. Our next-generation Monte Carlo code calculating the radio emission based on highly realistic CORSIKA-based air showers will also allow many new studies with unprecedented detail. Unlike 40 years ago, the study of radio emission today is making good progress. A possible application of the radio technique on large scales for hybrid detection of ultra-high energy cosmic rays is now being studied in the framework of the Pierre Auger Observatory. These activities are still in a relatively early phase, but with the experience gained in the LOPES and CODALEMA experiments, the challenges to unlock the great potential of the radio technique on large scales can be tackled. \\vspace{0.4cm} {\\small\\noindent Acknowledgements: LOPES was supported by the German Federal Ministry of Education and Research (Verbundforschung Astroteilchenphysik) and is part of the research programme of the Stichting voor Fundamenteel Onderzoek der Materie (FOM), which is financially supported by the Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO). The KASCADE-Grande experiment is supported by the German Federal Ministry of Education and Research, the MIUR of Italy, the Polish Ministry of Science and Higher Education and the Romanian National Academy for Science, Research and Technology.}" }, "0609/astro-ph0609388_arXiv.txt": { "abstract": "{ We study the distribution of fermionic dark matter at the center of galaxies using NFW, Moore and isothermal density profiles and show that dark matter becomes degenerate for particle masses of a few {\\rm keV} and for distances less than a few parsec from the center of our galaxy. A compact degenerate core forms after galaxy merging and boosts the growth of supermassive black holes at the center of galaxies. To explain the galactic center black hole of mass of $\\sim 3.5 \\times 10^{6}M_{\\odot}$ and a supermassive black hole of $\\sim 3 \\times 10^{9}M_{\\odot}$ at a redshift of 6.41 in SDSS quasars, we require a degenerate core of mass between $3 \\times 10^{3} M_{\\odot}$ and $3.5 \\times 10^{6}M_{\\odot}$. This constrains the mass of the dark matter particle between $0.6 \\ {\\rm keV}$ and $82~{\\rm keV}$. The lower limit on the dark matter mass is improved to ~{\\rm 7~keV} if exact solutions of Poisson's equation are used in the isothermal power law case. We argue that the constrained particle could be the long sought dark matter of the Universe that is interpreted here as a sterile neutrino. ", "introduction": "Precision observations of the cosmic microwave background and of large scale structure confirm the picture in which 96\\% of the matter density of the Universe is made of dark energy and dark matter (DM), which could be revealed only by their gravitational interaction (Spergel et al. \\cite{spergel03}). The nature of these two forms of matter is still unknown. Many candidates have been proposed for DM. These include Cold Dark Matter (CDM) particles of masses heavier than $1 {\\rm GeV/c^{2}}$ (Bertone et al. \\cite{bertone05}) and Warm Dark Matter (WDM) particles such as sterile neutrinos (Dodelson \\& Widrow \\cite{dw94}). Recently, there has been a renewed interest in sterile neutrinos as candidates for DM as they could be a natural extension of the minimal standard model (MSM) of electroweak interactions (Weinberg \\cite{weinberg67}; Glashow \\cite{glashow61}). One of these sterile neutrinos could be the DM, while the other two could help explain baryogenesis (Asaka \\& Shaposhnikov \\cite{asaka05}; Shaposhnikov \\cite{shapo06}). Numerical cosmological N-body simulations suggest DM density profiles which follow $\\rho \\sim r^{-\\gamma}$ law, with $\\gamma \\approx 3$ in the outer parts of the halos and $1 \\stackrel {\\textstyle <}{\\sim} \\gamma \\stackrel {\\textstyle <}{\\sim} 2$ inside a few kpc (Navarro et al. \\cite{nfw97}; Moore et al. \\cite{moore98}, \\cite{moore99}; Klypin et al. \\cite{klypin02} and Power et al. \\cite{power03}). Although the Milky Way is well studied in the inner 3 - 10 kpc region, little is known about the DM distribution on smaller scales, i.e. $r \\stackrel {\\textstyle <}{\\sim}0.1 \\ {\\rm pc}$ where the gravitational potential of baryons dominates over DM. In addition, there is a mounting evidence that a black hole of mass $M \\sim 3.5 \\ 10^{6}M_{\\odot}$ dominates the mass distribution in the inner one parsec of the Galaxy (Sch\\\"odel et al. \\cite{schodel03}; Ghez et al. \\cite{ghez05}). Thus the investigation of the distribution of DM around the black hole in the Galactic center is important as it could give some hints on the nature of the DM particle (Gnedin \\& Primack \\cite{gnedin04}; Bertone \\& Merritt \\cite{bm05}). Recently, we have established that the inner DM density profile scales as $\\rho \\sim r^{-3/2}$ under the assumption that the black hole feeds from degenerate fermionic DM (Munyaneza \\& Biermann \\cite{mb05}). It was then shown that a DM particle of order of $m_{s} \\sim 10$ keV could explain the growth of supermassive black holes of $10^{6}$ to $10^{9}M_{\\odot}$ from stellar seed black holes. It is interesting to note that a sterile neutrino mass in overlapping mass range, with a small mixing angle with the active neutrinos, has also been suggested by Kusenko (\\cite{kusenko04}) to explain the high velocities of pulsars at birth in supernovae. Moreover, a DM particle mass in the keV range can be constrained from X-ray background studies (Drees \\& Wright \\cite{dw00}; Abazajian et al. \\cite{abaza01} and Dolgov \\& Hansen \\cite{dh02}). Here, it is worth mentioning that an upper limit of a sterile neutrino mass of $6.3 \\ {\\rm keV}$ from X-ray background seems to be in conflict with the lower limit of $14 \\ {\\rm keV}$ from SDSS Lyman-alpha forest (Abazajian \\& Koushiappas \\cite{ak06}). However, the non-thermal phase distribution of the DM particles has the potential to modify the Lyman alpha limits, while leaving the X-ray limits intact, so eliminating the contradiction with this additional degree of freedom. Depending on the specific model for the production of the DM particles, their initial phase space distribution is possibly largely sub-thermal. Recently, Biermann \\& Kusenko (\\cite{bk06}) established that the decay of a such sterile neutrino could speed up the formation of molecular hydrogen and boost the early star formation and reionisation in agreement with the WMAP 3-year results (Spergel et al. \\cite{spergel06}). The purpose of this Letter is to study the constraints on the DM particle mass in the central region of the galaxy. Given that there is still uncertainty about the DM profile around the central black hole, we will assume standard Navarro-Frenk-White (NFW), Moore or isothermal gas sphere profiles and investigate what happens if the DM particles become degenerate and form a degenerate dark matter star also called a fermion ball (Munyaneza \\& Viollier \\cite{mv02}; Munyaneza \\& Biermann \\cite{bm05}). We then get the lower limits on the DM particle mass by assuming that the mass of the fermion ball cannot be more than the mass of $3.5 \\times 10^{6}M_{\\odot}$ for the Galactic center black hole. An upper limit to the DM mass will be established from the condition that the mass of the fermion ball should be greater than about $3 \\times 10^{3}M_{\\odot}$. The choice of this mass comes from the growth mechanism arguments to form supermassive black holes of mass of $10^{9}M_{\\odot}$ in SDSS quasars (Munyaneza \\& Biermann \\cite{mb05}). Moreover, a mass of $3 \\times 10^{3} M_{\\odot}$ is the upper limit of masses for which there is no black hole in galaxies such as M33 (Gebhardt et al. \\cite{geb01}, \\& Barth et al. \\cite{barth05}). Thus the Galactic center provides us with a fertile testing ground for DM theories and cosmological structure evolution. ", "conclusions": "We have determined the constraints on the DM mass under the assumption that the NFW, Moore or isothermal density profiles become degenerate near the center. Assuming that the mass of the degenerate fermion ball is between about $3 \\times 10^{3}M_{\\odot}$ and $3.5\\times 10^{6}M_{\\odot}$, we have found that the mass of the DM particle should have a lower limit of about 7 keV in the isothermal power law case. The above limit on the DM mass is obtained from a simple model which does not take into account baryonic matter and a proper model for DM with anisotropic distribution in phase space (Evans \\& An \\cite{evans06}; Hansen \\& Stadel \\cite{hs06}) and baryonic matter would modify these limits. The obtained mass range is in full agreement with the Tremaine-Gunn lower bound of about 0.5 {\\rm keV }on the mass of any fermionic DM when applied to dwarf spheroidal galaxies (Tremaine \\& Gunn \\cite{tg79}). The constraints obtained on DM particle mass overlap with the upper limit on sterile neutrino masses of about $8 \\ {\\rm keV}$ to obey the X-ray emission constraints from the Virgo cluster observations (Abazajian \\cite{abaza05}). In addition, a sterile neutrino mass in the mass range of 2-20 keVs was derived to explain the high velocities up to 1000 km/s experienced by pulsars at birth in supernovae (Kusenko \\cite{kusenko04}; Barkovitch et al. \\cite{barko04} and Fuller et al. \\cite{fuller03}). Moreover, the obtained constraints agree well with the lower bound of $m_{s} \\approx 2 \\ {\\rm keV}$ on the mass of sterile neutrinos obtained from the analysis of the Lyman- $\\alpha$ forest data (Hansen et al. \\cite{hansen02}; Viel et al. \\cite{viel05}, and Abazajian \\cite{abaza05}). Recently, Biermann \\& Kusenko (\\cite{bk06}) have shown that the decay of such a sterile neutrino could help initiate star formation in the early Universe and the detection of such an X-ray line would confirm whether such a sterile neutrino exists or not and this could be done by observing the X-ray decay line using XMM and Chandra satellites (Boyarski et al. \\cite{boya06}, Riemer-S$\\o$rensen et al. \\cite{riemer06}). To summarize, we find that NFW, Moore and isothermal density profiles become degenerate for fermion DM particle of ${\\rm keV}$ masses at a size of the order parsec. Assuming that supermassive black holes grow from degenerate fermion cores of masses of a few $10^{3} M_{\\odot}$ to $\\sim 10^{6}M_{\\odot}$ at galactic centers, we have found that the sterile neutrino mass should be in the range from 0.6~{\\rm keV} to about 82~{\\rm keV}, with an improved lower limit of about 7 keV in the isothermal power law density profile. As this range of sterile neutrino masses overlaps with other results on sterile neutrino masses discussed in this Letter, this leaves the sterile neutrino to be an excellent candidate for DM. The formation and growth details of degenerate fermion balls is beyond the scope of this Letter and will be discussed in another paper." }, "0609/astro-ph0609207_arXiv.txt": { "abstract": "We briefly discuss the past, present, and future state of astronomical science with laser guide star adaptive optics (LGS AO). We present a tabulation of refereed science papers from LGS AO, amounting to a total of 23 publications as of May 2006. The first decade of LGS AO science (1995--2004) was marked by modest science productivity ($\\approx$1~paper/year), as LGS systems were being implemented and commissioned. The last two years have seen explosive science growth ($\\approx$1~paper/month), largely due to the new LGS system on the Keck~II 10-meter telescope, and point to an exciting new era for high angular resolution science. To illustrate the achievable on-sky performance, we present an extensive collection of Keck LGS performance measurements from the first year of our brown dwarf near-IR imaging survey. We summarize the current strengths and weaknesses of LGS compared to {\\sl Hubble Space Telescope}, offer a list of desired improvements, and look forward to a bright future for LGS given its wide-scale implementation on large ground-based telescopes. ", "introduction": "\\label{sect:intro} % Astronomers have envisioned using laser guide star adaptive optics (LGS AO) to achieve diffraction-limited observations from ground-based telescopes for over two decades\\cite{1985A&A...152L..29F, 1987Natur.328..229T, 1994OSAJ...11..263H}. (See Refs.~\\citenum{2004aoa..book.....R} and~\\citenum{1998aoat.conf.....H} for a historical review of LGS AO development.) The realization of these visions has been an arduous effort, but we are now entering a new epoch as LGS systems are commissioned on the largest ground-based telescopes. The scientific promise of near diffraction-limited imaging and spectroscopy from the ground over most of the sky is finally being realized. At this key juncture, the purpose of this paper is to briefly review published LGS science to date; to provide a snapshot of the science that is being done with LGS AO; and to look ahead to a future path where LGS AO is a ubiquitous tool for observational astronomy. \\begin{table}[t] \\vskip -0.1in \\caption{Refereed science papers from LGS AO as of May 2006, listed by publication date. Some titles have been truncated and some title words abbreviated (``LGS'' and ``AO''). ``Field'' gives the area of study: ``SS'' = solar system, ``Gal'' = galactic, ``Xgal'' = extragalactic. ``$N_{obj}$'' indicates the number of science targets/fields observed with LGS.} \\label{tab:bib} \\small \\begin{center} \\vskip -0.2in \\begin{tabular}{c|p{1.2in}|p{2.75in}|p{0.33in}|c|c|c} \\hline \\hline \\rule[-1ex]{0pt}{3.5ex} \\# & Authors, Journal & Title & Facilty & Field & $\\lambda\\lambda$ & $N_{obj}$ \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 1 & McCullough \\etal\\ 1995, ApJ & Photoevaporating Stellar Envelopes Observed with Rayleigh Beacon Adaptive Optics & SOR & Gal & H$\\alpha$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 2 & Christou \\etal\\ 1995, ApJ & Rayleigh Beacon AO Imaging of ADS 9731: Measurements of the Isoplanatic Field of View & SOR & Gal & $IJH$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 3 & Drummond \\etal\\ 1995, ApJ & Full AO Images of ADS 9731 and $\\mu$ Cassiopeiae: Orbits and Masses & SOR & Gal & $IJH$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 4 & ten Brummelaar \\etal\\ 1996, AJ & Differential Binary Star Photometry Using the AO System at Starfire Optical Range & SOR & Gal & $ri$ & 10 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 5 & Glenar \\etal\\ 1997, PASP & Multispectral Imagery of Jupiter and Saturn Using AO and Acousto-Optic Tuning & SOR & SS & 0.7--1.0 \\micron & 2 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 6 & Koresko \\etal\\ 1997, ApJ & A Multiresolution Infrared Imaging Study of LkH$\\alpha$ 198 & SOR & Gal & $H$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 7 & Drummond \\etal\\ 1998, Icarus & Full AO Images of Asteroids Ceres and Vesta: Rotational Poles \\& Triaxial Ellipsoid Dimensions & SOR & SS & $i$ & 2 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 8 & Barnaby \\etal\\ 2000, AJ & Measurements of Binary Stars with the Starfire Optical Range AO System & SOR & Gal & $ri$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 9 & Hackenberg \\etal\\ 2000, AA & Near-Infrared AO Observations of Galaxy Clusters: Abell 262 at z=0.0157 ... & Calar Alto & Xgal & $K$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 10 & Perrin \\etal\\ 2004, Science & Laser Guide Star AO Imaging Polarimetry of Herbig Ae/Be Stars & Lick & Gal & $JHK_S$ & 2 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 11 & Melbourne \\etal\\ 2005, ApJL & Merging Galaxies in GOODS-S: First Extragalactic Results from Keck Laser AO & Keck & Xgal & $K^\\prime$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 12 & Gal-Yam \\etal\\ 2005, ApJL & A High Angular Resolution Search for the Progenitor of the Type Ic Supernova 2004gt & Keck & Xgal & $K_S$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 13 & Brown \\etal\\ 2005, ApJL & Keck LGS AO Discovery and Characterization of a Satellite to Large Kuiper Belt Object 2003 EL61 & Keck & SS & $K^\\prime$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 14 & Muno \\etal\\ 2005, ApJ & A Remarkable Low-Mass X-ray Binary within 0.1 pc of the Galactic Center & Keck & Gal & $K^{\\prime}L^{\\prime}$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 15 & Liu \\etal\\ 2005, ApJ & Kelu-1 is a Binary L Dwarf: First Brown Dwarf Science from Laser Guide Star AO & Keck & Gal & $JHK^\\prime$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 16 & Cohen \\etal\\ 2005, ApJL & To Be or Not to Be: Very Young Globular Clusters in M31 & Keck & Xgal & $K^\\prime$ & 6 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 17 & Ghez \\etal\\ 2005, ApJ & The First LGS AO Observations of the Galactic Center: Sgr A*'s Infrared Color ... & Keck & Gal & $K^{\\prime}L^{\\prime}$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 18 & Marchis \\etal\\ 2006, Nature & A Low Density of 0.8 g cm$^{-3}$ for the Trojan Binary Asteroid 617 Patroclus & Keck & SS & $HK^\\prime$ & 1\\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 19 & Brown \\etal\\ 2006, ApJ & Satellites of the Largest Kuiper Belt Objects & Keck & SS & $K^\\prime$ & 4 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 20 & Krabbe \\etal\\ 2006, ApJL & Diffraction Limited Imaging Spectroscopy of the SgrA* Region Using OSIRIS & Keck & Gal & 2.0--2.4 \\micron & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 21 & Gelino \\etal\\ 2006, PASP & Evidence of Orbital Motion in Binary Brown Dwarf Kelu-1AB & Keck & Gal & $HK^\\prime$ & 1\\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 22 & Liu \\etal\\ 2006, ApJ & SDSS J1534+1615AB: A Novel T Dwarf Binary Found with Keck LGS AO and the Role of Binarity in the L/T Transition & Keck & Gal & $JHK^\\prime$ & 1 \\\\ \\hline \\rule[-1ex]{0pt}{3.5ex} 23 & Sheehy \\etal\\ 2006, ApJ & Constraining the AO PSF in Crowded Fields: Measuring Photometric Aperture Corrections & Keck & Xgal & $H$ & 1 \\\\ \\hline \\end{tabular} \\end{center} \\end{table} ", "conclusions": "More than two decades since its conception, LGS AO is now entering a new phase in its growth, opening a new era in high angular resolution science. The first decade of astronomical science was marked by modest science productivity, as these LGS AO systems were commissioned and optimized ($\\approx$1~science paper/year). Benefiting from the fruits of this effort, the LGS AO system on the Keck~II 10-meter telescope has had a highly successful first 1.5 years of science ($\\approx$1~paper/month). Several LGS systems are planned to come online in the next two years, including those at the Gemini-North (2006), VLT (2006), Palomar 5-meter (2006), MMT (2006), Subaru (2007), Gemini-South (2007), and the Keck~I (2008) telescopes. Given the size and quality of the science communities associated with these new AO systems, we can look forward to even more significant growth in LGS science soon ($\\approx$1~paper/week?). We conclude with a non-comprehensive wishlist for future LGS developments, in order to highlight ongoing efforts and to review some outstanding challenges: \\begin{itemize} \\item {\\em Field of view:} Single LGS systems produce a corrected field of view the size of the isoplanatic angle. In fact, nearly all LGS science thus far has been restricted to objects that span $\\lesssim$10\\arcsec\\ across (and most typically $\\lesssim2$\\arcsec). The upcoming Gemini-South multi-conjugate AO system is notable, as it will be the first LGS AO system to correct much larger fields of view ($\\approx$1--2\\arcmin). \\item {\\em Observing efficiency:} Nighttime LGS AO operations currently require many more personnel compared to regular seeing-limited observations. (See also paper 6270-12 by LeMignant \\etal\\ in this Proceedings.) Greater automation will reduce this burden on telescope staff and also should lead to greater efficiency, as observing procedures are streamlined. For ``survey-style'' science programs, in some sense the science return is proportional to the number of targets observed; therefore more efficient observing is a significant benefit. \\item {\\em Robust real-time performance predictions:} LGS image quality varies depending on nightly weather conditions, laser performance, AO performance, sodium layer density and structure, and tiptilt star properties. To maximize the science return, it would be desireable to be able to robustly predict LGS performance for any given target on any given night. This capability would allow observations to be tailored to achieve the desired science goal. Queue-scheduled observing is a key element here, but also better real-time understanding of seeing conditions, LGS performance and the conditions in the sodium layer are need. \\item {\\em Near-IR tiptilt sensors:} While much of the sky is available for LGS, the most obscured regions (\\eg, star-forming regions) are not due to the lack of optically visible tiptilt stars. Tiptilt sensors working at near-IR wavelengths would open the door to studying the youngest stages of star and planet formation. \\item {\\em High quality catalogs for tiptilt stars:} Our brown dwarf imaging survey finds about 1 in 10 tiptilt stars with $R\\lesssim18$~mag from the USNO-B1.0 catalog\\cite{2003AJ....125..984M} are unsuitable, either because they turn out to be faint galaxies or they turn out to be much fainter than the reported magnitudes. The Pan-STARRS project\\cite{2002SPIE.4836..154K} will provide precise, multi-band photometry over the entire sky visible from Hawaii, with the initial PS-1 telescope beginning operations later this year. Combined with the SkyMapper Telescope\\cite{2005AAS...206.1509S} in the southern hemisphere, high quality all-sky catalogs should be available in a few years for robust selection of tiptilt stars. \\item {\\em Improved PSF stability and characterization:} The time-variability of the LGS PSF will remain a significant concern for the foreseeable future. Post-processing software techniques can provide some immediate assistance, \\eg, deconvolution and/or PSF modeling techniques.\\cite{1993ApJ...415..862J, 2000A&AS..147..335D, 2006astro.ph..4551S} Development of algorithms to use AO telemetry data to estimate the real-time PSF would be a valuable capability for LGS.\\cite{2005PASP..117..847S} On a somewhat longer timescale, instruments tailored to handle the challenges of LGS AO imaging, such as dual-channel imaging systems, can circumvent this problem for some types of science programs. Finally, next-generation LGS AO systems should produce higher Strehl imaging, leading to more stable and well-behaved PSFs. \\end{itemize} While the technology is far from mature, LGS AO is entering a phase of rapid growth. There is little doubt that it will quickly become a key capability for a very broad range of astrophysics, spanning the nearest solar system bodies to the highest redshift galaxies and the entire universe in-between." }, "0609/astro-ph0609800_arXiv.txt": { "abstract": "Observations of two of the formaldehyde (H$_2$CO) masers (A and D) in Sgr~B2 using the VLBA+Y27 (resolution $\\approx$ 0\\farcs01) and the VLA (resolution $\\approx 9\\arcsec$) are presented. The VLBA observations show compact sources ($\\la 10$ milliarcseconds, $\\la 80$~AU) with brightness temperatures $>10^8$~K. The maser sources are partially resolved in the VLBA observations. The flux densities in the VLBA observations are about 1/2 those of the VLA; and, the linewidths are about 2/3 of the VLA values. The applicability of a core-halo model for the emission distribution is demonstrated. Comparison with earlier H$_2$CO absorption observations and with ammonia (NH$_3$) observations suggests that H$_2$CO masers form in shocked gas. Comparison of the integrated flux densities in current VLA observations with those in previous observations indicates that (1) most of the masers have varied in the past 20 years, and (2) intensity variations are typically less than a factor of two compared to the 20-year mean. No significant linear or circular polarization is detected with either instrument. ", "introduction": "The nature of Galactic formaldehyde (H$_2$CO) masers is a growing mystery. While hundreds of Galactic OH, H$_2$O, and CH$_3$OH masers are known, only five Galactic star-forming regions have associated H$_2$CO maser emission. To date, this emission is seen only in the $1_{10}\\rightarrow 1_{11}$ transition at 6 cm wavelength. Shortly after the discovery of the first H$_2$CO maser in NGC~7538 (Downes \\& Wilson 1974; Forster et al.\\ 1980), a radiative pumping model was proposed (Boland and de Jong 1981). The H$_2$CO masers discovered subsequently did not meet the conditions required for this mechanism (Gardner et al.\\ 1986; Mehringer, Goss, \\& Palmer 1994, hereafter MGP94; Hoffman et al.\\ 2003; hereafter H03). Thus, twenty-five years after the discovery of the first H$_2$CO maser, these sources remain rare and the excitation mechanism remains unknown. Sgr~B2, the northernmost component of the extended Sgr~B radio source, is located within a few hundred pc of the Galactic center (Reid et al.\\ 1988). (The distance to the Galactic center is assumed to be 8.5~kpc in this paper.) Sgr~B2 is comprised of three main star-forming complexes designated north (N), middle or main (M), and south (S), and many smaller H{\\sc ii} regions. The H$_2$CO masers occur throughout Sgr~B2, shown in Figure~\\ref{fig1}. The heating mechanisms and complex chemistry of the region are subjects of ongoing study (e.g., Gaume \\& Claussen 1990; Goicoechea et al.\\ 2004). Sgr~B2 contains nine individual H$_2$CO maser regions, several of which have multiple velocity components. All of the masers are unresolved at 1\\arcsec\\ angular resolution, except for maser C which MGP94 suggest consists of several masers blended within the beam. These regions are near H{\\sc ii} regions distributed over the $\\sim$3.6 arcmin$^2$ complex (MGP94). Whiteoak and Gardner (1983) and MGP94 designated the maser regions with letters (Fig.\\ \\ref{fig1}). The H$_2$CO masers are observed over the velocity range $+40$~\\kms$\\lesssim v_{\\rm LSR} \\lesssim +80$~\\kms, while other species, such as H$_2$O masers, are observed over a larger range $-30$~\\kms$\\lesssim v_{\\rm LSR} \\lesssim +120$~\\kms\\ (Kobayashi et al.\\ 1989; McGrath, De Pree, and Goss 2004). Of the nine maser regions observed in Sgr~B2 by MGP94, the G maser was shown to be time variable, at least quadrupling in intensity over 10~yr. (Similarly, H03 found one NGC~7538 feature to triple in intensity over $\\approx 10$~yr.) As initially noted by Whiteoak and Gardner (1983), all of the Sgr~B2 H$_2$CO masers lie close to OH, H$_2$O, CH$_3$OH, and NH$_3$ masers. For most of the masers in MGP94, the separation to an OH maser was less than 0.05 pc. Recent successes in search techniques for new masers (Araya et al.\\ 2004, 2005, 2006a) and in high-resolution observational techniques (H03) promise to provide empirical constraints for the development of a realistic model for the Galactic H$_2$CO maser emission. The necessary steps in compiling an empirical picture of the H$_2$CO emission in Sgr~B2 are (1) detailed imaging of the masers in order to quantify the intrinsic properties of the emission (e.g., brightness temperature), (2) assessment of intensity variability in the masers, and (3) precise astrometry for elucidating spatial relationships between the H$_2$CO masers and more common masers (OH, H$_2$O, CH$_3$OH). In this paper, we present new observations of the H$_2$CO masers in Sgr~B2 using the the Very Long Baseline Array (VLBA) and Very Large Array (VLA) of the NRAO\\footnote{The VLA and VLBA are components of the National Radio Astronomy Observatory (NRAO), a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc.}. ", "conclusions": "We present VLBA+Y27 images of the Sgr~B2 A and D H$_2$CO masers. The measured sizes ($\\la 80$~AU) and brightness temperatures ($>10^8$~K) are comparable to those found in other VLBI studies of H$_2$CO masers. However, about half of the flux density from these regions is resolved out with the VLBA data. A comparison between VLA and VLBA observations shows that the missing flux density exhibits a broader linewidth than the emission from the compact VLBI source. We demonstrate quantitatively the applicability of a core-halo model for these masers. We also present new VLA observations of the H$_2$CO masers in Sgr~B2 with improved velocity resolution. We have detected variability in several of the masers. H$_2$O and CH$_3$OH masers discovered near the H$_2$CO masers may indicate associations among the species, suggestive of a related origin. The association of H$_2$CO masers in Sgr~B2 with hot NH$_3$ shells proposed by Mart\\'{\\i}n-Pintado et al. (1999), together with the arguments for shock excitation of the maser region in NGC~7538 in H03, provide an encouraging stepping stone toward a solution of the problem of the excitation of these poorly understood masers." }, "0609/gr-qc0609076_arXiv.txt": { "abstract": "By the assumption that the thermodynamics second law is valid, we study the possibility of $\\omega=-1$ crossing in an interacting holographic dark energy model. Depending on the choice of the horizon and the interaction, the transition from quintessence to phantom regime and subsequently from phantom to quintessence phase may be possible. The second transition avoids the big rip singularity. We compute the dark energy density at transition time and show that by choosing appropriate parameters we can alleviate the coincidence problem. ", "introduction": "Recent observations suggest that the universe is undergoing an accelerated expansion \\cite{acc}. This acceleration may be explained by the assumption that $70\\%$ of the universe is filled by a perfect fluid with negative pressure, dubbed dark energy. Some present data seem to favor an evolving dark energy, corresponding to an equation of state (EOS) parameter less than $\\omega=-1$ at present epoch from $\\omega>-1$ in the near past \\cite{Bo}. Many candidates for dark energy has been proposed such as the cosmological constant \\cite{dark}: A constant quantum vacuum energy density which fills the space homogeneously, corresponding to a fluid with a constant EOS parameter $\\omega=-1$; dynamical fields with a suitably chosen potential to make the vacuum energy vary with time \\cite{field}, and so on. Recently, using holographic principle, a new candidate for dark energy which is independent of any specific field has been suggested \\cite{Li},\\cite{hol}. Based on the holographic principle (which relates the number of degrees of freedom of a physical system to the area of its boundary), in order to allow the formation of black holes in local quantum field theory, Cohen et al \\cite{cohen} proposed a relationship between UV and IR cutoff. This yields an upper bound on the zero-point energy density, which by a suitable choice of the infrared cutoff, can be viewed as the holographic dark energy density. In \\cite{Li}, three candidates for the infrared cutoff was proposed: the Hubble radius, the particle horizon and the future event horizon. There was shown that among these options only the future event horizon may be identified with the desired infrared cutoff. To study the coincidence problem and also to have other choices for the infrared cutoff, e.g. the Hubble radius, interaction between dark matter and dark energy \\cite{inter} may be considered in the holographic dark energy model. As we have mentioned, based on astrophysical data, we may take into account the possibility of $\\omega=-1$ (phantom divide line) crossing. Therefore dark energy models which can describe phantom divide line crossing, has been also studied vastly in the literature \\cite{divide}. The phantom like behavior of interacting holographic dark energy was studied in \\cite{abd}, where it was claimed that by selecting appropriate interaction parameters the transition from the dark energy EOS parameter $\\omega_D>-1$ to $\\omega_D<-1$ is possible. Despite this, in \\cite{antabd} it was shown that the dark energy effective EOS parameter cannot cross $\\omega_D^{eff.}=-1$. In this paper we consider interacting holographic dark energy model and study the ability of the model to describe the transition from quintessence to phantom regime and vice versa. After preliminaries in section two, where we introduce the interacting holographic dark energy model and some of its general properties used in the subsequent sections, in section three we study the possibility of crossing $\\omega=-1$. In section four we derive necessary conditions for existence of two transitions in our model. The first transition is from quintessence to phantom phase and the second the transition from phantom to quintessence regime. The importance of the second transition lies on the fact that it avoids the big rip singularity. We discuss also the behavior of Hubble parameter and dark energy density at transitions times. We use $\\hbar=G=c=k_B=1$ units throughout the paper. ", "conclusions": "In this paper we considered the holographic dark energy model with a general interaction between dark matter and dark energy, see (\\ref{3}). We took the infrared cutoff as a linear combination of the future an particle horizon see (\\ref{9}). We derived an expression for EOS parameter of the universe, $\\omega$, in terms of the ratio of dark energy density and total energy density of the flat FRW space-time, $\\Omega_D$, see (\\ref{20}). Using (\\ref{20}) and the expression obtained for time derivative of $\\Omega_D$ (\\ref{21}), and by assumption that the thermodynamics second law is still valid, we studied the possibility (and necessary conditions) for quintessence to phantom phase transition and vice versa with a differentiable Hubble parameter. Using some theorems about solutions of cubic equation satisfied by $\\Omega_D^{1/2}$ see (\\ref{25}), we showed that such transitions occur provided we appropriately choose the parameters of the system see(\\ref{47}) and (\\ref{50})." }, "0609/astro-ph0609031_arXiv.txt": { "abstract": "Although observations point to the neutrality and lack of currents on large scales in the universe, many mechanisms are known that can generate charges or currents during the early universe. We examine the question of survivability of relic charges and currents in a realistic model of the universe. We show that the dynamics of cosmological perturbations drive the universe to become electrically neutral and current-free to a high degree of accuracy on all scales, regardless of initial conditions. We find that charges are efficiently driven away in a time small compared to the Hubble time for temperatures $100 \\, \\mathrm{GeV} \\gtrsim T \\gtrsim 1 \\, \\mathrm{eV}$, while the same is true for currents at all temperatures $T \\gtrsim 1 \\, \\mathrm{eV}$. The forced neutrality relaxes constraints on the generation of an electric charge in the early universe, while the forced erasure of currents disfavors many mechanisms for the early origins of large-scale magnetic fields. ", "introduction": " ", "conclusions": "" }, "0609/astro-ph0609540_arXiv.txt": { "abstract": "Brighter Type~Ia supernovae (SNe~Ia) have broader, more slowly declining \\Bband\\ light curves than dimmer SNe~Ia. We study the physical origin of this width-luminosity relation (WLR) using detailed radiative transfer calculations of Chandrasekhar mass SN~Ia models. We find that the luminosity dependence of the diffusion time (emphasized in previous studies) is in fact of secondary relevance in understanding the model WLR. Instead, the essential physics involves the luminosity dependence of the \\emph{spectroscopic/color} evolution of SNe~Ia. Following maximum-light, the SN colors are increasingly affected by the development of numerous Fe~II/Co~II lines which blanket the \\Bband\\ and, at the same time, increase the emissivity at longer wavelengths. Because dimmer SNe~Ia are generally cooler, they experience an earlier onset of Fe~III to Fe~II recombination in the iron-group rich layers of ejecta, resulting in a more rapid evolution of the SN colors to the red. The faster \\Bband\\ decline rate of dimmer SNe~Ia thus reflects their faster ionization evolution. ", "introduction": "Although normal Type~Ia supernovae (SNe~Ia) are generally considered a homogeneous class, they nevertheless show substantial ($\\sim 1$~mag) variations in luminosity at peak. The use of SNe~Ia for cosmology measurements thus relies on empirical calibration techniques. Most common among these is the width-luminosity relation \\citep{Phillips_1999}. Bright SNe~Ia generally have broad \\Bband\\ light curves (LCs) which decline slowly after peak. Dimmer SNe~Ia have narrower, more quickly declining LCs. Understanding the physical origin of the width-luminosity relation (WLR) is a primary goal of the theory and modeling of SNe~Ia. Unfortunately, a full theoretical description of the SN~Ia explosion mechanism is still lacking. Normal SNe~Ia are widely believed to be the thermonuclear disruption of carbon-oxygen white dwarfs near the Chandrasekhar limit, however a number of uncertainties remain regarding the structure of the progenitor, the precise ignition conditions, and the physics of the turbulent nuclear combustion that unbinds the star. Fortunately, these uncertainties need not be fully resolved in order to study the WLR. In SN~Ia explosions, hydrodynamic and nuclear burning processes last only $\\sim 1$~minute. The subsequent luminosity is powered entirely by the decay of radioactive elements synthesized in the explosion, in particular \\Nifs\\ in the decay chain $\\Nifs \\rightarrow\\ \\Cofs \\rightarrow\\ \\Fefs$. The LCs of SNe~Ia are therefore fully determined by the composition and density structure of the material ejected in the explosion. Naturally, the mass of \\Nifs\\ produced (\\Mni) is the primary determinate of the peak brightness of the event. Observations indicate that for normal objects, \\Mni\\ spans the range $0.4-0.9$~\\Msun, with a typical value $\\Mni \\approx 0.6~\\Msun$. The challenge thus falls to radiative transfer theory to explain why SNe~Ia with larger \\Mni\\ also have broader \\Bband\\ LCs. Most previous transfer studies have emphasized diffusion arguments of one sort or another \\citep{Hoeflich_WLR, Hoeflich_99by, Pinto-Eastman_WLR, Mazzali_WLR}. In SNe~Ia, the ejecta remain optically thick for the first several months after explosion. The width of the bolometric LC is related to the time scale for photons to escape the ejecta by diffusion. In principle, the WLR could be explained if brighter SNe~Ia have higher effective opacities and hence a longer diffusion time. Different authors have invoked different physical arguments to motivate this sort of opacity dependence (see \\S\\ref{Sec:Others}). In this paper, we stress that the WLR does not in fact hinge on the bolometric diffusion time, but is instead principally a \\emph{broadband} phenomenon. In particular, the \\Bband\\ LC depends sensitively on the rate at which the SN colors evolve progressively redward following maximum light. Dimmer SNe~Ia exhibit a faster color evolution than brighter SNe~Ia; this turns out to be a primary reason for their relatively faster \\Bband\\ decline rates. A relevant theoretical explanation of the WLR should therefore focus on the effect \\Mni\\ has on the \\emph{spectroscopic/color evolution} of SNe~Ia, rather than on the overall diffusion time-scale. In what follows, we use realistic time-dependent multi-group radiative transfer calculations to explain the physical origin of the WLR in a set of Chandrasekhar-mass SN~Ia models. We demonstrate that the faster color evolution of dimmer models reflects their faster \\emph{ionization evolution}. Following maximum-light, the SN colors are increasingly affected by the development of numerous Fe~II/Co~II lines which blanket the \\Bband\\ and, at the same time, increase the emissivity at longer wavelengths. Because dimmer SNe~Ia are generally cooler, they experience an earlier onset of Fe~III to Fe~II recombination in the iron-group rich layers of ejecta, resulting in a more rapid evolution of the SN colors to the red. This explains their faster \\Bband\\ decline rate. ", "conclusions": "" }, "0609/astro-ph0609295_arXiv.txt": { "abstract": "During the past decade $\\eta$ Car has brightened markedly, possibly indicating a change of state. Here we summarize photometry gathered by the Hubble Space Telescope as part of the HST Treasury Project on this object. Our data include STIS/CCD acquisition images, ACS/HRC images in four filters, and synthetic photometry in flux-calibrated STIS spectra. The HST's spatial resolution allows us to examine the central star separate from the bright circumstellar ejecta. {\\it Its apparent brightness continued to increase briskly during 2002--06, especially after the mid-2003 spectroscopic event.\\/} If this trend continues, the central star will soon become brighter than its ejecta, quite different from the state that existed only a few years ago. One precedent may be the rapid change observed in 1938--1953. We conjecture that the star's mass-loss rate has been decreasing throughout the past century. ", "introduction": "% Eta Carinae's photometric record is unparalleled among well-studied objects, especially since it has been near or exceeded the classical Eddington limit during the past two or three centuries. From 1700 to 1800 it gradually brightened from 4th to 2nd magnitude, and then experienced its famous Great Eruption or ``supernova impostor event'' beginning about 1837. For twenty years it was one of the brightest stars in the sky, rapidly fluctuating between magnitudes 1.5 and 0.0, briefly attaining $V \\approx -1.0$. After 1858 it faded below 7th magnitude, presumably enshrouded in the nascent Homunculus nebula. Subsequent behavior, however, has been more complex than one might have expected. A mysterious secondary eruption occurred in 1887--1900; then the apparent brightness leveled off around $m_{pg} \\approx 8$ for about 40 years, followed by a rapid increase in 1938--53; after that it brightened at a fairly constant rate for another 40-year interval, and most recently the rate accelerated in the 1990's. Some, but not all, of the secular brightening can be attributed to decreasing obscuration as the Homunculus nebula expands. However, in truth this is more complex than it appears. The star's physical structure has been changing in a decidedly non-trivial way which is, at best, only dimly understood. For historical and observational details see \\citet{dh97,frew05,dvauc52, oconnell56,feinstein67,feinstein74,aavso,kdetal99a,kdetal99b, vangenderen99, sterken99,phot1}. Spectroscopic changes have occurred along with the brightness variations. The 5.5-year spectroscopic/photometric cycle \\citep{gaviola53,zanella84,whitelock94,whitelock04,damineli96,phot1} is not apparent in data obtained before the 1940's \\citep{feast01,rmh05}. Brief ``spectroscopic events'' marking the cycle are most likely mass-ejection or wind-disturbance episodes, probably regulated by a companion star \\citep{zanella84,kd99,smith03,heii}. At visual wavelengths, the associated ephemeral brightness changes represent mainly emission lines in the stellar wind, while the longer-term secular brightening trend involves the continuum \\citep{phot1,martinconf}. \\citet{rmh05}, \\citet{halpha05}, and \\citet{kd05} have speculated that the four obvious disruptions in the photometric record -- c.\\ 1843, 1893, 1948, and 2000 -- might indicate a quasi-periodicity of the order of 50 years.\\footnote{ So far as we know, this idea was first voiced by Humphreys at two meetings in 2002, but it did not appear in the published proceedings} In any case the star has not yet recovered from its Great Eruption seen 160 years ago. The Hubble Space Telescope (HST) Treasury Program for $\\eta$ Car was planned specifically to study the 2003.5 spectroscopic event. We employed the Space Telescope Imaging Spectrograph (STIS) and Advanced Camera for Surveys (ACS), following earlier STIS observations that began in 1998 \\citep{etatp04}. Fortuitously, the STIS data almost coincide with a rapid secular brightening which began shortly before 1998 (see Section 4 below). Those and the ACS images are of unique photometric value for at least two reasons: \\begin{enumerate} \\item{At visual wavelengths, normal ground-based observations have been dominated by the surrounding Homunculus ejecta-nebula, which, until recently, appeared much brighter than the central star and which has structure at all size scales from 0.1 to 8 arcseconds. {\\it So far, only the HST has provided well-defined measurements of just the central star.\\/}\\footnote{ At least this is true for visual and UV wavelengths. The near-infrared photometry reported by \\citet{whitelock94} and \\citet{whitelock04} may be strongly dominated by the central star. Those observations probably represent free-free emission in the wind at larger radii than the visual wavelength data. They show both the spectroscopic events and the brightening trend better than other ground-based measurements.} The Homunculus is primarily a reflection nebula, but the Homunculus/star brightness ratio has changed substantially. During 1998-99, for instance, the star nearly tripled in apparent brightness while ground-based observations showed only about a 0.3-magnitude brightening of Homunculus plus star \\citep{kdetal99b}. This rather mysterious development is known from HST/STIS and HST/ACS data. } \\item{Numerous strong emission lines perturb the results for standard photometric systems. H$\\alpha$ and H$\\beta$ emission, for example, have equivalent widths of about 800 and 180 \\mbox{\\AA} respectively in spectra of $\\eta$ Car. Broad-band $U$, $B$, $R$, and $I$ magnitudes, and most medium-band systems as well, are therefore poorly defined for this object. Photometry around 5500 \\mbox{\\AA}, e.g. broad-band $V$, is relatively free of strong emission lines, but transformations from instrumental magnitudes to a standard system require the other filters \\citep{kdetal99b,sterken01,vangenderen03}. This difficulty is somewhat lessened for HST observations restricted to the central star, whose spectrum has fewer emission lines than the bright ejecta; and some of the HST/ACS filters are fairly well-adapted to the case. At any rate the STIS and ACS data appear to be stable and internally consistent. Detector and filter systems used in most ground-based work, on the other hand, require fluctuating instrumental and atmospheric corrections, and do not give any major advantage for this object. } \\end{enumerate} In this paper we present the complete set of photometric data gathered for the $\\eta$ Car HST Treasury Project. The central star has brightened, especially in the UV, since the HST results described by \\citet{phot1}. We report three types of later measurements: \\begin{enumerate} \\item{The star's brightness in acquisition images made with STIS before that instrument's failure in early 2004. These images represent a broad, non-standard wavelength range from 6500 to 9000 {\\AA}.} \\item{The star's brightness in ACS/HRC images made in four filters (F220W, F250W, F330W, \\& F550M) from October 2002 to the present.} \\item{Synthetic photometry in flux-calibrated STIS spectra.} \\end{enumerate} After presenting the data below, we briefly discuss the observed trends and what they bode for the near future of $\\eta$ Car. Our principal reason for reporting these data now is that the Treasury Program observations have been completed; future HST/ACS observations are possible but not assured. These last Treasury Program observations are essential for demonstrating the secular trend in brightness (see Section \\ref{trend}), that these brightness changes suggest a fundamental change in the state of the star (see Section \\ref{discussion}) and what the near future may hold for $\\eta$ Carinae (\\ref{predictions}). ", "conclusions": "} % The observed brightening of $\\eta$ Car is not easy to explain. It cannot signify a major increase in the star's luminosity, because that would exceed the Eddington limit, producing a giant eruption. It cannot be a standard LBV-like eruption; in that case the energy distribution should have shifted to longer wavelengths, the Balmer emission lines should have decreased, and the spectrum should have begun to resemble an A- or F-type supergiant \\citep{lbv94}. In fact, qualitatively the star's spectrum has changed little in the past decade, and it has become bluer, not redder.\\footnote{ The change in color is modest, however, too small to confidently quote here. Dust near $\\eta$ Car has long been known to have an abnormally small reddening/extinction ratio, see \\citet{dh97}, \\citet{fos95}, and refs.\\ cited therein.} The most obvious remaining explanation involves a change in the circumstellar extinction, which, in turn, probably requires a subtle change in the stellar wind. Mere ``clearing of the dust'' -- i.e., motion of a localized concentration of dusty ejecta -- cannot occur fast enough \\citep{kdetal99b}. Therefore one must consider either destruction of dust grains, or a decrease in the formation of new dust, or both; and, if these account for the observations, why should they happen now? \\subsection{Dust Near the Star} The hypothetical decreasing extinction probably occurs within 2000 AU ($\\sim$ 1{\\arcsec}) of the star, and preferably closer, because: \\begin{enumerate} \\item{In various observations between 1980 and 1995, the star did not appear as bright as expected relative to the Weigelt blobs; the discrepancy was a factor of the order of 10, based on simple theoretical arguments \\citep{dh86,fos95}. Evidently, then, our line of sight to the star had substantially larger extinction at visual wavelengths, even though its projected separation from the blobs was less than 0.3{\\arcsec}. The required extra extinction was of the order of 3 magnitudes. Since then the star has brightened far more than the Weigelt objects have; therefore, if this involves localized extinction, its size scale must be a fraction of an arcsec, only a few hundred AU. } \\item{No known process seems likely to destroy dust more than 2000 AU from the star in a timescale of only a few years.} \\item{Ground-based photometry and HST images have shown only a modest, fraction-of-a-magnitude increase in the brightness of the large-scale Homunculus lobes during the past decade \\citep{laplata,phot1}. } \\end{enumerate} Dust grains should condense in $\\eta$ Car's wind at a distance of 200--600 AU, 2 to 10 years after the material has been ejected.\\footnote{ Here, lacking a specific dust-formation model for the unusual case of $\\eta$ Car, we suppose that appreciable grain condensation begins in the outward flowing material at the location where the equilibrium grain temperature is around 1000 K. This is a fairly conventional assumption and the precise choice of temperature has little effect on our reasoning. The quoted time-after-ejection assumes typical ejecta speeds of 200--700 km s$^{-1}$.} Since newly-formed dust moves outward in a timescale of several years, the circumstellar extinction seen at any time depends partly on the current dust formation rate. This, in turn, depends on local wind density, radiation density, etc., and newly formed hot grains ($T_d > 800$ K) are susceptible to destruction. The dust column density can thus be sensitive to small changes in the stellar parameters. Moreover, the wind is latitude-dependent and our line of sight is close to the critical latitude where wind parameters can vary rapidly \\citep{smith03}. All these factors appear suitable for the proposed explanation. On the other hand, near-infrared observations imply that extinction within $r < 2000$ AU has been quite small along most paths outward from the star. In Fig.\\ 3 of \\citet{cox95}, for instance, the 2--6 $\\mu$m flux indicates the high end of the dust temperature distribution. Modeling this in a conventional way, we find that less than 5\\% of the total luminosity was absorbed and re-emitted by inner dust with $T_d > 500$ K during the years 1973 to 1990 when those observations were made.\\footnote{ The measured flux was approximately a power law $f_{\\nu} \\sim {\\nu}^{-3.7}$ at wavelengths around 4 ${\\mu}$m. Assuming a typical emission efficiency dependence $Q_{\\nu} \\sim {\\nu}$, the observed spectral slope can be explained by a grain temperature distribution $dN/dT \\sim T^{-8.7}$. The result noted in the text is obtained by normalizing this to match the observed flux around 4 or 5 $\\mu$m and then integrating the total emitted flux at all wavelengths due to grains above 500 K.} Therefore, {\\it our line of sight must be abnormal in order to have a large amount of extinction near the star.\\/} In principle one might view this as an argument against our proposed scenario, but no plausible alternative has been suggested to explain the apparent faintness of the central star before 1999 and its ratio to the Weigelt blobs \\citep{speckle86,dh86,speckle88,fos95,kdetal99b}, The spatial distribution of dust is probably quite inhomogeneous near the star. The Homunculus lobes have a conspicuously ``granular'' appearance; the equatorial ejecta are clumpy, including the Weigelt knots; and stars near and above the Eddington limit tend to produce clumpy outflows \\citep{shaviv05}. Consequently the radiative transfer problem includes macroscopic effects which have not yet been modeled. If the grain albedo is sufficient, light may escape mainly by scattering through interstices between condensations. In that case, high-extinction lines of sight may be fairly common in the inner region even though most of the light escapes along other paths, not necessarily radial. Incidentally, the near-infrared photometric trends reported by \\citet{whitelock94,whitelock04} are not straightforward to interpret. The fairly-constant 3.5 $\\mu$m flux, for instance, represents a complicated mixture of dust formation parameters and does not necessarily indicate a constant amount of dust; see comments by \\citet{kdetal99b}. \\subsection{The Role of the Stellar Parameters \\label{stellarparams}} If the observed brightening represents a decrease in circumstellar extinction, the likeliest reason for this to occur is through some change in the star -- no one has yet proposed a suitable alternative. The most relevant stellar parameters are the radius, current luminosity, and surface rotation rate, which together determine the wind's velocity, density, and latitude structure. All of these may still be changing today, 160 years after the Great Eruption; thermal and rotational equilibrium in particular are likely to be poor assumptions for the star's internal structure \\citep{smith03,kd05}. As a working hypothesis to explain $\\eta$ Car's photometric and spectroscopic record in the past 100 years, let us tentatively suppose that {\\it the mass-loss rate is gradually decreasing,\\/} while the surface rotation rate may be increasing. Historical considerations include: \\begin{enumerate} \\item{High-excitation \\ion{He}{1} emission, now observed at most times, was consistently absent before 1920 \\citep{feast01} and probably before 1940 \\citep{rmh05}. If a hot companion star is present as most authors suppose, then the most obvious way to hide or suppress its helium ionization is to immerse the entire system in an extremely dense wind -- i.e., the primary star's mass-loss rate must have been larger then. This idea is far from straightforward \\citep{kd99}, but so far as we know it is the only qualitative explanation yet proposed. Informally, based on Zanstra-style arguments (i.e., assessing the volume emission measure $n_{He} n_e V$ needed to absorb all the photons above 25 eV), we estimate that a rate of the order of 10 times the present value, i.e. $\\sim 10^{-2}$ $M_\\odot$ y$^{-1}$, would have been required early in the twentieth century in order to suppress the helium recombination emission. } \\item{Twenty years ago the amount of fresh dust, indicated by the near-infrared flux, appeared consistent with a mass-loss rate somewhat above $10^{-3}$ $M_{\\odot}$ yr$^{-1}$ \\citep{iue86}. This absorbed only a small fraction of the luminosity (Section 5.1 above), but the substantially higher mass-loss rate suspected for earlier times would have produced enough hot inner dust to absorb a non-negligible fraction. } \\item{The brightness observed between 1900 and 1940 is rather mysterious. Judging from its mass and present-day optical thickness, around 1920 the Homunculus (then only half as large as it is today) should have had at least 5 magnitudes of visual-wavelength extinction; in a simple model the object should have been fainter than 10th magnitude instead of $m_{pg} \\approx 8$ as was observed. No doubt the inhomogeneities mentioned earlier played a role, but no model has been calculated. Moreover, why did the brightness remain fairly constant even though the Homunculus expanded by about 70\\% in 1900--1940? This interesting problem has received practically no theoretical attention.} \\item{\\ion{He}{1} emission first appeared, and $\\eta$ Car's brightness suddenly increased, between 1938 and 1953 as we mentioned in Section 1. This might conceivably be explained by a decrease in the wind density; but \\citet{kd05} and \\citet{halpha05} have conjectured that 1940--1950 may have been the time when rotation became fast enough to produce latitude structure in the wind. If so, a higher-excitation, lower-density zone then developed at low latitudes \\citep{smith03}. } \\end{enumerate} The above points inspire two hypotheses that may explain the rapid brightening trend shown in Fig.\\ \\ref{plot1}. First, if the mass-loss rate has been decreasing, this tends to reduce the column density of recently-formed dust along our line of sight. Meanwhile (or alternatively), perhaps the wind's latitude structure is continuing to evolve so that its dense zone is now moving out of the line of sight. HST data suggest that our line of sight has been fairly close to the critical boundary latitude separating the two phases \\citep{smith03}. A small increase in surface rotation rate, or some other parameter change, might conceivably move the dense zone to higher latitudes, decreasing the amount of dust that forms along our line of sight. This idea is appealing because it suggests a way in which the effective extinction may be very sensitive to the stellar parameters. This problem obviously requires detailed models far beyond the scope of this paper, combining the star's changing structure, its wind, dust formation, and possibly dust destruction. \\subsection{Concerning the 5.5-year Cycle} Figs.\\ \\ref{plot1} and \\ref{plot2} reveal no major surprises about the 2003.5 spectroscopic event, but several comments are worthwhile. First, the sharp drop in UV brightness (filters F220W and F250W) is qualitatively understood and does not involve circumstellar dust. During both the 1998 and the 2003 events, STIS data showed very heavy ultraviolet blanketing by ionized metal lines; indeed the star became quite dark at some wavelengths between 2000 and 3000 {\\AA} \\citep{ironcurtain}. We further note that just before the spectroscopic event, a slight increase occurred at wavelengths below 4000 {\\AA} (filters F220W, F250W, F330W), but not at visual and far-red wavelengths (F550M and F25ND3). Ground-based visual-wavelength and near-IR photometry showed a qualitatively similar effect \\citep{vangenderen03,whitelock04,laplata}. The brightening is particularly prominent in J, H, and K which are dominated by free-free emission \\citep{whitelock04}. The ACS F550M and STIS F25ND3 data primarily measure the continuum brightness, while the other HST filters are heavily influenced by strong emission or absorption lines. At about the same time \\ion{He}{1} emission in the central star also went through a similar increase in brightness \\citep{martinconf}. The minor pre-event brightening thus appears to represent an increase in some emission features implying a temporary increase in ionizing flux. The primary star may provide additional UV photons or the hypothetical hot companion star may excite the primary wind more than usual at that time (just before periastron), but no quantitative model has been attempted. Figs.\\ \\ref{plot1} and \\ref{plot2} contain interesting hints about the timescale for the star's post-event recovery. Four months after the 2003.5 event, for instance, the 2--10 keV X-ray flux had increased almost to a normal level \\citep{xray}. The HST/ACS F220W and F250W brightnesses, however, were still quite low at that time, and they required about eight months to recover. This timescale must be explained in any valid model for the spectroscopic events. \\citet{halpha05} noted serious differences between STIS spectra of the 1998.0 and 2003.5 events, and interpreted them as evidence for a rapid secular physical change in $\\eta$ Car. \\citet{damineli99} had earlier found that \\ion{He}{1} emission became progressively weaker after each of the last few spectroscopic events. These clues are obviously pertinent to our comments in Section \\ref{stellarparams} above. Fluctuations {\\it between\\/} spectroscopic events have received little attention in the past. For instance, Fig.\\ \\ref{plot1} shows a brief 0.2-magnitude brightening at 2001.3; measured by the STIS in both imaging and spectroscopic mode. It was correlated with the behavior of a strange unidentified emission line near 6307 {\\AA}, and with other subtle changes in the spectrum \\citep{uemit}. This is interesting because mid-cycle events have not been predicted in any of the competing scenarios for the 5.5-year cycle. Perhaps the effects seen in 2001 indicate the level of basic, LBV-like fluctuations in $\\eta$ Car. \\subsection{Eta Carinae in the Near Future \\label{predictions}} The appearance of $\\eta$ Car and the Homunculus nebula has changed dramatically. Twenty years ago the entire object could have been described as ``a bright, compact nebula having an indistinct eighth-magnitude central core;'' but a few years in the future, if recent trends continue, it will be seen instead as ``a fifth- or even fourth-magnitude star accompanied by some visible nebulosity.'' Meanwhile the color is gradually becoming bluer. This overall development has long been expected \\citep{kd87}, but now appears to be occurring 20 years ahead of schedule. If it signals an irregularity in the star's recovery from the Great Eruption, then this may be a highly unusual clue to the highly abnormal internal structure. Unsteady diffusion of either the thermal or the rotational parameters would be significant for stellar astrophysics in general. There are several practical implications for future observations of this object. Valid ground-based spectroscopy of the star (strictly speaking its wind) is becoming feasible for the first time, as its increased brightness overwhelms the emission-line contamination by inner ejecta. Unfortunately this implies that the inner ejecta -- particularly the mysterious Weigelt knots -- are becoming difficult to observe. In fact, since the HST/STIS is no longer available, they are now practically impossible to observe. When some new high-spatial-resolution spectrograph becomes available in the future, the inner ejecta will probably be much fainter than the star. The expected future of the larger-scale Homunculus nebula is also interesting. At present it is essentially a reflection nebula. However, based on the presence of high-excitation emission lines such as [\\ion{Ne}{3}] close to the star, the system almost certainly contains a source of hydrogen-ionizing photons with energies above 13.6, and helium-ionizing photons above 25 eV. (See, e.g., \\citet{zanella84}; most recent authors assume that this source is a hot companion star.) Eventually, when circumstellar extinction has decreased sufficiently due to expansion and other effects, the UV source will begin to photoionize the Homunculus. This is especially true if the primary stellar wind is weakening as we conjectured above. First the inner ``Little Homunculus'' will become a bright compact \\ion{H}{2} region, and then the bipolar Homunculus lobes. The time when that will occur is not obvious, but it may be within the next few decades if current trends continue." }, "0609/astro-ph0609776_arXiv.txt": { "abstract": "{} {\\hete\\ is the seventh known X-ray transient accreting millisecond pulsar and has been in outburst for more than one year. We compared the data on \\hete\\ with other similar objects and made an attempt at deriving constraints on the physical processes responsible for a spectral formation.} {The broad-band spectrum of the persistent emission in the 2--300 keV energy band and the timing properties were studied using simultaneous {{\\it INTEGRAL} } and publicly available {{\\it RXTE}} data obtained in October 2005. The properties of the X-ray bursts observed from \\hete\\ were also investigated.} {The spectrum is well described by a two-component model consisting of a blackbody-like soft X-ray emission at 0.8 keV temperature and a thermal Comptonized spectrum with electron temperature of 30 keV and Thomson optical depth $\\tau_{\\rm T} \\sim 2$ for the slab geometry. The source is detected by {{\\it INTEGRAL}} up to 200 keV at a luminosity of $5\\times10^{36}$ erg s$^{-1}$ (assuming a distance of 5 kpc) in the 0.1--200 keV energy band. We have also detected one type I X-ray burst which shows photospheric radius expansion. The burst occurred at an inferred persistent emission level of $\\sim$ 3--4\\% of the Eddington luminosity. Using data for all X-ray bursts observed to date from \\hete, the burst recurrence time is estimated to be about 2 days. No pulsations have been detected either in the {{\\it RXTE}} or in the {{\\it INTEGRAL}} data which puts interesting constraints on theories of magnetic field evolution in neutron star low-mass X-ray binaries. } {} ", "introduction": "\\label{sec:intro} The detection of X-ray millisecond pulsation in persistent emission from low-mass X-ray binaries (LMXBs) remained elusive for many years until the discovery of the first accreting millisecond pulsar (MSP) by \\citet{wvdk98}. Since that time, a total of seven accreting MSP transients have been detected. They are weakly magnetized ($\\sim 10^{8}-10^9$ G) neutron stars (NS) with spin frequencies in the 180--600 Hz range and orbital periods between 40 min and 5 hr \\citep[see reviews by][]{w05,p06}. Their companion stars have been found to be either highly evolved white or brown dwarfs. For the first time, the predicted decrease of the NS spin period during accretion was measured in the accreting MSP IGR~J00291+5934 \\citep{mfb05,burd06}. This provided a strong confirmation of the theory of `recycled' pulsars in which old neutron stars in LMXBs become millisecond radio pulsars through spin-up by transfer of angular momentum by the accreting material. MSP energy spectra can be well described by a two-component model consisting of a soft black body (or multi-color blackbody) and a hard power-law like tail. The soft thermal component could be associated with radiation from the accretion disc and/or the heated NS surface around the shock \\citep[see e.g.][]{gp05,p06}. The hard emission is likely to be produced by thermal Comptonization in the hot accretion shock on the NS surface \\citep{gdb02,pg03} with seed photons coming from the stellar surface. The observed hard spectra are similar to the spectra observed from atoll sources in their hard, low-luminosity state \\citep{b00}. \\hete\\ was discovered during a bright X-ray burst by the {\\it High Energy Transient Explorer 2} ({\\it HETE-2}) on 14 June 2005 \\citep{vanderspek05}. Followup observations with the {\\it Rossi X-ray Timing Explorer} ({\\it RXTE}) identified the source as the seventh X-ray accreting millisecond pulsar, with a pulse frequency of 377.3 Hz, an orbital period of 83 min, and most likely a 0.016--0.07 M$_{\\odot}$ brown dwarf companion \\citep{kaaret06}. The detected burst was consistent with a type I X-ray burst with photospheric radius expansion. Assuming that the bolometric burst peak luminosity during photospheric radius expansion saturated at the Eddington limit, \\citet{kawai05} estimated the distance to the source to be $\\sim5$ kpc assuming helium burst burning and canonical NS values. An optical counterpart candidate had an $R$-band magnitude of 18.02 and a broad HeII emission line spectrum \\citep[][]{fox05,steeghs05a}. A similar line was previously observed in IGR~J00291+5934 \\citep{rjs04,ffc04}. The optical counterpart of the X-ray source is located at coordinates $\\alpha_{\\rm J2000} = 19^{\\rm h}00^{\\rm m}08\\fs65$ and $\\delta_{\\rm J2000} = -24{\\degr}55\\arcmin13\\farcs7$ with an uncertainty of $0\\farcs2$. Near-infrared observations detected the optical candidate at a constant magnitude of J=17.6. No radio counterpart consistent with the \\hete\\ coordinates was detected by the VLA \\citep[][]{steeghs05b,rupen05}. One year after the discovery, \\hete\\ is still active (see Fig. \\ref{fig:asm}). Compared to other accreting MSPs with outburst periods of a few days to a month, \\hete\\ shows evidence of being a ``quasi-persistent'' X-ray source. This source also has other properties atypical for accreting MSPs. During the first $\\sim30$ days after the discovery, \\hete\\ showed significant flux emission variability in the fractional rms amplitude \\citep{galloway06b}. On July 8, 2005 (MJD 53559) during the flux brightening, the observed pulse frequency decreased by $\\Delta\\nu/\\nu\\sim6\\times10^{-7}$, and in the subsequent observations, the pulsations were suppressed \\citep{kaaret06}. In this paper we report the {\\it INTEGRAL} observations of \\hete\\ obtained simultaneously with {\\it RXTE}. We study the broad-band spectral and timing properties of the source. The X-ray burst properties are also investigated. \\begin{figure} \\centerline{\\epsfig{file=6457fig1.eps,width=8.0cm} } \\caption{{\\it RXTE}/ASM light curve for \\hete\\ averaged over 1-day intervals from May 5, 2005 (53500 MJD). The count rate has been converted into flux using 1 Crab Unit for 75 cts s$^{-1}$ \\citep{l96}. The arrows indicate the times of the detected X-ray bursts \\citep[][and this work]{vanderspek05,barbier05,galloway05}. } \\label{fig:asm} \\end{figure} ", "conclusions": "\\label{sec:conclusions} We have found that the source spectrum is similar to other accreting X-ray millisecond pulsars having a high plasma temperature around 30 keV and a Thomson optical depth $\\sim2$. This source differs from other MSP in requiring thermal soft X-ray emission with nearly double the temperature. From our spectral fits we infer that this emission is not likely be produced in a multi-temperature accretion disk but more likely arises from thermal emission at the NS surface. One might expect that the lack of pulsations could be due to a particularly high optical depth, but our spectral fits rule out this possibility. The reason for the lack of coherent pulsations in the persistent emission from LMXBs is still an open question. Different explanations have been put forward to explain this phenomenon, including models which invoke gravitational lensing, electron scattering, or weak surface magnetic fields due to magnetic screening \\citep[][and references therein]{w88,brainerd87,titarchuk02,cumming01}. The high accretion rate inferred for \\hete\\ relative to the other known MSP transients suggests that we may be observing the evolution of the NS's magnetic field due to magnetic screening in this source \\citep{cumming01}. The transition of \\hete\\ from an X-ray millisecond pulsar to a persistent LMXB could indicate that there is a population of suppressed X-ray millisecond pulsars among the non-pulsating LMXBs. Detailed observations of this source at the epoch of pulsation suppression can help to solving the long-standing issue of missing pulsations in persistent LMXB emission." }, "0609/astro-ph0609406_arXiv.txt": { "abstract": "This paper reports on a near-infrared survey of early-type galaxies designed to provide information on bar strengths, bulges, disks, and bar parameters in a statistically well-defined sample of S0-Sa galaxies. Early-type galaxies have the advantage that their bars are relatively free of the effects of dust, star formation, and spiral structure that complicate bar studies in later type galaxies. We describe the survey and present results on detailed analysis of the relative Fourier intensity amplitudes of bars in 26 early-type galaxies. We also evaluate the {\\it symmetry assumption} of these amplitudes with radius, used recently for bar-spiral separation in later-type galaxies. The results show a wide variety of radial Fourier profiles of bars, ranging from simple symmetric profiles that can be represented in terms of a single gaussian component, to both symmetric and asymmetric profiles that can be represented by two overlapping gaussian components. More complicated profiles than these are also found, often due to multiple bar-like features including extended ovals or lenses. Based on the gravitational bar torque indicator $Q_b$, double-gaussian bars are stronger on average than single-gaussian bars, at least for our small sample. We show that published numerical simulations where the bar transfers a large amount of angular momentum to the halo can account for many of the observed profiles. The range of possibilities encountered in models seems well-represented in the observed systems. ", "introduction": "S0 galaxies were introduced to the Hubble sequence by Hubble (1936) as a means of bridging the apparently catastrophic gap between E7 and Sa galaxies. After real examples were discovered (see Sandage 1961), the hallmark of the class became a disk shape (definitely ``later than'' E6 or E7) and an absence of spiral arms or star formation. SB0 galaxies were originally classified as SBa by Hubble (1926) even though they also lacked arms. This inconsistency was corrected in Sandage (1961). Barred S0 galaxies are extremely interesting because they help to take some of the mystery out of S0s in general: a bar is usually a disk feature that is closely related to spiral structure in many galaxies (e.g., Kormendy and Norman 1979). In addition, ring features are directly related to bars in spirals, and SB0s may show vestiges of similar rings. One could therefore ask whether bar properties in S0s might provide any clues as to how S0s and spirals might be related in an evolutionary sense. We are interested in the distribution of bar strengths in S0 galaxies, a property of very early-type galaxies that has not yet been tapped for what it might tell us about the evolutionary history of S0 galaxies. Early-type galaxies have a remarkable array of bar morphologies whose Fourier and other properties are worth characterizing in more detail. The advantage we have for examining these issues is that early-type galaxies show their bars largely unaffected by dust, star formation, and spiral structure. We can look for subtle structural differences between different bars and isolate possible different bar types. We are also interested in comparing early-type barred galaxies with the models of Athanassoula (2003, 2005), who demonstrated the importance of angular momentum transfer to the halo as a means of producing strong bars. The Fourier profile information we provide here is ideally suited to comparison with $n$-body models, and for evaluating the symmetry assumption used for bar-spiral separation (Buta, Block, \\& Knapen 2003). The Near-InfraRed S0 Survey (NIRS0S) is an attempt to obtain a statistically well-defined database of images of S0s from which the properties of S0 bars may be fairly compared to those of spirals. The Ohio State University Bright Galaxy Survey (OSUBGS, Eskridge et al. 2002) provides a valuable dataset for studying the properties of spiral galaxies, and has been fully tapped for bar strength studies by Block et al. (2002), Buta, Laurikainen, \\& Salo (2004=BLS04), Laurikainen, Salo, \\& Buta (2004a), Laurikainen et al. (2004b), and Buta et al. (2005). The questions we address with the NIRS0S sample are: (1) how strong do S0 bars get compared to spiral bars? (2) how does the distribution of bar strengths in S0 galaxies compare with that for spirals? (3) what characterizes the morphology of bars in S0 galaxies? and (4) what are the near-IR luminosity ratios and profile characteristics of bulges in S0 galaxies? In describing early-type disk galaxies, we will use specific terminology. An ``ansae\" bar is one showing bright enhancements near the ends (Sandage 1961). A ``regular\" bar does not show such enhancements. Ansae bars are preferentially found in early-type galaxies while regular bars are preferentially found in later types, although no statistical study has quantified the difference. A lens is a feature showing a shallow brightness gradient interior to a sharp edge (Kormendy 1979). An oval is a broad elongation in the light distribution; it differs from a conventional bar in lacking higher-order Fourier terms. If intrinsically elongated, a lens can also be an oval. Ovals are discussed further by Kormendy and Kennicutt (2004). In this paper, we focus on a set of 26 galaxies from the NIRS0S having ovals and/or bars, and investigate radial profiles of relative Fourier intensity amplitudes. We seek to examine the diversity in early-type bars according to the symmetry of such profiles. In section 2, we describe first the rationale for the survey and the sample selection criteria. In section 3, we describe the observations made with different instruments and detectors. The study of the relative Fourier intensity amplitudes is presented in section 5. A discussion of the results is presented in section 6. ", "conclusions": "Interpretation of the results in this paper requires a theory that accounts not only for bar strength and pattern speed, but also the varied shapes of bars. Studies by Athanassoula (2005 and references therein) show that angular momentum exchange is at the heart of all of these issues. Critical to the properties of bars is how much angular momentum is transferred to the halo. The effect depends on the density of matter in halo resonances and on how cold or hot the resonant material is. In order to absorb angular momentum a halo must be ``live\", as opposed to a rigid halo that cannot interact with other galaxy components. Cold, live halos can absorb so much angular momentum that a bar can grow very strong. Weaker bars form in warmer, smaller halos or rigid halos, while in hot disks, mainly ovals will form. The extreme effects of a live halo interaction has led to the possibility that bars might be found in systems lacking a background disk in the bar region. Gadotti and de Souza (2003) present two possible cases of this, although LSB05 present contrary evidence that does not support their claim. Athanassoula computes relative Fourier mass profiles for her models that can be compared to the $I_m/I_0$ profiles presented in this paper. Provided the mass-to-light ratio in the $K_s$ band is relatively constant, such a comparison should be fair. The problems of comparing $n$-body models with real bars are summarized by Athanassoula \\& Misiriotis (2002), who discuss the types of bars that develop in massive halo (MH), massive disk (MD), and intermediate models. Considering that the models are not of any specific galaxy, and are only a few of a large number of models actually computed, the relative Fourier mass profiles well resemble the observed profiles for some of our galaxies. The MH models show strong $m$=2, 4, 6, and 8 components while these are much weaker for the MD models, with $m$=6 and 8 being in the noise. The MH profiles most resemble those for NGC 1452, 2983, 4608, and 4643 in our sample. The model bars have a fairly sharp outer edge, as is seen in these galaxies. Athanassoula (2003) shows how the mass of the halo impacts the Fourier terms. More massive halos lead to stronger bars. Her most massive halo model (M$\\gamma$3) has Fourier profiles similar to those of NGC 1452 and 4643, two of the strongest bars in our sample. Model MH1, with a less massive halo, has profiles similar to NGC 936 and 2787, while model MH2, with the lowest halo mass in the three illustrated, shows mainly an oval bar and has profiles resembling those of NGC 1302 and 2781. The good agreement between the models and the observations is further shown by the simulations of Athanassoula, Lambert, \\& Dehnen (2005), who evaluate the effects of a central mass concentration (CMC) on the evolution of the bar models in the previous papers. They show that a CMC has an effect on the higher $m$ terms, weakening their importance and leading to more $m$=2-dominated bars. Before the introduction of a CMC in their massive halo (MH) models, the relative Fourier mass profiles (their Figure 4) strongly resemble those we observe for NGC 4643, which has the sharpest outer edge of all the bars in our sample. In this model, the mass ratios $A_4/A_2$ = 0.69, $A_6/A_0$ = 0.47, and $A_8/A_0$ = 0.36, are very comparable with the intensity ratios found using the first gaussian component of NGC 4643: $I_4/I_2$=0.66, $I_6/I_2$=0.45, and $I_8/I_2$=0.29, respectively (from Table 3). Comparable values are found for NGC 1452 and NGC 4608. In their MH models with a CMC, the profiles are more symmetric and the ratios above are considerably reduced. This is as observed for most of the other galaxies in our sample whose bars are weaker. Massive disk (MD) models in this paper also show similar effects, but the bars are weaker than for the MH models and the profile asymmetries are less important. Bureau \\& Athanassoula (2005) present three $n$-body models designed for deducing diagnostics of edge-on bars. These models are similar to those used by Athanassoula \\& Misiriotis (2002) and include a weak bar, an intermediate strength bar, and a strong bar, with each model bar surrounded by a strong (and largely circular) stellar inner ring. The Fourier decompositions of these barred galaxy models again resemble the galaxies described above. In the strong bar model, the outer end of the bar has a sharp edge, similar to what is seen in NGC 1452 and 4643 in our sample. The strongest circular inner rings in our sample are seen in NGC 1452 and 4608 (Figure~\\ref{images}). Weaker circular inner rings are seen in NGC 936, 1317, and 2787, but in several of our galaxies the rings are highly elongated. Early-type barred galaxies have a wide range in intrinsic inner ring shapes, as is typical of such features (Buta 1995). The relative Fourier intensity profiles of the bars in our sample could also include some of the effects of bar destruction. The bars in our sample that include overlapping extended ovals could be cases where the bar destruction was in progress at the time the galaxies became relatively deficient in gas. Bournaud \\& Combes (2002) suggested that bar destruction in the absence of external gas accretion can lead to an oval lens feature (see also Kormendy 1979). An extreme example of this in our sample could be NGC 2859, where the impact of the oval lens extends considerably beyond the ends of the bar. Our Fourier analysis is efficient for separating the lens from the bar in this case, based on the disappearance of the higher order Fourier modes in the lens. Such an extended oval is only weakly seen in NGC 1452, 4608, or 4643, as if these bars were at their peaks and had not begun to disintegrate. These same three galaxies also have little or no central oval or secondary bar." }, "0609/astro-ph0609630_arXiv.txt": { "abstract": "{The context of this paper is buoyant toroidal magnetic flux ropes, which is a part of flux tube dynamo theory and the framework of solar-like magnetic activity.} {The aim is to investigate how twisted magnetic flux ropes interact with a simple magnetized stellar model envelope---a magnetic ``convection zone''---especially to examine how the twisted magnetic field component of a flux rope interacts with a poloidal magnetic field in the convection zone.} {Both the flux ropes and the atmosphere are modelled as idealized 2.5-dimensional concepts using high resolution numerical magneto-hydrodynamic (MHD) simulations.} {It is illustrated that twisted toroidal magnetic flux ropes can interact with a poloidal magnetic field in the atmosphere to cause a change in both the buoyant rise dynamics and the flux rope's geometrical shape. The details of these changes depend primarily on the polarity and strength of the atmospheric field relative to the field strength of the flux rope. It is suggested that the effects could be verified observationally.} {} ", "introduction": "Buoyant magnetic flux tubes are an essential part of the framework of current theories of dynamo action in both the Sun and solar-like stars: it is widely believed that flux tubes are formed by some combination of rotational shear and turbulent convection near the bottom of the convection zones (CZ) of these stars. When sufficiently buoyant the magnetic flux rise in the form of tubular (toroidal) $\\Omega$-shaped loops of magnetic field lines. Rising under the influence of rotational forces, the loops finally emerge as slightly asymmetric and tilted bipolar magnetic regions at the surface, see e.g.\\ \\cite{Fan+ea94} and \\cite{Caligari+ea95}. During the last decade, it has been established that if they do exist, these flux tubes must in fact be flux {\\em ropes} of intertwined or twisted field lines; otherwise they would quickly be disrupted by a magnetic ``mushrooming'' instability, see e.g.\\ \\cite{Emonet+Moreno1998} and \\cite{Dorch+Nordlund1998}. It has been shown that the instability is inhibited if the degree of systematic twist is sufficiently high, corresponding to a critical value of the field line pitch angle $\\Psi_c$ approximately determined by equality between the energy density of the twisted field component and the ram pressure due to the rise of the flux rope e.g.\\ \\cite{Moreno+Emonet1996}. A fundamental ingredient in all theories of solar-like dynamos is a process where a relatively weak poloidal magnetic field is turned into a toroidal field by differential rotation: it is this toroidal field that eventually gives rise to the flux ropes, which at the end of their existence return a poloidal field component to the CZ, thereby completing the magnetic activity cycle. This dynamo process is gradual and different cycles overlap. Buoyant magnetic flux ropes have been extensively study through numerical simulations, cf.\\ the reviews in \\cite{Dorch2002} and \\cite{Fan2004}, and the most recent high resolution study in two dimensions by \\cite{Cheung+ea2006}. However, so far studies of the interaction of flux ropes with poloidal fields have solely dealt with flux ropes that emerge into the solar corona, e.g.\\ \\cite{Archontis+ea04}, i.e.\\ not with a magnetic CZ. The question addressed in this paper is how the polarity and strength of the field in a magnetized convection zone may affect the rise and evolution of twisted flux ropes. When the predominantly toroidal buoyant flux ropes rise through the CZ, they encounter a poloidal magnetic field that has a component perpendicular to the ropes' axes. However, since the flux ropes are twisted, the transversal field components may be either parallel or anti-parallel to the average magnetic field in the CZ, i.e.\\ the field in the two-dimensional plane perpendicular their toroidal axis. Several questions relevant to the theory of flux tube dynamos then emerge: e.g.\\ is the rise faster or slower and do more or less magnetic flux reach the surface in the presence of a magnetized convection zone; how does a magnetized convection zone affect the required amount of twist. In this paper I attempt to shed light on these questions by presenting results from idealized numerical 2.5-dimensional MHD simulations of the cross-sections of buoyant flux ropes that interact with a horizontal magnetic layer while they rise. ", "conclusions": "\\label{discussion.sec} It can be argued that two-dimensional models have several restrictions compared to the full three-dimensional case: e.g.\\ there is a tendency for enhanced flows due to the lack of the extra degrees of freedom associated with three dimensions. Therefore, one can consider the questions posed in this paper to be addressed here only to a certain order. On the one hand, the initial naive assumption that one could have---that the effect of the magnetized layer would be to destroy a flux rope with a twist anti-parallel to the polarity of the layer---is wrong. In fact, the opposite is true: even though reconnection within the magnetic layer reduces the twist of the flux ropes, it is easier for them to penetrate the poloidal layer, because the reconnection of field lines allows the ropes to ``carve'' their way through the layer. This is provided, however, that the flux ropes contain enough transversal flux. On the other hand, the simple estimate of $\\chi_{\\rm c} \\approx 5$ in Eq.\\ \\ref{chi_c.eq} seems to hold in the simulations: in the models with $\\chi < \\chi_{\\rm c}$, the ropes' buoyancy are strongly damped. However, while ropes with $\\chi > \\chi_{\\rm c}$ are allowed to rise, they may still be halted in the CZ, but the more correct critical limit will have to be a function also of twist, i.e.\\ $\\chi_{\\rm c} = \\chi(\\epsilon)$. It has been shown here that the flux ropes' dynamics depend on both the strength and sign of the poloidal field: for a relatively strong poloidal field the rise is slower and therefore the flux ropes reach a lower height in the same amount of time. In certain cases the rise can be completely halted. It turns out that anti-parallel twisted ropes reach higher and faster. When it comes to the flux ropes' topology, the geometrical shape of ropes also depends on the strength and sign of the poloidal field: the apex of anti-parallel twisted ropes are flatter and have steeper magnetic gradients in their axial field. Ideally this could be observed as a seemingly faster passage through horizontal layers (emergence) of the axial part of the flux rope, relative to cases with more moderate vertical gradients in the axial field: e.g.\\ \\cite{Archontis+ea04} have shown that in fact a steep axial gradient is required by flux emergence (their Eq. 10). As to the flux ropes' twisted field components, reconnection reduces the twist of anti-parallel ropes. If ropes on either sides of the equator are oppositely twisted, this could then in principle be falsified observationally by comparing observations of poloidal field strength in emerging northern and southern active regions. The most fundamental problems remaining are those of the origin of the twist, and the question of how it arises. One may speculate that twisted field lines could be generated in large-scale flux bundles located near the bottom of the convection zone connecting across the solar equator: such flux bundles would experience a rotating motion since their lower parts are located in a region rotating slower than their uppermost parts. This rotation would twist the magnetic field lines and this twist could be transmitted to the parts of the flux bundle at slightly higher latitudes, thereby possibly giving rise to a twisted toroidal flux system. The sign of a twisted component generated in this geometric manner would always be pointing in the same direction (northwards in the Sun). Hence, in consecutive 11-year half-cycles of the solar dynamo, the twist would alternate between being anti-parallel and parallel to the poloidal field (from which the corresponding toroidal field was generated in the respective cycle). Such an alternation would result a 22-year cycle in the amount of twist in the emerging parts of the toroidal flux ropes, e.g.\\ in the helicity of bipolar magnetic regions. Fully three-dimensional MHD simulations extending the present study are under way, using the numerical scheme of \\cite{Archontis+ea04}." }, "0609/hep-ph0609305_arXiv.txt": { "abstract": "The LVD detector, located in the INFN Gran Sasso National Laboratory (Italy), studies supernova neutrinos through the interactions with protons and carbon nuclei in the liquid scintillator and interactions with the iron nuclei of the support structure. We investigate the effect of neutrino oscillations in the signal expected in the LVD detector. The MSW effect has been studied in detail for neutrinos travelling through the collapsing star and the Earth. We show that the expected number of events and their energy spectrum are sensitive to the oscillation parameters, in particular to the mass hierarchy and the value of $\\theta_{13}$, presently unknown. Finally we discuss the astrophysical uncertainties, showing their importance and comparing it with the effect of neutrino oscillations on the expected signal. ", "introduction": "There are many experimental works suggesting neutrino conversion among flavors in the recent few years, through the study of atmospheric \\cite{skatm}, solar \\cite{Cl} \\cite{Ga} \\cite{Sage} \\cite{Kam} \\cite{sksol} \\cite{sno}, reactor \\cite{KamLAND} and accelerator \\cite{K2K} neutrinos. The interpretation of all these phenomena in terms of neutrino oscillations is rather robust, because it is able to include all the experimental data (except the ``not yet confirmed'' LSND \\cite{lsnd} signal), even if the expected oscillatory behavior, in terms of the observable $L/E$, has not been yet experimentally observed (preliminary results that show a low significance hint for a oscillatory behavior have been found by a re--analysis of the SK data \\cite{skatm}). An interesting fact is that the inclusion of the MSW effect \\cite{msw} permits a consistent interpretation of KamLAND results and the `high energy' solar neutrino data \\cite{Kam,sksol,sno}. In the standard three flavor scenario, six parameters must be determined by oscillation experiments: 3 mixing angles ($\\theta_{{\\rm sol}}$, $\\theta_{13}$, $\\theta_{{\\rm atm}}$ ), 2 squared mass differences ($\\Delta m^2_{{\\rm sol}}$ and $\\Delta m^2_{{\\rm atm}}$) and 1 CP-violation phase $\\delta$. A recent analysis of all the available experimental data \\cite{VissStrum05} constrains the ``atmospheric'' and ``solar'' parameters to be in the following $99\\% ~C.L.$ ranges (compare also with the results in \\cite{FLglobal}): \\begin{table}[h] $$\\begin{array}{lrlc} \\hbox{Oscillation parameter}&\\multicolumn{2}{c}{\\hbox{central value}} &\\hbox{$99\\%$ C.L. range}\\\\ \\hline \\hbox{solar mass splitting} & \\Delta m^2_{{\\rm sol}} ~= & (8.0\\pm 0.3)\\,10^{-5}~\\eV^2 & (7.2\\div 8.9)\\,10^{-5}\\eV^2 \\\\ \\hbox{atm. mass splitting~~~~} & |\\Delta m^2_{{\\rm atm}}| ~= & (2.5\\pm 0.3) \\,10^{-3}~\\eV^2~~~ & (1.7\\div 3.3) \\,10^{-3}\\eV^2\\\\ \\hbox{solar mixing angle} & \\tan^2 \\theta_{{\\rm sol}} ~=& 0.45\\pm0.05 & 30^\\circ < \\theta_{{\\rm sol}}<38^\\circ \\\\ \\hbox{atm. mixing angle} & \\sin^2 2\\theta_{{\\rm atm}} ~= & 1.02\\pm 0.04 &36^\\circ <\\theta_{{\\rm atm}}< 54^\\circ\\\\ \\end{array}$$ \\label{tab1} \\end{table} However the other parameters are not completely determined: the $\\theta_{13}$ mixing angle is only upper limited, mainly by the Chooz experiment data \\cite{Chooz} ($\\sin^2 \\theta_{13} < 3.~ 10^{-2}$ at the $99 \\%~ C.L.$), the sign of $\\Delta m^2_{{\\rm atm}}$ (that fixes the so--called mass hierarchy) is completely unknown, as well as the CP--violation phase $\\delta$. Because of the wide range of matter density in the stellar envelope, a supernova explosion represents a unique scenario for further study of the neutrino oscillation mixing matrix. Indeed neutrinos can cross two resonance density layers and therefore the resulting possible mixing scenarios are different from the solar ones. The emerging neutrino spectra are sensitive to the sign of $\\Delta m^2_{{\\rm atm}}$ and to the value of $\\theta_{13}$. Before proceeding, it is important to recall that, at present, there is not a unique theory of supernova explosions. Till now, numerical investigations of the ``standard model'' based on a delayed scenario of the explosion \\cite{Olga0} failed to reproduce the explosion. On top of that, other models are being studied where rotation \\cite{Olga1} or magnetic field \\cite{Olga2} play an essential role. In the following, we will use a simple description of the neutrino flux that does not contradict the SN1987A events seen by Kamiokande-II \\cite{Olga3}, IMB \\cite{Olga4} and Baksan \\cite{Olga5}, see e.g. \\cite{Olga6} for a discussion, although it is not able to take into account the events seen in Mont-Blanc observatory \\cite{Olga7,Oscar0}. This ``standard'' description, however, corresponds to the expected neutrino emission in the delayed scenario and in the last phase of the collapse with rotation \\cite{Olga8}. For this reason, we take it as a useful starting point for the investigation of the impact of oscillations in the neutrino signal from a supernova. The main aim of this paper is to show how neutrino oscillations affect the signal detected by the LVD observatory in the INFN Gran Sasso National Laboratory, Italy. We also evaluate the impact on the signal of the astrophysical parameters of the supernova explosion mechanism, such as the total energy emitted in neutrinos, the star distance, the neutrino--sphere temperatures and the partition of the energy among the neutrino flavors. In section \\ref{se:sn} we describe the characteristics of the neutrino fluxes emitted during a gravitational core collapse. In section \\ref{se:osc} the neutrino oscillation mechanism is shown, in particular the peculiarities of the MSW effect in the supernova matter and in the Earth. The LVD detector and the relevant neutrino interactions both in the liquid scintillator and in the iron support structure are described in section \\ref{se:lvd}. The impact of neutrino oscillations in the signal expected in the LVD detector is presented in section \\ref{se:res} while the uncertainties in the astrophysical parameters and their effect on the results are discussed in section \\ref{se:ap}. Finally, the conclusions are drawn in section \\ref{se:sum}. Two appendices complete this work, describing in more detail the MSW effect calculation in the Earth (A) and the neutrino interaction with the iron of the LVD support structure (B). Preliminary results have been presented previously in \\cite{taup01}, \\cite{icrc03} and \\cite{Giacobbe}. ", "conclusions": "\\label{se:sum} The main aim of this paper was to show how neutrino oscillations affect the signal expected in the LVD detector at the occurence of the next galactic supernova. The LVD detector has been described in its main components. It is able to detect neutrinos of all flavors, by studying them in the various CC and NC channels. All the neutrino interactions that occur in the liquid scintillator as well as in the iron support structure have been studied in detail taking into account the neutrino energy threshold, cross section and detection efficiency. We assumed a galactic supernova explosion at a typical distance of $D = 10$~kpc, parametrized with a pure Fermi--Dirac energy spectrum ($\\eta = 0$) with a total energy $E_b = 3 \\cdot 10^{53}$ erg and perfect energy equipartition $f_{\\nu_e}=f_{\\bar\\nu_e}=f_{\\nu_x}=1/6$; we fixed $T_{\\nu_x} / T_{\\bar{\\nu}_e} = 1.5$, $T_{\\nu_e} / T_{\\bar{\\nu}_e} = 0.8$ and $T_{\\bar{\\nu}_e} =5~{\\rm MeV}$. We considered neutrino oscillations in the standard three-flavor scenario. The MSW effect has been studied in detail for neutrinos travelling through the supernova matter. We also considered the distortion in the expected neutrino spectra induced by a possible path inside the Earth before their detection. For the chosen supernova parameters, it results that the expected number of events and their energy spectrum depend on the unknown oscillation parameters: the mass hierarchy and the value of $\\theta_{13}$. In particular, the inverse beta decay interactions ($\\anue p, e^+ n$) are highly sensitive to the mass hierarchy: for adiabatic transition, the number of events increases of $\\sim 25 \\%$ in the IH case, with respect to the NH one, since the detected $\\anue$ completely come from the higher energy $\\nux$. The mean energy of the detected positrons is correspondingly increased. The total number of $(\\nu_e + \\bar\\nu_e)$ CC interaction with $^{12}$C nuclei is highly increased taking into account neutrino oscillations, because of their high energy threshold. For adiabatic transition the expected number of events is higher than the non adiabatic one, because at least one specie (between $\\nu_e$ or $\\bar\\nu_e$) comes significantly from the original and higher--energy $\\nux$ in the star. However, if it is not possible to discriminate between $\\nu_e$ and $\\bar\\nu_e$, the normal and inverted hierarchy cases present similar results. Indeed, in the NH (IH) case, the increase in $\\nue$ ($\\anue$) is compensated by a decrease in $\\anue$ ($\\nue$). The neutrino interactions with the iron of the support structure, which are studied in detail in this work, are also incread by the oscillations. The efficiency for the detection of the produced charged leptons and gammas in the active part of the detector has been calculated with a full simulation of the apparatus. The contribution of $(\\nu_e+\\bar \\nu_e)$ {\\rm Fe} interactions can be as high as $17\\%$ of the total number of events (in the adiabatic NH case) and they contribute mostly to the high energy part of the spectrum. With respect to the previous detection channels, the number of NC interactions with $^{12}$C nuclei does not depend on oscillations. In principle they could be used as a reference to identify the $\\nux$--sphere temperature. However, this is partly limited by the uncertainties in the other astrophysical parameters. We completed the calculations taking into account the effect of the passage of neutrinos through the Earth before their detection. This induces a characteristic modulation in the energy spectrum; however, given the expected number of events and the assumed oscillation parameters, the effect is quite weak. In conclusion, for the choice of the astrophysical parameters adopted in this work, the expected signal of neutrinos in the LVD detector from a supernova core collapse greatly benefits of the neutrino oscillation mechanism, practically in all the possible detection channels, especially if the transition is adiabatic and the hierarchy inverted (since in LVD the most relevant signal is given by $\\anue$). However, being aware of the fact that the astrophysical parameters of the supernova mechanism are up to now not well defined, we performed the same calculations using different values of them. The resulting differences are in fact important; they are mainly due to the poor theoretical knowledge of the physics of the gravitational collapse. This will be hopefully improved after the occurence and detection of the next galactic supernova, to which the LVD experiment can give a significant contribution, thanks to its cabability to observe and measure neutrino events of several types. \\appendix \\newpage {\\bf \\Large Appendices}" }, "0609/nucl-th0609074_arXiv.txt": { "abstract": "A method for calculation of Gamow-Teller transition rates is developed by using the concept of the Projected Shell Model (PSM). The shell model basis is constructed by superimposing angular-momentum-projected multi-quasiparticle configurations, and nuclear wave functions are obtained by digonalizing the two-body interactions in these projected states. Calculation of transition matrix elements in the PSM framework is discussed in detail, and the effects caused by the Gamow-Teller residual forces and by configuration-mixing are studied. With this method, it may become possible to perform a state-by-state calculation for $\\beta$-decay and electron-capture rates in heavy, deformed nuclei at finite temperatures. Our first example indicates that, while experimentally known Gamow-Teller transition rates from the ground state of the parent nucleus are reproduced, stronger transitions from some low-lying excited states are predicted to occur, which may considerably enhance the total decay rates once these nuclei are exposed to hot stellar environments. ", "introduction": "The knowledge on weak interaction processes is one of the most important ingredients for resolving astrophysical problems. The first systematical work on stellar weak-interaction rates was performed by Fuller, Fowler, and Newman \\cite{FFN1,FFN2,FFN3,FFN4}, who recognized the decisive role played by the Gamow-Teller (GT) transitions. Due to their pioneering work, the study of stellar weak interaction rates requested by astrophysics becomes essentially a nuclear structure problem, in which the actual decay rates are determined by the microscopic inside of nuclear many-body systems. It has been suggested that the nuclear shell model, i.e. a full diagonalization of an effective Hamiltonian in a chosen model space, is the most preferable method for GT transition calculations. This was noticed early by Aufderheide {\\it et al.} \\cite{ABRM93}, and has recently been emphasized by Langanke and Mart\\'inez-Pinedo \\cite{LM03}. For a theoretical model employed in GT transition calculations, it is generally required that the model can reproduce a wide range of structure properties of relevant nuclei. It has been shown that the state-of-the-art shell-model diagonalization method is indeed capable of performing such calculations. For example, Wildenthal and Brown \\cite{Wild84,BW85} obtained nuclear wave functions in the full $sd$-shell model space, which were successfully applied to calculation of GT rates in the $sd$ shell nuclei \\cite{Oda94}. By using the method developed by the Strasbourg-Madrid group \\cite{SM94}, Langanke and Mart\\'inez-Pinedo \\cite{LM01} made the shell-model GT rates available also for the $pf$ shell nuclei. Still, these sophisticated calculations are tractable only for nuclei up to the mass-60 region, and cannot be applied to heavier nuclei which play important roles in the nuclear processes in massive stars. In the long history of the nuclear shell-model development, tremendous effort has been devoted to extending the shell-model capacity from its traditional territory to heavier shells. Despite the great progress made in recent years, it seems impossible to treat an arbitrarily large nuclear system in a spherical shell model framework due to the unavoidable problem of dimension explosion. One is thus compelled to seek judicious schemes to deal with large nuclear systems. The central issue has been the shell-model truncation. There are many different ways of truncating a shell-model space. While in principle, it does not matter how to prepare a model basis, it is crucial in practice to use the most efficient one. In this regard, we recognize the fact that except for a few lying in the vicinity of shell closures, most nuclei in the nuclear chart are deformed. This naturally suggests for a shell model calculation to use a {\\it deformed} basis to incorporate the physics in large systems. That is the philosophy that the Projected Shell Model (PSM) \\cite{PSM} and the several important generalizations \\cite{Sheikh99,Chen01,Sun02,Sun03,Gao06} are based on. The present article reports on the new development for calculation of GT transition rates in the PSM framework. Before a detailed description of the work, we mention a few attractive features in our approach, which may be relevant for future astrophysical applications. \\begin{itemize} \\item The PSM utilizes single particle bases generated by deformed mean-field models yet carries out a shell-model diagonalization like the conventional shell model. Conceptually, the PSM bridges two important nuclear structure methods: the deformed mean-field approach and the conventional shell model, and takes the advantages of both. On the one hand, as a shell model, the PSM can be applied to any heavy, deformed nuclei without a size limitation. On the other hand, unlike the mean-field models or models with an average nature, the PSM wave functions contain correlations beyond mean-field and the states are written in the laboratory frame having definite quantum numbers such as angular-momentum and parity. These are needed properties when the wave functions are employed in transition calculations. \\item Because of the way the PSM constructs its basis, the dimension of the model space is small (usually in the range of $10^2 - 10^4$). With this size of basis, a state-by-state evaluation of GT transition rates is computationally feasible. This feature is important because in stellar environments with finite temperatures, the usual situation is that the thermal population of excited states in a parent nucleus sets up connections to many states in a daughter by the GT operator. However, our current knowledge on GT transitions from excited nuclear states is very poor, and in many cases, it must rely on theoretical calculations. \\item The calculation of forbidden transitions involves nuclear transitions between different harmonic oscillator shells and thus requires multi-shell model spaces. Such a calculation is not feasible for most of the conventional shell models working in one-major shell bases. The PSM is a multi-shell shell model. This feature is desired particularly when forbidden transitions are dominated. \\item Isomeric states belong to a special group of nuclear states because of their long half-lives. The existence of isomeric states in nuclei could alter significantly the elemental abundances produced in nucleosynthesis. There are cases in which an isomer of sufficiently long lifetime can change the paths of reactions taking place and lead to a different set of elemental abundances \\cite{Sun05a,Sun05b}. It has been shown that the PSM is indeed capable of describing the detailed structure of isomeric states \\cite{Sun04}. \\end{itemize} The paper is organized as follows. In Sec. II, we briefly introduce the PSM concept and describe how shell model diagonalization is carried out in the angular-momentum-projected bases. Technical details of calculation of transition matrix elements in the projected bases are given in Sec. III. Our first example of a GT transition calculation is illustrated in Sec. IV, where we first validate the model by comparing the calculated structure properties (energy levels and electromagnetic transitions) with experiment. The obtained GT transition results are then discussed and the effects caused by the Gamow-Teller residual forces and by configuration mixing are studied. Finally, the work is summarized and an outlook on future applications is given in Sec. V. ", "conclusions": "In this article, we have presented the new development of a shell model method for calculation of Gamow-Teller transition rates. The method is based on the Projected Shell Model. Different from the conventional shell model, which builds its states in a spherical basis, the PSM constructs its states in a deformed basis in which important nuclear correlations are taken into account very efficiently. Therefore, it is possible for a shell model diagonalization in the PSM to be carried out in a manageable space for medium to heavy, and even for super-heavy nuclei. We have shown how the GT transition matrix elements are calculated in the PSM framework. A computer code has been developed and been tested. One nontrivial test has been done through the Ikeda sum-rule. We have obtained a reasonable distribution of the B(GT) strength with a fulfillment of the sum-rule. We have presented the first example from the rare earth region. In the calculation of $^{164}$Ho $\\rightarrow$ $^{164}$Dy electron-capture process, we have predicted the GT transition rates for the excited states. Such rates should be included as part of the total rate when these states are thermally populated in hot stellar environments. In the calculation discussed so far, we have considered only allowed $\\beta$-decay for the low-lying states. Calculation of allowed $\\beta$-decay for states with high excitation and the forbidden transitions is possible in the PSM framework. Study of GT giant resonance is also under consideration. The method described in the present article can be applied to various fields such as nuclear astrophysics and fundamental physics, where weak interaction processes take place in nuclear systems \\cite{ALW05}. In particular, one may find interesting applications to cases where a laboratory measurement for certain weak interaction rates is difficult and where the conventional shell model calculations are not feasible. Potential applications in nuclear astrophysics are calculations of $\\beta$-decay rates for the r-process \\cite{r-process} and the rp-process \\cite{rp-process} nucleosynthesis, and electron-capture rates for the core collapse supernova modelling \\cite{LM03,SN}. In the double-$\\beta$ decay theory, theoretical calculations for the nuclear matrix elements are needed, for which one has relied on the Quasiparticle Random Phase Approximation \\cite{Civitarese98,double-beta}, particularly when heavy nuclei are involved. We expect that the method presented here can make important contributions to all these studies." }, "0609/astro-ph0609156_arXiv.txt": { "abstract": "{} {We use 9.5-yr of BiSON Sun-as-a-star data to search for dependence of solar-cycle parameter changes on the angular degree, $\\ell$, of the data. The nature of the Sun-as-a-star observations is such that for changes measured at fixed frequency, or for changes averaged across the same range in frequency, any $\\ell$ dependence present carries information on the latitudinal distribution of the agent (i.e., the activity) responsible for those changes.} {We split the 9.5-yr timeseries into contiguous 108-d pieces, and determine mean changes in the damping of, power in, and energy supplied to the modes through the solar cycle. We also apply a careful correction to account for the deleterious effects of the ground-based BiSON window function on the results.} {From our full analysis we obtain a marginally significant result for the damping parameter, where the mean change is found to be weakest at $\\ell=0$. The other parameters show hints of some dependence in $\\ell$.} {Our main conclusion is that the mean fractional solar-cycle change in the $\\ell=0$ damping rates is approximately 50\\,\\% smaller than was previously assumed. It had been common practice to use an average over all low-$\\ell$ modes; our downward revision of the radial-mode value has implications for comparisons with models of the global solar cycle changes, which are usually based on a spherically symmetric geometry.} ", "introduction": "\\label{sec:intro} The fact that damping rates and powers of the global p modes change through the solar cycle is now well established \\citep[e.g., ][]{chaplin00,komm00,salabert03,chano03,chano04,salabert06}. At least where the main part of the p-mode spectrum is concerned damping gets heavier, and observed power gets weaker, as the level of solar activity increases. Information on the damping and power parameters comes straightforwardly from the observations. The damping rates are assumed to be linearly related to the linewidths of the resonant peaks in the frequency power spectrum; while the powers are proportional to the product of the peak widths and heights (the latter more formally termed the maximum power spectral densities). \\citet{komm02} took advantage of the large number of components available to analyze in the medium-degree range (from $\\ell=40$ to 80) the latitudinal dependence of the width and height changes. They found that the changes were concentrated in latitudes occupied by the active regions. The damping rates and powers, like the mode frequencies, therefore seemed to be responding in some fashion to changes wrought on the active regions by the changing magnetic fields. Most of the low-$\\ell$ results have come from the Sun-as-a-star data (e.g., from BiSON, GOLF, IRIS and VIRGO/SPM; or from GONG and MDI data combined into a Sun-as-a-star-like proxy). These data show only the even $\\ell+m$ components; the odd components are so weak as to be unobservable. Since it is the outer, sectoral components (with $\\ell=|m|$) that appear most prominently, the Sun-as-a-star mode parameters estimated by the usual analysis methods are dominated by, and therefore \\emph{close to}, the sectoral values. This characteristic gives a noticeable change in the latitudinal sensitivity from $\\ell=0$ to $\\ell=1$, 2 or 3. The spatial sensitivity of Sun-as-a-star data at the latter three values of $\\ell$ is weighted toward the lower latitudes, where the active regions reside. The Sun-as-a-star parameters at these $\\ell$ are therefore more sensitive to the solar cycle than are the $\\ell=0$ data. Evidence for $\\ell$ dependence in the shifts of the Sun-as-a-star mode parameters therefore carries information on the latitudinal distribution of the agent responsible for those shifts. (Changes in inertia, at fixed frequency, between these $\\ell$ are very small indeed.) It is much more difficult to uncover spatial dependence of the mode parameter changes in the low- than in the medium-$\\ell$ data because there are far fewer components to analyze, and uncertainties on the results are commensurately larger. When changes to damping and power were first uncovered in the Sun-as-a-star data, results were averaged over $\\ell$ to reduce errors. Given the modest precision in the results, it was assumed the $\\ell$-averaged values also provided a working proxy of the radial-mode shifts. And so these averages were used as meaningful comparisons for models of global changes in damping \\citep[see, e.g.,][]{houdek01}, models set up in spherical geometry, i.e., pertinent only to the $\\ell=0$ case. With better analysis, and more data, it has become possible to give results for each $\\ell$, and to therefore test whether the $\\ell=0$ shifts really are weaker than an average across $\\ell=0$ to 2 or 3. We made a first cut at such an analysis in \\citet{chaplin03a}. Our results, from BiSON Sun-as-a-star data, suggested very strongly that the $\\ell=0$ linewidth changes were indeed significantly weaker than at $\\ell=1$ and 2. The clear implication was that results of the theoretical models would now need to be compared to this new, smaller shift. In the analysis of the BiSON data, we had to allow for the corrupting influence on the results of the ground-based BiSON window function. We did so by applying corrections that were designed originally for a different study, using data from a different period. In this paper we revisit our analysis. We design and implement a correction procedure for the 9.5-yr BiSON dataset in question. We find that implementation of this internally consistent, and more accurate, correction gives little change to the results, and reinforces our earlier conclusion. In summary, the mean fractional solar-cycle change in the $\\ell=0$ damping rates is approximately 50\\,\\% smaller than was previously assumed. ", "conclusions": "We have analyzed some 9.5-yr of BiSON Sun-as-a-star data -- collected in solar cycles 22 and 23 -- to search for dependence of the mode excitation and damping parameter changes on the angular degree, $\\ell$, of the data. The nature of the Sun-as-a-star observations is such that for changes measured at fixed frequency, or for changes averaged across the same range in frequency, any $\\ell$ dependence present carries information on the latitudinal distribution of the agent (i.e., the activity) responsible for those changes. We split the 9.5-yr timeseries into contiguous 108-d pieces, and determined mean changes in the damping of, power in, and energy supplied to, the modes through the solar cycle. We also applied a careful correction to account for the deleterious effects of the ground-based BiSON window function on the results. This correction was calibrated by, and then fully tested on, artificial seismic data generated by the solarFLAG mode simulation code. From our full analysis we obtained a marginally significant result for the damping parameter, where the mean change was found to be weakest at $\\ell=0$, and higher in data on the other $\\ell$. The result implies the damping is strongest in the active regions, confirming results on the more numerous higher-$\\ell$ data \\citep{komm02}. The other excitation and damping parameters we investigated showed hints of some dependence in $\\ell$, but nothing that could judged as statistically significant. Our main conclusion is that the mean fractional solar-cycle change in the $\\ell=0$ damping rates is approximately 50\\,\\% smaller than was previously assumed. It had been common practice to use an average over all low-$\\ell$ modes (where mean solar-cycle shift values have averaged about 18\\,\\% \\citep{chaplin00,salabert03,chano03,chano04}. Our downward revision of the radial-mode value has implications for comparisons with models of the global solar cycle changes, which are usually based on a spherically symmetric geometry." }, "0609/astro-ph0609683_arXiv.txt": { "abstract": "We explore the possibility of having a composite (self-conserved) dark energy (DE) whose dynamics is controlled by the quantum running of the cosmological parameters. We find that within this scenario it is feasible to find an explanation for the cosmological coincidence problem and at the same time a good qualitative description of the present data. ", "introduction": "Independent data from different observations\\,\\cite{Supernovae, WMAP3Y,LSS} provide strong support for the existence of DE and seem to agree that it presently constitutes $\\sim70\\%$ of the total energy density. Nevertheless, the nature of DE remains unclear. If we identify it with a cosmological constant (CC) arising from the quantum field theory (QFT) vacuum energy, as done in the $\\CC$CDM model\\,\\cite{Peebles84}, we are led to a value many orders of magnitude greater than the measured one, what has been called the CC problem \\cite{weinRMP,Copeland06}. This problem could be alleviated by means of a dynamical DE. This possibility is supported by some recent studies\\,\\cite{Alam,Jassal1} and has been exploited profusely in various forms \\cite{Copeland06}. Among them the scalar fields are the most paradigmatic scenario. It must be stressed though that the presence of a scalar field is not essential for a model to be described in terms of an effective EOS, $\\pD=\\we\\,\\rD$ (for instance, this has been proven for any model with variable cosmological parameters in \\cite{SS12}). We present here a model in which the DE, in addition of being dynamical, is allowed to be composite. This model may offer an explanation to the ``cosmological coincidence problem''\\,\\cite{weinRMP} -i.e. to the fact that the DE and matter densities happen to be similar precisely at the present epoch- by keeping the ratio between these two densities bounded and of order 1 during the entire Universe existence. At the same time, the effective EOS of our model can match the available data. This feature is exemplified through specific numerical examples. ", "conclusions": "We have shown that the $\\CC$XCDM model can be in good agreement with present data and provide a solution to the cosmological coincidence problem as well as a clear signature. In our opinion the next generation of high precision cosmology experiments (DES, SNAP, PLANCK)\\,\\cite{SNAP} should consider the possibility of a composite DE with dynamics controlled by the running of the cosmological parameters. \\begin{theacknowledgments} This work has been supported in part by Ministerio de Eduaci\\'on y Ciencia of Spain (MEC) and FEDER under project 2004-04582-C02-01, and also by DURSI under 2005SGR00564. JG was also supported by MEC under BES-2005-7803. The work of HS is financed by the MEC and he thanks the Dep. ECM of the UB for the hospitality. \\end{theacknowledgments} \\newcommand{\\JHEP}[3]{{\\sl J. of High Energy Physics } {JHEP} {#1} (#2) {#3}} \\newcommand{\\NPB}[3]{{\\sl Nucl. Phys. } {\\bf B#1} (#2) {#3}} \\newcommand{\\NPPS}[3]{{\\sl Nucl. Phys. Proc. Supp. } {\\bf #1} (#2) {#3}} \\newcommand{\\PRD}[3]{{\\sl Phys. Rev. } {\\bf D#1} (#2) {#3}} \\newcommand{\\PLB}[3]{{\\sl Phys. Lett. } {\\bf #1B} (#2) {#3}} \\newcommand{\\EPJ}[3]{{\\sl Eur. Phys. J } {\\bf C#1} (#2) {#3}} \\newcommand{\\PR}[3]{{\\sl Phys. Rep } {\\bf #1} (#2) {#3}} \\newcommand{\\RMP}[3]{{\\sl Rev. Mod. Phys. } {\\bf #1} (#2) {#3}} \\newcommand{\\IJMP}[3]{{\\sl Int. J. of Mod. Phys. } {\\bf #1} (#2) {#3}} \\newcommand{\\PRL}[3]{{\\sl Phys. Rev. Lett. } {\\bf #1} (#2) {#3}} \\newcommand{\\ZFP}[3]{{\\sl Zeitsch. f. Physik } {\\bf C#1} (#2) {#3}} \\newcommand{\\MPLA}[3]{{\\sl Mod. Phys. Lett. } {\\bf A#1} (#2) {#3}} \\newcommand{\\CQG}[3]{{\\sl Class. Quant. Grav. } {\\bf #1} (#2) {#3}} \\newcommand{\\JCAP}[3]{{\\sl JCAP} {\\bf#1} (#2) {#3}} \\newcommand{\\APJ}[3]{{\\sl Astrophys. J. } {\\bf #1} (#2) {#3}} \\newcommand{\\AMJ}[3]{{\\sl Astronom. J. } {\\bf #1} (#2) {#3}} \\newcommand{\\APP}[3]{{\\sl Astropart. Phys. } {\\bf #1} (#2) {#3}} \\newcommand{\\AAP}[3]{{\\sl Astron. Astrophys. } {\\bf #1} (#2) {#3}} \\newcommand{\\MNRAS}[3]{{\\sl Mon. Not.Roy. Astron. Soc.} {\\bf #1} (#2) {#3}}" }, "0609/astro-ph0609360_arXiv.txt": { "abstract": "We present the age distributions for star clusters and individual stars in the Small Magellanic Cloud (SMC) based on data from the Magellanic Clouds Photometric Survey by Zaritsky and collaborators. The age distribution of the SMC clusters shows a steep decline, $dN_{cluster}/d\\tau \\propto \\tau^{-0.85\\pm0.15}$, over the period $10^7 \\lea \\tau \\lea 10^9$~yr. This decline is essentially identical to that observed previously for more massive clusters in the merging Antennae galaxies, and also for lower-mass embedded clusters in the solar neighborhood. The SMC cluster age distribution therefore provides additional evidence for the rapid disruption of star clusters (``infant mortality''). These disrupted clusters deliver their stars to the general field population, implying that the field star age distribution, $dN_{fld star}/d\\tau$, should have an inverse relation to $dN_{cluster}/d\\tau$ if most stars form initially in clusters. We make specific predictions for $dN_{fldstar}/d\\tau$ based on our cluster disruption models, and compare them with current data available for stars in the SMC. While these data do not extend to sufficiently young ages for a definitive test, they are consistent with a scenario wherein most SMC stars formed in clusters. Future analyses of $dN_{fldstar}/d\\tau$ that extend down to ages of $\\sim$ few million years are needed to verify the age relationship between stars residing in clusters and in the field. ", "introduction": "The age distribution of a population of star clusters contains important information about the formation and disruption of the clusters\\footnote{We use the term ``cluster'' generically to mean any aggregate of stars -- regardless of mass, size, or age, and whether bound or unbound -- with a density significantly higher than the local stellar background, making it recognizable as a distinct entity.}. The number of embedded clusters in the solar neighborhood (with masses $\\sim10^2-10^3~M_{\\odot}$) declines rapidly with age, and it has been estimated that only $4-7$\\% of these clusters will survive for 100~Myr (see the review by Lada \\& Lada 2003). We recently found that much more massive (compact) clusters ($\\geq3\\times10^4~M_{\\odot}$) in the merging ``Antennae'' galaxies also show a rapid decline in their age distribution, with $dN_{cluster}/d\\tau\\propto\\tau^{-1}$ (Fall, Chandar, \\& Whitmore 2005, hereafter FCW05). We interpret this decline in numbers as due to a high rate of early disruption (``infant mortality'') for massive clusters in the Antennae. Taken together, these age distributions suggest that compact, massive clusters in the chaotic environment of two merging disk galaxies disrupt at approximately the same rate as lower mass, embedded clusters in the more quiescent solar neighborhood. This suggests that star clusters disrupt rapidly due to processes internal to the clusters themselves (e.g., the removal of interstellar material due to the mass and energy input from massive stars; Hills 1980; Fall 2004; FCW05), and implies that the age distribution should be roughly $dN_{cluster}/d\\tau \\propto \\tau^{-1}$ in all star-forming galaxies. In contrast to the results for clusters in the Milky Way and in the Antennae, the cluster age distribution in the Small Magellanic Cloud (SMC) has been reported as $dN_{cluster}/d\\tau\\propto\\tau^{-2.1}$ (Rafelski \\& Zaritsky 2005; hereafter RZ05). However, this age distribution is actually the number of clusters divided by the number of individual stars of the same age, i.e., it is a {\\it normalized} cluster age distribution (RZ05). Does the proposed $\\tau^{-2}$ power law for the normalized cluster age distribution in the SMC differ from the $\\tau^{-1}$ form found in the Antennae and solar neighborhood because of the normalization procedure, because clusters in the SMC are forming and/or dissolving at a different rate, or for some other reason? The motivation for normalizing the number of clusters by the number of individual stars was to assess whether the formation of stars in clusters parallels the formation of stars in the field. An underlying assumption made by RZ05, then, is that star formation occurs in {\\it both} clusters and in the field simultaneously. However, our study of the Antennae galaxies suggests that at least 20\\% and possibly all of the $H\\alpha$ flux is emitted by compact clusters, which disrupt rapidly (FCW05). This picture implies instead that at least 20\\% and possibly all stars form in clusters, with stars being subsequently delivered from clusters to the field through the process of cluster disruption. This in turn implies that $dN_{cluster}/d\\tau$ and $dN_{fldstar}/d\\tau$ should have an inverse relationship to each other. If the SMC cluster system is found to have an unnormalized age distribution $\\propto \\tau^{-2.1}$, then our picture of clusters disrupting at essentially the same rate in all star-forming environments may have to be significantly revised. What is the unnormalized SMC cluster age distribution, does it differ from the normalized cluster age distribution, and what is the relationship between field and cluster stars? Addressing these questions is the main focus of this Letter. ", "conclusions": "We have shown here that the (unnormalized) cluster age distribution in the SMC has the form $dN_{cluster}/d\\tau \\propto \\tau^{-0.85\\pm0.15}$ and that overall the field star age distribution is nearly flat (i.e., $dN_{fldstar}/d\\tau\\propto \\tau^{0}$, over the age range $10^7 \\lea \\tau \\lea 10^9$~yr). The {\\it normalized} SMC cluster age distribution (the number of clusters divided by the number of individual stars of the same age) therefore must have the same form as the unnormalized distribution [i.e., $(dN_{cluster}/d\\tau)_{norm} \\propto \\tau^{-0.85\\pm0.15}$]. Our conclusion contradicts that of RZ05, who found that the normalized cluster age distribution is $(dN_{cluster}/d\\tau)_{norm} \\propto \\tau^{-2.1}$ from the same data. The difference in the power-law exponent between our work and that of RZ05 comes about because we divided the cluster age distribution ($dN_{cluster}/d\\tau$) directly by that for the field stars ($dN_{fldstar}/d\\tau$), whereas RZ05 divided $dN_{cluster}/d\\mbox{log}\\tau$ by $dN_{fldstar}/d\\mbox{log}\\tau$, and then appear to have introduced an extraneous factor of $\\tau^{-1}$. Our result that the age distribution of SMC clusters has a power-law form $dN_{cluster}/d\\tau \\propto \\tau^{-0.85\\pm0.15}$ has important implications regarding the physical processes that disrupt star clusters. The fact that clusters in vastly different environments (including the chaotic merging of two galaxies, a quiescent spiral disk, and a gas-rich dwarf galaxy) have age distributions with nearly identical forms provides strong evidence that star clusters disrupt rapidly due to processes {\\it internal} to the clusters themselves. Over the first few million years photoionization, stellar winds, jets and supernovae can remove much of the interstellar matter from clusters leaving the member stars freely expanding (e.g., Hills 1980; Boily \\& Kroupa 2003a,b). These expanding clusters will be observable for $\\sim$~few~$\\times 10^{7}$~yrs after becoming unbound, until their surface brightness becomes so low that they become indistinguishable from statistical fluctuations of stars in the field (FCW05). In addition to the removal of interstellar material, mass loss from stars themselves will likely contribute to the continued unbinding of clusters (e.g., Applegate 1986; Chernoff \\& Weinberg 1990; Fukushige \\& Heggie 1995). Finally, two-body evaporation will destroy any remaining clusters on longer timescales (e.g., Vesperini 1997; Baumgardt 1998; Fall \\& Zhang 2001). The early disruption (``infant mortality'') appears to operate relatively independent of cluster mass, while two-body evaporation is mass-dependent. The stars from disrupted clusters add to the general field, requiring a close relationship between the age distributions of clusters and field stars if most stars form initially in clusters. In this work we compared age distributions for clusters and individual stars in the SMC with model predictions to further test this general framework. Our models suggest that the rapid drop in the cluster age distribution should be reflected in a rapid rise in the field star age distribution. However this rapid rise in the number of field stars is only observable for the first $\\sim10^7$~yrs {\\it regardless of the fraction of stars that initially formed in clusters}. Because the present data on SMC stars begin just as the predicted field star distributions flatten, and because there is no discrimination between stars belonging to clusters versus to the field, they do not provide an unambiguous answer. However the observed relations are consistent with a scenario where most stars form in clusters that are subsequently disrupted by the mechanisms discussed above. The models also predict that even if all stars form initially in clusters, galaxies will contain a relatively small number of observable clusters at any given time. Future studies with the {\\it Hubble Space Telescope} (which allow better discrimination between stars in clusters and in the general field), and in the infrared (which can probe very young, deeply embedded stars and clusters) could be used to search for the expected rise in the field star age distribution at $\\tau \\lea 10^7$~yrs. A measurement of the rise in $dN_{fldstar}/d\\tau$ would provide a direct estimate of the fraction of stars which originally form in clusters versus in the field." }, "0609/astro-ph0609010_arXiv.txt": { "abstract": "The Southern Galactic Plane Survey (SGPS) is a radio survey in the 21 cm \\ion{H}{1} line and in 1.4~GHz full-polarization continuum, observed with the Australia Telescope Compact Array and the Parkes 64m single dish telescope. The survey spans a Galactic longitude of $253\\degr < l < 358\\degr$ and a latitude of $|b|<1\\degr$ at a resolution of 100 arcsec and a sensitivity below 1~mJy/beam. This paper presents interferometer only polarized continuum survey data and describes the data taking, analysis processes and data products. The primary data products are the four Stokes parameters I, Q, U, and V in 25 overlapping fields of $5\\degr.5$ by $2\\degr$, from which polarized intensity, polarization angle and rotation measure are calculated. We describe the effects of missing short spacings, and discuss the importance of the polarized continuum data in the SGPS for studies of fluctuations and turbulence in the ionized interstellar medium and for studying the strength and structure of the Galactic magnetic field. ", "introduction": "Galactic magnetism is one of the major components in the Milky Way, mostly in equipartition with gas and cosmic rays. Interstellar magnetic fields are believed to profoundly influence the ionized interstellar medium (ISM) through flux freezing and energy dissipation, affect star formation, determine the trajectories and acceleration of low and medium energy cosmic rays and play a major role in the turbulent gas dynamics (see e.g. reviews by Ferri\\`ere 2001, Scalo \\& Elmegreen 2004, Elmegreen \\& Scalo 2004). Knowledge about the strength and structure of the Galactic magnetic field is still sketchy. Yet, the field has received increasing attention, not only with the objective of studying magnetic fields in galaxies but also because the Galactic magnetized ISM forms a polarized foreground which needs to be determined for Cosmic Microwave Background Polarization \\citep{dto03} and Epoch of Reionization studies \\citep{mh04}. The only methods to probe Galactic magnetic fields in diffuse ionized gas over a large range of spatial scales are by way of radio polarization and Faraday rotation. Right and left circularly polarized components of radio emission experience birefringence while propagating through a magnetized and ionized medium. This causes the polarization angle of linearly polarized emission $\\phi$ to rotate as a function of wavelength $\\lambda$ as $\\Delta\\phi = \\mbox{RM}\\lambda^2$, where the rotation measure RM is RM~$=0.81\\int n_e \\mathbf{B}\\cdot\\mathbf{ds}$, $n_e$ is the thermal electron density in cm$^{-3}$, $\\mathbf{B}$ is the magnetic field vector in microGauss, $\\mathbf{ds}$ is the path length vector through the medium in parsecs, and the integral is along the line of sight from the observer to the source of polarized emission. Therefore, Faraday rotation measurements allow estimation of the magnetic field component along the line of sight, weighted by the electron density, and integrated over the pathlength. Depolarization characteristics can be used to determine the scale and amplitude of fluctuations in the medium. The observed polarized radiation used to trace the Galactic magnetic field can come from pulsars, polarized extragalactic sources or diffuse Galactic synchrotron emission (including supernova remnants). All of these sources have their own advantages and disadvantages. Pulsars are unique because model-dependant distance estimates allow constraining of the path length, and because a dispersion measure can be calculated, which in combination with RM yields a direct measure of the magnetic field averaged over the path length. However, they are scarce and distributed mainly in the Galactic plane. Unresolved extragalactic sources, on the other hand, are distributed all over the sky. But they have an intrinsic RM contribution, and currently published datasets yield an average of one source~deg$^{-2}$ in the Galactic plane (Brown et al.\\ 2003, Brown et al., in prep), and only 0.02-0.03 source~deg$^{-2}$ in the rest of the sky (e.g.\\ Simard-Normandin et al.\\ 1981, Broten et al.\\ 1988). Only diffuse synchrotron emission provides a pervasive background of polarized radiation which can be used to form RM maps of large fields in the sky with high resolution \\citep{hkb03a, hkb03b, ulg03, r04}. Diffuse synchrotron emission does suffer from depolarization, which decreases the possible measurements of RM, but which in itself can yield information about the fluctuations in the magneto-ionized interstellar medium \\citep{gdm01}. Recently the whole sky has been mapped in absolutely calibrated polarized continuum at 1.4~GHz (see Wolleben et al.\\ 2005 for the Northern sky, and Testori et al.\\ 2004 in the South) at a resolution of about half a degree. More than half of the Galactic plane is being surveyed at the much higher resolution of an arcmin, in two separate polarization surveys. The Canadian Galactic Plane Survey (CGPS, Taylor et al.\\ 2003) covers the Northern Sky at Galactic longitudes $74.2\\degr