{ "0601/astro-ph0601272_arXiv.txt": { "abstract": "We consider the properties of a hyperaccretion model for gamma-ray bursts (GRBs) at late times when the mass supply rate is expected to decrease with time. We point out that the region in the vicinity of the accretor and the accretor itself can play an important role in determining the rate of accretion, and its time behavior, and ultimately the energy output. Motivated by numerical simulations and theoretical results, we conjecture that the energy release can be repeatedly stopped and then restarted by the magnetic flux accumulated around the accretor. We propose that the episode or episodes when the accretion resumes correspond to X-ray flares discovered recently in a number of GRBs. ", "introduction": "Gamma-Ray Bursts (GRB) are generally believed to be powered by hyperaccretion onto a compact, stellar mass object. The total amount of the available fuel is considered to be the key factor determining the burst duration. Within merger scenarios for short-duration GRBs, a neutron star (NS) is accreted onto another NS or onto a stellar mass black hole (BH; e.g, Paczy\\'nski 1986, 1991; Eichler et al. 1989; Narayan et al. 1992; Fryer et al. 1999). Within the collapsar model for long-duration GRBs, up to 20 $\\MSUN$ of a stellar envelope collapses onto the star's core which is a NS or a BH (e.g., Woosley 1993; Paczy\\'nski 1998; MacFadyen \\& Woosley 1999; Popham, Woosley, \\& Fryer 1999; Proga et al. 2003). For short- and long-duration events, the accretion rate, $\\MDOT_a$ must be of order of 1~$\\MSUN~{\\rm s^{-1}}$, yielding a duration of less than a few seconds for the former and a duration as long as tens to hundreds of seconds for the latter. These duration estimates are made under the assumption that all the available fuel is accreted during the GRB activity at a time-averaged constant rate. Recent GRB observations obtained with {\\it Swift} motivate us to review the above assumption and some other aspects of GRB models. In particular, early X-ray afterglow lightcurves of nearly half of the long-duration GRBs show X-ray flares (Burrows et al. 2005; Romano et al. 2006; Falcone et al. 2006). X-ray flares are also found to follow the short-duration GRB 050724 (Barthelmy et al. 2005) whose host galaxy is early-type, which is consistent with the merger origin. The flares generally rise and fall rapidly, with typical rising and falling time scales much shorter than the epoch when the flare occurs. This time behavior strongly supports the ``internal'' origin of the flares (Burrows et al. 2005; Zhang et al. 2006; Fan \\& Wei 2005), in contrast to the ``external'' origin of the power-law decay afterglows. The internal model not only offers a natural interpretation of the rapid rise and decay behavior of the flares, but also demands a very small energy budget (Zhang et al. 2006). Within this picture, the data require a {\\it restart} of the GRB central engine (i.e., a restart of accretion). Fragmentations in the collapsing star (King et al. 2005) or in the outer parts of the accretion disc (Perna et al. 2006) have been suggested to be responsible for the observed episodic flaring behavior. These two flare models appeal to one of the basic ingredients of an accretion powered engine -- the mass accretion rate -- and conjecture that the episodic energy output is driven by changes in the mass supply and subsequently accretion rate. In this picture, the inner part of the accreting system {\\em passively} responds to changes in the accretion flow at larger radii. Here, we point out that the region in the vicinity of the accretor and the accretor itself can play an important role of determining the rate and time behavior of the accretion and the energy output. In particular, we conjecture that the energy release can be repeatedly stopped and then restarted, provided the mass supply rate decreases with time even if the decrease is smooth. For both merger and collapsar GRB models, a decrease of the mass supply rate is expected, especially in the late phase of activity, because the mass density decreases with increasing radius. In our model, we appeal to the fact that, as mass is being accreted onto a BH, the magnetic flux is accumulating in the vicinity of the BH. Eventually, this magnetic flux must become dynamically important and affect the inner accretion flow, unless the magnetic field is very rapidly diffused. In the remaining part of the paper we list and discuss theoretical arguments and results from a variety of numerical magnetohydrodynamic (MHD) simulations of accretion flows that support our model. We also provide analytic estimates to show that our model can quantitatively account for the observed features of the flares. \\section[]{Magnetic model for GRBs and their flares} \\subsection{Insights from numerical models} Generally, our model for the flares is based on the results from the numerical simulation of an MHD collapsar model for GRBs carried out by Proga et al. (2003) and the results from a number of simulations of radiatively inefficient accretion flows (RIAFs) onto a BH (Proga \\& Begelman 2003, PB03 hereafter; Igumenshchev, Narayan, \\&Abramowicz 2003, INA03 hereafter). Proga et al. (2003) performed time-dependent axisymmetric MHD simulations of the collapsar model. These MHD simulations included a realistic equation of state, neutrino cooling, photodisintegration of helium, and resistive heating. The progenitor was assumed to be spherically symmetric but with spherical symmetry broken by the introduction of a small, latitude-dependent angular momentum and a weak split-monopole magnetic field. The main conclusion from the simulations is that, within the collapsar model, MHD effects alone are able to launch, accelerate and sustain a strong polar outflow. The MHD outflow provides favorable initial conditions for the subsequent production of a baryon-poor fireball (e.g., Fuller, Pruet \\& Abazajian 2000; Beloborodov 2003; Vlahakis \\& K$\\ddot{\\rm o}$nigl 2003; M\\'{e}sz\\'{a}ros 2002), or a magnetically dominated ``cold fireball'' (Lyutikov \\& Blandford 2002), though the specific toroidal magnetic field geometry Proga et al. derived differs from some of these models (e.g., Vlahakis \\& K$\\ddot{\\rm o}$nigl 2003; Lyutikov \\& Blandford 2002). The latest Swift UV-Optical Telescope (UVOT) observations indicate that the early reverse shock emission is generally suppressed ( Roming et al. 2005), which is consistent with the suggestion that at least some GRBs are Poynting-flux-dominated outflows (Zhang \\& Kobayashi 2005). To study the extended GRB activity, one would like to follow the collapse of the entire star. However, such studies are beyond current computer and model limits. Therefore, we explore instead the implications of the published simulations and consider the physics of the collapsing star to infer the properties and physical conditions in the vicinity of a BH during the late phase of evolution, i.e., when a significant fraction of the total available mass is accreted. The long time evolution of axisymmetric MHD accretion flows was studied by PB03 who explored simulations very similar to those performed by Proga et al. (2003) but with much simpler micro physics (i.e., an adiabatic equation of state, no neutrino cooling or photodisintegration of helium). Proga et al. (2003) found that despite the more sophisticated micro physics of the MHD collapsar simulations the flow cooling is dominated by advection not neutrino cooling. As a result, the early phase of the time evolution, and the dynamics of the innermost flow, are very similar in both the RIAF simulations and the collapsar simulations. In particular, after an initial transient behavior, the flow settles into a complex convolution of several distinct, time-dependent flow components including an accretion torus, its corona and outflow, and an inflow and outflow in the polar funnel (see the left panel in Fig. 1 for a schematic picture of such a flow). The accretion through the torus is facilitated by the magnetorotational instability (MRI, e.g., Balbus \\& Halwey 1991) which also dominates the overall dynamics of the inner flow. In the remaining part of the paper, we will assume that the late evolution of the MHD collapsar simulations is similar to the late evolution of the RIAFs simulations. This assumption is justifiable because the flows in the collapsar and RIAFs simulations are similar during the early phase of the evolution (i.e., their dymanics and cooling are dominated by MRI and advection, respectively) The late evolution of RIAFs shows that the torus accretion can be interrupted for a short time by a strong poloidal magnetic field in the vicinity of a BH. This result is the main motivation for this paper, as it shows that the extended GRB activity may be a result of an accretion flow modulated by the ``magnetic-barrier'' and gravity. Because this barrier halts the accretion flow intermittently (see Figs.~6 \\& 8 in PB03), the accretion rate is episodic (see Fig.3 of PB03). This potentially gives a natural mechanism for flaring variability in the magnetic-origin models of GRBs as we first mentioned in Fan, Zhang \\& Proga (2005; see the middle panel of Fig. 1 here, for a cartoon picture of the accretion halted by the magnetic-barrier.) The importance of accumulating of the magnetic flux has been explored and observed by others in various astrophysical contexts (e.g., Bisnovatyi-Kogan \\& Ruzmaikin 1974, 1976; Narayan, Igumenshchev \\& Abramowicz 2003; INA03). In particular, INA03 carried out a three-dimensional (3D) MHD simulation (their model B) to late model times. They found that the magnetic flux accumulates, initially near the BH and then farther out, and the field becomes dynamically dominant. At late times, mass is able to accrete only via narrow streams, in a highly nonaxisymmetric manner (see also Narayan et al. 2003). The main difference between PB03's and INA03's results is the extent and duration of the magnetic dominance. In PB03, the magnetic dominance is a {\\em transient} whereas in INA03 is a {\\em persistent} state. The reason for this difference is the treatment of the magnetic field: for the initial conditions, PB03 used the split-monopole magnetic field and any changes in the magnetic flux near the BH during the evolution are due to the chaotic, small-scale fields generated in the disc. The detailed analysis show that the disc properties in PB03's simulations are determined by MRI. In particular, MRI is responsible for the complex field structure and for the disc toroidal field being one or even two orders of magnitude higher than the poloidal field (see figs. 9 and 10 in PB03 and fig. 3 in Proga et al. 2003). On the other hand, in their model B, INA03 set up a poloidal field configuration in the injected gas in such a way that the portion of the material that accretes always carries in the same sign of the vertical component of the magnetic field. The simulations carried out by PB03 and INA03 differ also in the assumed geometry (axisymmetric versus fully 3D). INA03 and PB03 do not explore all cases including the case where the external or initial field has zero net flux or the field with the poloidal component changing sign on length scales much smaller than the size of the mass reservoir \\footnote{ In the case where the initial or external flux has zero-net flux, a large scale coherent field might in some circumstances be generated by MRI (e.g., Livio, Pringle, \\& King 2003). If so the central magnetic flux could vary with time but still be dynamical signifacant for some periods of time.}. Additionally, these simulations also do not give definitive answers to the problems for which they were designed. Nevertheless, they give interesting insights into the general problem of MHD accretion flows. In particular, they suggest that magnetic fields can provide an important parameter determining the time scale for the accretion; i.e., it can be significantly longer than the local dynamical time scale. This can have important implications for the observed X-flares in GRBs, as we argue here, and for X-ray spectral states for BH binaries as discussed by Spruit \\& Uzdensky (2005, SU05 hereafter). In fact, the work by SU05 describes very well the general physics and theory of magnetic flux accumulated by an accretion flow. Therefore we now turn our attention to some theoretical aspects of the problem as presented by SU05. \\subsection{Theory of the magnetic barrier and accretion flow} SU05 considered a new mechanism of efficient inward transport of the large-scale magnetic field through a turbulent accretion disc. The key element of the mechanism is concentration of the external field into patches of field comparable in strength to the MRI turbulence in the disc. They focused on how to increase the magnetic flux at the center in the context of BH binaries. In particular, they argue that the capture of external magnetic flux by accretion disc and its subsequent compression in the inner regions of the disc may explain both changes in the radiation spectrum and jet activity in those objects. However, their model and physical arguments are generic and applicable to our problem. One can expect that as the strength of the magnetic field increases at the center, the field may eventually suppress MRI turbulence and reduce the mass accretion rate and the power in the outflow. This should be the case especially for GRBs because the mass inflow rate at the late time is most likely much lower than at the early time. The disc may become a Magnetically-Dominated Accretion Flow (MDAF) as proposed by Meier (2005) or the fields in the polar funnel can expand toward the equator and reconnect as in PB03's simulations. In the latter, the torus is pushed outward by the magnetic field. At this time, the gas starts to pile up outside the barrier; eventually it can become unstable to interchange instabilities at the barrier outer edge as suggested by SU05 or the gas in the torus can squash the magnetic field (compare Fig.~5 and 6 in PB05 or the middle and right panel in Fig.1 here). When interchange instabilities operate, magnetic flux from the bundle mixes outward into the disc while the disc material enters the barrier. In the accretion disk context, interchange instabilities have been studied by a few authors (e.g., Spruit et al. 1995; Lubow \\& Spruit 1995; Stehle 1996; Stehle \\& Spruit 2001; Li \\& Narayan 2004). These studies showed that the onset of small-scale modes typical of interchanges (as in Rayleigh-Taylor instabilities) takes place only at rather large field strengths, due to a stabilizing effect of the Keplerian shear. The interchange instability operates at moderate field strengths, but only at low shear rates (less than Keplerian). However for most of the time, we expect high shear rates in a torus because a low shear torus quickly becomes Keplerian due to MRI (e.g., PB03 and Proga et al. 2003). We note that SU05 interpreted INA03 accretion through the barrier, in the form of blobs and streams as a product of interchange instabilities. SU05 also suggested that the field strength at which these instabilities become effective is most usefully expressed in terms of the degree of support against gravity provided by the magnetic stress $B_R B_Z$. According to SU05, the instabilities become effective when the radial magnetic force, $F_m\\sim 2 B_R B_Z/4\\pi$, is of the order of a few percent of the gravitational force, $F_g=GM\\Sigma/R^2$, where $M$ is the central mass, $R$ is the radius, and $\\Sigma$ is the surface density. For $B_R \\approx B_Z$, there is a range in field strengths between the value at which MRI turbulence is suppressed and the value where dynamical instability of the barrier itself sets in, where no known instability operates (Stehle \\& Spruit 2001). In this range, the disk material cannot mix or penetrate the magnetic field accumulated at the center (e.g., the middle panel of Fig. 1). Instead, mass builds up outside a region with such field strengths until the magnetic field at the center is compressed enough for instability to set in. Thus, both numerical work and theoretical models of magnetized accretion flows show that the inner most part of the flow and accretor can respond {\\em actively} to changes of the accretion flow at larger radii. In particular, the inner most accretion flow can be halted for a very long time as shown by INA03 or it can be repeatedly halted and reactivated as shown in PB03. \\subsection{Analytic estimates} We finish this section with order-of-magnitude estimates of a few key features of our X-ray flare model. We start by estimating the strength and flux of magnetic field required to support the gas. The gas of the surface density, $\\Sigma$ can be supported against gravity by the magnetic tension if $F_g\\sim F_m$. The surface density can be estimated from $\\Sigma=\\MDOT /2\\pi R \\epsilon v_{ff}~{\\rm g~cm^{-2}}$, where $\\epsilon v_{ff}$ is the flow radial velocity assumed to be a fraction $\\epsilon$ of the free fall velocity, $v_{ff}$. Assuming $B_r \\approx B_z=B$, the force balance yields the field strength $B \\sim 2\\times 10^{16}~\\epsilon_{-3}^{-1/2} r^{-5/4} \\MDOT_1^{1/2} M_3^{-1} $~G, where $\\epsilon_{-3} \\equiv 10^{3} \\epsilon$, $r \\equiv R/R_S=R/(2GM_{BH}/c^2)$, $\\MDOT_1=\\MDOT/1~\\MSUN~{\\rm s^{-1}}$, and $M_3=M/3 \\MSUN$. We estimate the magnetic flux as $\\Phi\\sim\\pi r^2 R_S^2 B(r)= 5\\times10^{28}~\\epsilon_{-3}^{-1/2}r^{3/4}\\MDOT_ 1^{1/2} M_3~{\\rm cm^2~G}$ from which we obtain an estimate to the magnetospheric radius $r_m \\approx 60~\\epsilon_{-3}^{2/3} \\MDOT_1^{-2/3} M_3^{-4/3} \\Phi_{30}^{4/3}$, where $\\Phi_{30}\\equiv \\Phi/(10^{30}~{\\rm cm^2 G})$. Substituting the expression for $B$ into the expression for the surface density, one finds that a given magnetic flux can support the gas with the surface density of $\\Sigma_B=5\\times10^{19}~\\Phi_{30}^2 M_3^{-3}r^{-2}~{\\rm g~cm^{-2}}$. To stop accretion with the hyper rate of $1\\MSUN~{\\rm s^{-1}}$ onto a 3$\\MSUN$ black hole at r=3 (i.e., for $r_m$ to be 3), the magnetic flux of order $\\Phi_{30} \\sim 0.11$ is required. We now assume that such a magnetic flux is accumulated during hyperaccretion and that it does not change with time. Under these assumptions, $r_m=300$ for the mass supply rate of $10^{-3}~\\MSUNYR$ representative of the late time evolution . This relatively large radius demonstrates one of our key points that the innermost part an accreting system can actively respond, via magnetic fields, to changes in the inflow at large radii. To estimate the conditions needed to restart accretion, the accretion energetics and related time scales, we ask what is the mass of a disc with $\\Sigma$ high enough to reduce $r_m$ from 300 to 3 or so. To answer this question, we adopt Popham et al.' (1999) model of neutrino-dominated discs. Popham et al. assumed that neutrino cooling produces a thin disc (Shakura \\& Sunyaev 1973) for accretion rates require to power GRBs. Using the disc solution for the density and height (eqs. 5.4 and 5.5 in Popham et al. 1999), we can express the disc surface density as $\\Sigma_\\alpha=1.8\\times10^{19}~\\alpha^{-1.2} M_3^{-0.8}\\MDOT_1r^{-1.25}$~g, where $\\alpha$ is the dimensionless parameter scaling the stress tensor and the gas pressure (Shakura \\& Sunyaev 1973). Equating $\\Sigma_B$ with $\\Sigma_\\alpha$, one can estimate the mass accretion rate of an $\\alpha$ disc and compute $M_D$ by integrating $\\Sigma_\\alpha$ over radius. For $\\Phi_{30}=0.11$ and $\\alpha=10^{-2}$ the accretion rate through the $\\alpha$ disc is $0.03~\\MSUN~s^{-1}$ and $M_D$ for $r$ between 3 and 300 is 0.32~$\\MSUN$. This mass accretion rate is more than one order of magnitude lower than the rate of $\\sim 1~\\MSUN~s^{-1}$ typical for the early time evolution. Thus, our estimates are consistent with the fact that the X-ray flare luminosity is at least one or two orders of magnitude lower the prompt gamma-ray emission (see section 3). If this disc mass is a result of slow mass accumulation during the late evolutionary stage, then it will take about 400 s to rebuild the disc for the mass supply rate of $10^{-3}~\\MSUN~s^{-1}$ and 12 s to accrete all this mass at the disc accretion rate of $0.03~\\MSUN~s^{-1}$. The latter is a lower estimate for the flare duration because, for simplicity, we assumed a relatively high, {\\em constant} disc accretion rate. It is very likely that the rate changes with time as the shape of the light curve during the flares indicates. In our model, the mass supply rate controls the epochs when the flares happen: the disc is rebuilt on the time scale which increases with time because the mass supply slowdowns. Additionally, the flare duration is coupled to the epoch through the mass of the rebuilt disc. Thus our model is capable of accounting for the observed duration - time scale correlation. ", "conclusions": "The detailed analysis of the X-ray flares revealed that they generally have lower luminosities (by at least one or two orders of magnitude) than the prompt gamma-ray emission. Additionally, the total energy of the flare is also typically smaller than that of the prompt emission, although in some cases both could be comparable (e.g. for GRB 050502B, Falcone et al. 2006). Moreover multiple flares are observed in some GRBs and the durations of these flares seem to be positively correlated with the epochs when the flares happen, i.e. the later the epoch, the longer the duration (O'Brien et al. 2005; Falcone et al. 2006; Barthelmy et al. 2005). The flare analysis also showed that the later the epoch the lower the flare luminosity. The above qualitative properties of the flares provide important constraints on models of them. Perna et al.'s (2006) disc fragmentation model promises to account for the duration - time scale correlation and the duration - peak luminosity anticorrelation. However, the physical process or processes causing fragmentation are uncertain. It is also uncertain that the conditions for the disc fragmentation are met in GRB progenitors. This seems to be the case especially for the collapsar model as a relatively high rotation of the progenitor is required. We also note that magnetic fields can suppress or even prevent disc fragmentation (e.g., Banerjee \\& Pudritz 2006). Here, we propose that the X-ray flares in GRBs are consequences of the fact that during the late time evolution of a hyperaccretion system the mass supply rate should decrease with time while the magnetic flux accumulating around a BH should increase. In particular, we point out that the flux accumulated during the main GRB event can change the dynamics of the inner accretion flow. We argue that the accumulated flux is capable of halting intermittently the accretion flow. In our model, the episode or episodes when the accretion resumes correspond to X-ray flares. A comparison of our analytic estimates from Section 2.3 with the observed X-ray flare characteristics, shows that our model is not only physically based but also can both qualitatively and quantitatively account for some aspects of the flares -- such as the peak times. In general, our model fits under the general label of the magnetic jet model for GRBs as we appeal to the magnetic effects to play the key role not only during the main event but also during the late evolution. The importance of the magnetic effects for the X-ray flares can be argued based on energy budget of the accretion model (Fan et al. 2005). The X-ray flares discovered in GRBs are relatively new and unexpected phenomena. They give a strong incentive to apply the existing models of hyperaccretion systems to circumstances where the mass supply is reduced. Studies of this kind should reveal whether one needs to introduce additional physics in order to explain the flares. If so one should explore the effects of this on the early evolution of GRBs and check whether they are consistent with GRBs observations. Our X-ray flare model has the advantage that it is essentially the same as the MHD collapsar model for GRBs, with only one justifiable change in a key physical property of the collapsar model: a decrease of the mass supply rate with time." }, "0601/astro-ph0601044_arXiv.txt": { "abstract": "We study the bending of light for static spherically symmetric (SSS) space-times which include a dark energy contribution. Geometric dark energy models generically predict a correction to the Einstein angle written in terms of the distance to the closest approach, whereas a cosmological constant $\\Lambda$ does not. While dark energy is associated with a repulsive force in cosmological context, its effect on null geodesics in SSS space-times can be attractive as for the Newtonian term. This dark energy contribution may be not negligible with respect to the Einstein prediction in lensing involving clusters of galaxies. Strong lensing may therefore be useful to distinguish $\\Lambda$ from other dark energy models. ", "introduction": "It is still unclear what drives the universe into acceleration recently. While a cosmological constant $\\Lambda$ is the simplest explanation, its value seems completely at odd with the naive estimate of the vacuum energy due to quantum effects. An alternative to $\\Lambda$ is obtained considering a dynamical degree of freedom added to the primordial soup. Since dynamical dark energy (dDE) varies in time and space, its fluctuations are potentially important in order to distinguish it from $\\Lambda$ \\cite{CDS}. Another possibility for the explanation of the present acceleration of the universe is given by the geometry itself, through a modification of Einstein gravity at large distances \\cite{DDG}: these are known as geometric dark energy (gDE) models. A non vanishing mass for the graviton is among these possibilities \\cite{GG}. {\\em Cosmological} observations, such as those coming from Supernovae, cosmic microwave background (CMB) anisotropies and large scale structure (LSS), have not been able to discriminate among dDE models yet (see \\cite{UIS} for updated constraints and forecasts). It is therefore important to explore observational tests at {\\em astrophysical} level for objects which are detached from the cosmological expansion. ", "conclusions": "We have discussed light bending in SSS with DE. The importance of general relativistic tests, such as the perihelion precession, has been already emphasized for the DGP model \\cite{DGZ,LS}, while little attention was previously paid to light bending. These tests are complementary to the observational signatures of dark energy in cosmological context, mainly based on the behaviour of perturbations. In cosmology, an important difference between $\\Lambda$ and dDE or gDE is the presence of DE perturbations in the latter case, which are at least gravitationally coupled to the other types of matter. Such DE perturbations are therefore a key point to distinguish $\\Lambda$ from dDE or gDE in CMB and LSS, and sometimes may become so important to strongly constrain models \\cite{CDS,CF,koyama} with respect to what Supernovae data can do. In this article we have shown that in objects which have detached from the expansion of the universe, $\\Lambda$ may be distinguishable from other DE models through the bending of light. In order to link our findings with observations, we should insert $\\varphi$ in the lens equation, e.g. \\cite{schneider}: \\begin{equation} \\theta-\\beta=\\frac{d_{SL}}{d_{OS}} \\varphi \\end{equation} where $\\theta$ and $\\beta$ are the angular positions of the image and of the source measured respect to the line from the observer to the lens; $d_{LS}$ and $d_{OS}$ are the angular diameter distances between the lens and the source and between the observer and the source, respectively. On considering for simplicity alignment between the lens and the source, an Einstein ring forms with angle $\\theta_{E} = \\theta (\\beta = 0)$. From our results, $\\theta_E$ is affected by both the non-perturbative SSS potential around the lens ($ \\varphi \\ne 4 G M/r_0$ if DE $\\ne \\Lambda$) and the cosmology of a given model. The gDE corrections to the Einstein deflection angle for clusters in Eq. (\\ref{dephi}) are as important as the cosmology for an observable as $\\theta_E$. The differential of $\\theta_E$ is \\be \\frac{\\Delta \\theta_E}{\\theta_E} = \\frac{\\Delta \\varphi}{\\varphi} + \\Delta \\ln \\frac{d_{SL}}{d_{OS}} \\,, \\ee which reveals how cosmological information is encoded just in the second term to the right. By considering the cosmology of the DGP model for instance \\cite{LUE}, one finds that the second term is $\\sim -0.06 (\\Delta \\Omega_{\\rm M}/\\Omega_{\\rm M}) + \\Delta H_0/H_0$ for a source and a lens located at $z = 1$ and $z =0.3$, respectively ($\\Omega_{\\rm M} \\sim 0.3$ and $H_0$ are the present matter density and Hubble parameter, respectively). On considering the uncertainties on the cosmological parameters of the order of percent, this simple quantitative example shows how the corrections to the Einstein deflection angle we have found in Table I should be taken into account in the study of strong lensing by clusters. We believe that results similar to what we have found here for gDE models, might occur for dDE scenarios as well, in which the non-asymptotically flat term is due to the non-perturbative clumping of DE into objects detached from the cosmological expansion. However, dDE models may be less predictive than gDE models: gDE contain the same number of parameters of $\\Lambda$CDM, while dDE may need more. Let us end on noting that some of the gDE models considered here may have serious theoretical issues \\cite{BD,koyama2}, whose resolution clearly go beyond the present project. However, the main result in Eq. (\\ref{dephi}) of this paper remains valid: models alternative to general relativity with a cosmological constant predict a correction to the Einstein angle, which can be used to distinguish $\\Lambda$ from other DE models. \\vspace{.5cm} {\\bf Acknowledgements} We wish to thank Lauro Moscardini for many discussions and suggestions on galaxies, groups and clusters. We are grateful to Robert R. Caldwell and one of the anonymous referees for valuable comments on this project. FF and MG are partially supported by INFN IS PD51; FF and and AG are partially supported by INFN IS BO11. The authors thank the Galileo Galilei Institute for Theoretical Physics for the hospitality and the INFN for partial support during the developments of this project." }, "0601/astro-ph0601334_arXiv.txt": { "abstract": "We present a {\\it Hubble Space Telescope} Advanced Camera for Surveys (ACS) weak-lensing study of RX J0849+4452 and RX J0848+4453, the two most distant (at $z=1.26$ and $z=1.27$, respectively) clusters yet measured with weak-lensing. The two clusters are separated by $\\sim4\\arcmin$ from each other and appear to form a supercluster in the Lynx field. Using our deep ACS $i_{775}$ and $z_{850}$ imaging, we detected weak-lensing signals around both clusters at $\\sim4\\sigma$ levels. The mass distribution indicated by the reconstruction map is in good spatial agreement with the cluster galaxies. From the singular isothermal sphere (SIS) fitting, we determined that RX J0849+4452 and RX J0848+4453 have similar projected masses of $(2.0\\pm0.6)\\times10^{14} M_{\\sun}$ and $(2.1\\pm0.7)\\times10^{14} M_{\\sun}$, respectively, within a 0.5 Mpc ($\\sim60\\arcsec$) aperture radius. In order to compare the weak-lensing measurements with the X-ray results calibrated by the most recent low-energy quantum efficiency determination and time-dependent gain correction, we also re-analyzed the archival $Chandra$ data and obtained $T=3.8_{-0.7}^{+1.3}$ and $1.7_{-0.4}^{+0.7}$~keV for RX J0849+4452 and RX J0848+4453, respectively. Combined with the X-ray surface brightness profile measurement under the assumption of isothermal $\\beta$ model, the temperature of RX J0849+4452 predicts that the projected mass of the cluster within $r=0.5~$Mpc is $2.3_{-0.4}^{+0.8}\\times10^{14} M_{\\sun}$, consistent with the weak-lensing analysis. On the other hand, for RX J0848+4453 we find that the mass derived from this X-ray analysis is much smaller ($6.3_{-1.5}^{+2.6}\\times10^{13} M_{\\sun}$) than the weak-lensing measurement. One possibility for this observed discrepancy is that the intracluster medium (ICM) of RX J0848+4453 has not yet fully thermalized. Although this interpretation is rather simplistic, the relatively loose distribution of the cluster galaxies in part supports this possibility of low degree of virialization. We also discuss other scenarios that might give rise to the discrepancy. ", "introduction": "It has become clear that massive clusters are not extremely rare at high redshifts ($z>0.8$) and the presence of these large collapsed structures when the age of the Universe is less than half its present value is no longer in conflict with our current understanding of the structure formation, especially in a $\\Lambda$-dominated flat cosmology. Pursuit of galaxy clusters to higher and higher redshift is important in the extension of the evolutionary sequences to earlier epochs, when the effect of the different cosmological frameworks becomes more discriminating. A great deal of observational efforts have been made in the last decade in enlarging the sample of high-redshift clusters. X-ray surveys have provided an efficient method of cluster identification and probe of cluster properties because a hot intracluster medium (ICM) within the cluster generates strong diffuse X-ray emission and is believed to be in quasi-equilibrium with gravity. However, it is questionable how well the clusters selected by their X-ray excess can provide the unbiased representation of the typical large scale structure at the cluster redshift. If the degree of the virialization decreases significantly with redshift and is strongly correlated with X-ray temperature, the cosmological dimming $\\sim (1+z)^{-4}$ can bias our selection progressively towards higher and higher mass, relaxed structures. Among other important approaches to detect high-redshift clusters is a red-cluster-sequence (RCS) survey using the distinctive spectral feature in cluster ellipticals. This so-called 4000\\AA~break feature is well-captured by a careful combination of two passbands, and Gladders \\& Yee (2005) recently reported 67 candidate clusters at a photometric redshift of $0.9 < z < 1.4$ from the $\\sim10$\\% subregion of the total $\\sim100~\\mbox{deg}^2$ RCS survey field. A related method but giving a higher contrast of cluster members with respect to the background sources is to use deep near-infrared (NIR) imaging (e.g., Stanford et al. 1997) for the selection of cluster candidates. High-redshift clusters identified in these color selection methods are expected to serve as less biased samples encompassing the lower mass regime at high redshifts. In the current paper, we study two $z\\sim1.3$ clusters, namely RX J0849+4452 and RX J0848+4453 (hereafter Lynx-E and Lynx-W for brevity), using the deep F775W and F850LP (hereafter $i_{775}$ and $z_{850}$, respectively) images obtained with the Advanced Camera for Surveys (ACS) on the $Hubble$ $Space$ $Telescope$ ($HST$). Interestingly, although these two clusters are separated by only $\\sim4\\arcmin$ from each other, they were discovered by different methods. Standford et al. (1997) discovered Lynx-W in a NIR survey as an overdense region of the $J-K > 1.9$ galaxies and spectroscopically confirmed 8 cluster members. They also analyzed the archival ROSAT-PSPC observation of the region and found diffuse X-ray emission near the confirmed cluster galaxies. However, they could not rule out the possibility that the X-ray flux might be coming from the foreground point sources because of the PSPC PSF is too broad to identify such objects. The subsequent study of the field using the $Chandra$ observations showed that, although the previous ROSAT-PSPC observation is severely contaminated by the X-ray point sources adjacent to the cluster, the cluster is still responsible for some diffuse X-ray emission. Both the X-ray temperature and luminosity of the cluster appear to be low ($T_X\\sim1.6$~keV and $L_{bol}\\sim0.69\\times10^{44} ~ \\mbox{ergs}~\\mbox{s}^{-1}$; Stanford et al. 2001). Lynx-E was, on the other hand, first discovered in the ROSAT Deep Cluster Survey (RDCS) as a cluster candidate and follow-up near-infrared imaging showed an excess of red ($1.81$ we expect to have a growing list of low-mass clusters that are also X-ray dark because of the evasively low-temperature, as well as the substantial cosmological dimming. Therefore, it is plausible to suspect that these two $z\\sim1.3$ clusters (especially Lynx-W, the poorer X-ray system) might lie on a border where the X-ray observations alone start to become insufficient to infer the mass properties. Weak-lensing provides an alternative approach to deriving the mass of a gravitationally bound system without relying on assumptions about the dynamical state. This can help us to probe the properties of the high-redshift clusters in lower mass regimes, where the X-ray measurements alone may not provide useful physical quantities. In our particular case, weak-lensing is an important tool to test how the masses of the two Lynx clusters at $z\\sim1.3$ compare with their X-ray measurements. Especially for Lynx-W, weak-lensing seems to be the unique route for probing the cluster mass, considering the poor and amorphous X-ray emission. Another interesting question is whether the low X-ray temperature of Lynx-W arises simply from a low mass or from a yet poor thermalization of the ICM. However, the detection of weak-lensing signal at $z\\sim1.3$ is difficult and much more so if the lens is not very massive. In our previous investigation of the two $z\\sim0.83$ high-redshift clusters (Jee et al. 2005a, hereafter Paper I; Jee et al. 2005b, hereafter Paper II), we were able to detect clear lensing signals. They revealed the complicated dark matter substructure of the clusters in great detail. The effective source plane (defined by the effective mean redshift of the background galaxies) in Paper I and II is at $z_{eff}\\sim1.3$, corresponding to the redshift of the lenses targeted in the current paper! Therefore, the number density of background galaxies decreases substantially compared to our $z\\sim0.8$ studies and, in addition, the higher fraction of non-background population in our source sample inevitably dilutes the resulting lensing signal quite severely. Furthermore, the accurate removal of instrumental artifacts becomes more critical as stronger signals come from more distant, and thus fainter and smaller galaxies. They are more severely affected by the point-spread-function (PSF). Nevertheless, our analyses of RDCS 1252-2927 at $z=1.24$ (Lombardi et al. 2005; Jee et al. in preparation) demonstrate that weak-lensing can still be applied to clusters even at these redshifts and reveals the cluster mass distribution with high significance. Returning to the X-ray properties, the low-energy quantum efficiency (QE) degradation of the $Chandra$ instrument can cause noticeable biases in cluster temperature measurements. Although there have been many suggestions regarding this issue, it was not until recently that a convergent prescription to remedy the situation has become available from the $Chandra$ X-ray Center\\footnote{see http://cxc.harvard.edu/ciao3.0/threads/apply\\_acisabs/ or http://cxc.harvard.edu/ciao3.2/releasenotes/}. Because we suspect that the previous X-ray analyses of the Lynx clusters suffered from the relatively insufficient understanding of this problem, we have also re-analyzed the archival $Chandra$ data to enable a fairer comparison between the weak-lensing and X-ray measurements. Throughout the paper, we assume a $\\Lambda$CDM cosmology favored by the Wilkinson Microwave Anisotropy Probe (WMAP), where $\\Omega_M$, $\\Omega_{\\Lambda}$, and $H_0$ are 0.27, 0.73, and 71 $\\mbox{km}~\\mbox{s}^{-1}~\\mbox{Mpc}^{-1}$, respectively. All the quoted uncertainties are at the 1 $\\sigma$ ($\\sim68$\\%) level. ", "conclusions": "} We have presented a weak-lensing analysis of the two Lynx clusters at $\\bar{z}=1.265$ using the deep ACS $i_{775}$ and $z_{850}$ images. Our mass reconstruction clearly detects the dark matter clumps associated with the two high-redshift clusters and other intervening objects within the ACS field, including the known foreground cluster at $z=0.57$. In order to verify the significance of the cluster detection and to separate the high-redshift signal from the low-redshift contributions, we performed a weak-lensing tomography by selecting an alternate lower-redshift source plane. This second mass reconstruction does not show the mass clumps around the high-redshift clusters, while maintaining most of the other structures seen in the first mass map. This experiment strongly confirms that the weak-lensing signals observed in the first mass reconstruction are real and come from the high-redshift Lynx clusters. Interestingly, both clusters are found to have similar weak-lensing masses of $\\sim 2.0\\times 10^{14} M_{\\sun}$ within 0.5 Mpc ($\\sim60\\arcsec$) aperture radius despite their discrepant X-ray properties. Our re-analysis of the Chandra archival data with the use of the latest calibration of the low-energy QE degradation shows that Lynx-E and W have temperatures of $T=3.8_{-0.7}^{+1.3}$ and $1.7_{-0.4}^{+0.7}$~keV, respectively. Combined with the X-ray surface brightness profile measurements, the X-ray temperature of Lynx-E gives a mass estimate in good agreement with the weak-lensing result. On the other hand, the X-ray mass of Lynx-W is much smaller than the weak-lensing estimation nearly by a factor of three. According to our experiment in \\textsection\\ref{section_mass_estimate}, it is unlikely that any foreground contamination or cosmic shear effect in weak-lensing measurement causes this large discrepancy. Apart from a simplistic, but valid possibility that Lynx-W might have a filamentary structure extended along the line of sight, yielding a substantial, projected mass but with yet only low-temperature thermal emission, we can also consider the self-similarity breaking (e.g., Ponman et al. 1999; Tozzi \\& Norman 2001; Rosati, Stefano, \\& Norman et al. 2002) typically observed for low-temperature X-ray systems. There have been quite a few suggestions that a non-gravitational heating (thus extra entropy) might prevent the ICM from further collapsing at the cluster core. The effect is supposed to be more pronounced in colder systems whose virial temperature is comparable to the temperature created by this non-gravitational heating, leading to shallower gas profiles than those of high-temperature systems (e.g., Balogh et al. 1999; Tozzi \\& Norman 2001). Interestingly, our determination of the surface brightness profile of Lynx-W is much shallower ($\\beta=0.42\\pm0.07$) than that of Lynx-E ($\\beta=0.71\\pm0.12$) (however, Ettori et al. (2004) obtained $\\beta=0.97\\pm0.43$ for Lynx-W). The relatively loose distribution of the cluster galaxies in Lynx-W without any apparent BCG defining the cluster center leads us to consider another possibility that the system might be dynamically young and the ICM has not fully thermalized within the potential well. If we imagine that the ICM is not primordial, but has been ejected from the cluster galaxies at some recent epoch, it is plausible to expect that the X-ray temperature of the ICM might yet under-represent the depth of the cluster potential well. Tozzi et al. (2003) investigated the iron abundance in the ICM at $0.3$350,000 stars seen towards M31. One of the brightest of these, LGGS 004341.84+411112.0, turns out to have a spectrum that is remarkably similar to that of the P Cygni, one of the best studied LBVs in the Milky Way. The star has been relatively constant in $B$ over the past 40 years, with variations $<0.2$~mag, but with evidence of variations of 0.05~mag during a year. Much the same can be said for P Cygni (Israelian \\& de Groot 1999). However, the star has likely had an outburst two millennia ago, as judged by the fact that the star is slightly extended {\\it HST} images, indicative of a 0.5~pc nebula, similar to the 0.2~pc nebula seen around P~Cygni. This has consequences for interpreting the nature of objects found with spectroscopic similarity to other known LBVs; these stars are considered LBV ``candidates\" until variability is established, but the present study emphasizes that this may require more than a lifetime. If we assume that the spectral similarity of our star to that of P Cygni is indicative of a similar effective temperature (and hence intrinsic color and bolometric correction), we can place the star on the H-R diagram using our photometry. The star falls on the evolutionary track for a 85$M_\\odot$ star at the extreme end its evolution to cooler temperature, just as the evolutionary tracks turns to lower luminosity (Meynet \\& Maeder 2005). The star is likely in a transition between an O star and a WNL. This provides strong vindication of the current generation of stellar evolutionary models that include rotation, but it also notes the difficulties in interpreting the evolution phases implied by the models. An atmospheric analysis of the star is planned, and long-term follow-up (both photometric and spectroscopic) is warranted." }, "0601/astro-ph0601428_arXiv.txt": { "abstract": "Among the light elements created in the Big Bang, deuterium is one of the most difficult to detect but is also the one whose abundance depends most sensitively on the density of baryons. Thus, although we still have only a few positive identifications of D at high redshifts---when the D/H ratio was close to its primordial value---they give us the most reliable determination of the baryon density, in excellent agreement with measures obtained from entirely different probes, such as the anisotropy of the cosmic microwave background temperature and the average absorption of the UV light of quasars by the intergalactic medium. In this review, I shall relate observations of D/H in distant gas clouds to the large body of data on the local abundance of D obtained in the last few years with {\\it FUSE\\/}. I shall also discuss some of the outstanding problems in light element abundances and consider future prospects for advances in this area. ", "introduction": "The measurement of the interstellar abundance of deuterium was one of the main science drivers of the \\emph{FUSE} mission right from its inception. Five years on, this promise has been amply fulfilled, as demonstrated by the numerous talks and posters at this meeting devoted to \\emph{FUSE} results on D/H. The importance of deuterium stems from the fact that, among the light elements created in Big Bang nucleosynthesis (BBNS), it is the one whose primordial abundance responds most sensitively to cosmological density of baryons, $\\Omega_{\\rm b}$. While $^4$He is the most abundant, because it soaks up essentially all the available neutrons, this property also makes it a rather insensitive `baryometer'. The quantity $\\Omega_{\\rm b}$, or more precisely the baryon to photon ratio $\\eta$, only affects $Y_{\\rm p}$ (the primordial mass fraction in $^4$He) by determining the time delay before BBNS can set in---and thus the time available for neutrons to decay---in the first few minutes of the universe history. The more fragile deuterium, on the other hand, is easily destroyed by two-body reactions with protons, neutrons and other nuclei so that its abundance relative to hydrogen when BBNS is over, (D/H)$_0$ or D$_0$ for short, shows a steep, inverse, dependence on $\\Omega_{\\rm b}$. $^7$Li is less useful than D in this respect because it is far less abundant, by about five orders of magnitude, and its dependence on $\\Omega_{\\rm b}$ is double-valued because it can be synthesised via different nuclear reactions in the high and low baryon density regimes. The detection of interstellar D was among the first discoveries made by \\emph{FUSE}'s predecessor, the \\emph{Copernicus} satellite. Rogerson \\& York (1973) resolved the isotope shift in the Ly$\\gamma$ line seen towards the bright B1 III star $\\beta$~Centauri, and deduced $N$(\\ion{D}{1})/$N$(\\ion{H}{1})$\\,= (1.4 \\pm 0.2) \\times 10^{-5}$. Three decades later, the mean of 21 measurements of D/H in the `Local Bubble' (the nearby region of the Milky Way disk) is in excellent agreement with \\emph{Copernicus}' first detection: $\\langle$D/H$\\rangle = (1.56 \\pm 0.04) \\times 10^{-5}$ (Wood et al. 2004).\\footnote{An interesting observation is that the distance to $\\beta$~Cen has `doubled' since 1973. The \\emph{Hipparcos} parallax to this star implies a distance $d = 161\\,$pc, while the parallactic distance available to Rogerson \\& York (1973) was $d = 81$\\,pc. This is a clear demonstration that it is easier for astronomers to measure chemical abundances than distances, even to the brightest stars.} An important property of deuterium is that it is only destroyed whenever interstellar gas is cycled through stars (a process termed astration), so that its abundance relative to H steadily decreases with the progress of galactic chemical evolution. Analytically, this reduces to a simple expression for the time evolution of D: \\begin{equation} {\\rm (D/D_0)}~ = ~f^{(1/\\alpha~ -1)}~ = ~e^{-Z(1/\\alpha~ -1)} \\end{equation} where $f$ is the gas fraction, $Z$ the metallicity (in units of the yield of a primary element such as oxygen) and $\\alpha$ is the mass fraction which is locked up in long-lived stars and stellar remnants whenever a quantity $M$ of interstellar matter is turned into stars (Ostriker \\& Tinsley 1975; Pagel 1997). Equation (1) is valid in the simplest case of a `closed-box' model of chemical evolution. More realistic models which include inflow and/or outflow generally result in lower reductions of the primordial D/H as a function of either $f$ or $Z$ (Edmunds 1994). We cannot measure the lock-up fraction $\\alpha$ directly, but only deduce it theoretically by assuming a distribution of stellar masses (the IMF) and guessing at what fraction of its initial mass each star returns to the interstellar medium (ISM). In principle, the degree of astration suffered by D in the Milky Way, where $f = 0.1 - 0.2$, could be anywhere between 20\\% and 90\\%, depending on the uncertain value of $\\alpha$. Consequently, Rogerson and York could only use their measurement of D/H in the ISM today to place a lower limit on (D/H)$_0$ and a corresponding upper limit $\\Omega_{\\rm b} \\leq 0.0675$ (for $h = 0.7$ where, as usual, $h$ is the Hubble constant in units of 100\\,km~s$^{-1}$~Mpc$^{-1}$. In this article I shall use $h = 0.7$ throughout and dispose of the $h^2$ term in the value of $\\Omega_{\\rm b}$). Note that the above upper limit on $\\Omega_{\\rm b}$, even without any correction for D astration, implies that most of the matter in the universe is non-baryonic. ", "conclusions": "We have come a long way since that pioneering measurement by Rogerson \\& York of the interstellar abundance of deuterium with {\\em Copernicus\\/} more than thirty years ago. The number of such measurements now approaches fifty, thanks in particular to the capabilities of {\\em FUSE\\/} and the GHRS and STIS instruments on the {\\em Hubble Space Telescope\\/}. With large ground-based telescopes we have been able to probe high redshift clouds where D is still close to its primordial abundance. New methods have been exploited to determine the density of baryons, the most impressive of which is the mapping of the temperature anisotropies in the cosmic microwave background over a wide range of angular scales. Bringing all of these developments together we find that many aspects of the overall picture fit together remarkably well, giving us confidence in the validity of the whole cosmological framework. Others still provide challenging puzzles, particularly the unexplained dispersion in the local determinations of D/H and the very low abundance of $^7$Li in some of the oldest stars of our Galaxy. But I am optimistic that we will not have to wait another three decades to iron out these remaining wrinkles in our understanding of the abundances of the light elements." }, "0601/astro-ph0601664_arXiv.txt": { "abstract": "{We investigated abundance ratios along the profiles of six high-redshift Damped Lyman-$\\alpha$ systems, three of them associated with H$_2$ absorption, and derived optical depths in each velocity pixel. The variations of the pixel abundance ratios were found to be remarkably small and usually smaller than a factor of two within a profile. This result holds even when considering independent sub-clumps in the same system. The depletion factor is significantly enhanced only in those components where H$_2$ is detected. There is a strong correlation between [Fe/S] and [Si/S] abundances ratios, showing that the abundance ratio patterns are definitely related to the presence of dust. The depletion pattern is usually close to the one seen in the warm halo gas of our Galaxy. ", "introduction": "\\label{sec:introduction} Damped Lyman-$\\alpha$ systems (hereafter DLAs) observed in QSO spectra are characterized by strong H~{\\sc i}$\\lambda$1215 absorption lines with broad damping wings. Although the definition has been restricted for historical reasons to absorptions with log~$N$(H~{\\sc i})~$>$~20.3 (Wolfe et al. 1986), damping wings are easily detected in present-day, high quality data for much lower column densities (down to log~$N$(H~{\\sc i})~$\\sim$~18.5). A more appropriate definition should be related to the physical state of the gas. If we impose the condition that the gas must be neutral, then the definition should be limited to systems with log $N$(H~{\\sc i})~$>$~19.5 (e.g. Viegas 1995). Since their discovery twenty years ago \\citep{Wolfe86}, DLAs clearly have something to do with galaxy formation. What kind of galaxy DLAs are associated to is, however, still a matter of debate. Some authors identify these systems with large rotating discs \\citep{Prochaska97,Hou01}, while others think that DLAs arise mostly either in dwarf galaxies \\citep{Centurion} or galactic building blobs \\citep{Haehnelt98,Ledoux98}. The answer is probably not unique. In any case, DLAs represent the major reservoir of neutral hydrogen at any redshift \\citep{Storrie2000}, and they probe the chemical enrichment and evolution of the neutral Universe (see Pettini et al. 1994; Lu et al. 1996; Prochaska et al. 1999; Ledoux et al. 2002a and references therein). Since abundances can be measured very accurately in DLAs, we can both discuss the connection between observed abundance ratios and dust content and to trace the nucleosynthesis history of the dense gas in the universe. In this context, it is helpful to compare these results with measurements in the ISM of our Galaxy. Refractory elements that condense easily into dust grains - namely, Cr, Fe, Ni- are strongly depleted (up to a factor hundred) in the ISM, while non-refractory elements remain in its gaseous phase - S, Zn, and partially Si-. The amount of depletion depends on the physical condition of the gas. Thus, different depletion patterns are observed depending on whether the gas is cold or warm and/or whether the gas is located in the disc or the halo of the Galaxy \\citep{Savage96}. The LMC and SMC also exhibit different gas-phase abundance ratios \\citep{Welty99}. However, a particular nucleosynthesis history can give rise to peculiar metallicity patterns and mimic the presence of dust. \\citet{Tinsley} suggested that type Ia supernova are the major producers of Fe. An enhancement in [$\\alpha$/Fe] ratios ($\\alpha$-elements are mostly O, S, Si) could reflect an IMF skewed to high masses and therefore a predominant role of type II supernova. For very low metallicity stars ([Fe/H]~$<$~$-$3) in the Galaxy, large variations in several abundance ratios have been reported \\citep{william}, which suggests that peculiar nucleosynthesis processes and inhomogeneous chemical enrichment are probably taking place. As mentioned above, DLAs trace the chemical evolution of galaxies at early epochs in the universe. Many detailed studies have been performed so far, revealing that their metallicities range between 1/300 $Z_{\\odot}$ and solar values. The abundance pattern is fairly uniform and compatible with low dust content (see Pettini et al. 1994, Lu et al. 1996, Prochaska et al. 1999, Ledoux et al. 2002a). This uniformity in the relative abundance patterns observed from one DLA to the other has been emphasized by Prochaska \\& Wolfe (2002) and suggests that protogalaxies have common enrichment histories. Few studies have adressed the question of the homogeneity inside each particular system. Prochaska \\& Wolfe (1996) first studied chemical abundance variations in a single DLA, and showed that the chemical abundances were uniform to within statistical uncertainties. Lopez et al. (2002) confirmed this finding from analysis of another DLA using Voigt profile decomposition. Petitjean et al. (2002) and Ledoux et al. (2002b) showed that the depletion patterns in subcomponents were very similar along DLA profiles except in the components where molecular hydrogen is detected and where depletion is larger. More recently, \\citet{Prochaska03}, performed a study of 13 systems concluding that the majority of DLAs have very uniform relative abundances. This contrasts in particular with the dispersion in nucleosynthetic enrichment of the Milky Way as traced by stellar abundances. Here, we use the best data from our survey of DLAs (Ledoux et al. 2003) to investigate this issue further using an inversion method to derive the velocity profiles in different abundance ratios. In particular, we investigate the consequence of the presence of molecular hydrogen in some of the DLAs. The paper is structured as follows: we describe the data in Sect. 2; in Sect. \\ref{sec:method} we briefly introduce the method used for the analysis; and results are presented and discussed in Sects. \\ref{sec:results} and \\ref{sec:discussion}. ", "conclusions": "\\label{sec:discussion} \\subsection{Relative abundance ratios }\\label{subsec:histo} \\begin{figure*} \\centering \\includegraphics[height=5.5cm,width=7.5cm]{2504f13a.ps} \\includegraphics[height=5.5cm,width=7.5cm]{2504f13b.ps} \\includegraphics[height=5.5cm,width=7.5cm]{2504f13c.ps} \\includegraphics[height=5.5cm,width=7.5cm]{2504f13d.ps} \\includegraphics[height=5.5cm,width=7.5cm]{2504f13e.ps} \\includegraphics[height=5.5cm,width=7.5cm]{2504f13f.ps} \\caption{Depletion of heavy elements relative to Sulfur (except for Q~1157+014, where Zinc was used instead) in the different subclumps of the six DLA systems studied here. Filled symbols represent the different sub-systems considered (see Table~\\ref{tab:res}). Error bars correspond to typical scatter for each sub-system. The histograms show the observed values in the cold (solid line) or warm (dashed line) disc clouds and in halo clouds (dotted line) of the Galaxy.} \\label{fig:depletion_pattern} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[height=10.5cm,width=13.5cm]{2504f14.ps} \\caption{[Fe/S] vs [Si/S] for all the subclumps analyzed in this paper. Different symbols represent different DLA: squares for Q~0013$-$004, diamonds for Q~0528$-$250, triangles for Q~0405$-$443, and circles for Q~1037$-$270. Open symbols are used to distinguish subclumps where H$_2$ is detected. Otherwise symbols are filled. We also indicate the typical [Fe/S] vs [Si/S] values observed in the cold, warm ISM, and halo of our Galaxy from \\citet{Welty99}.} \\label{fig:correl} \\end{figure*} \\begin{figure*} \\centering \\includegraphics[height=5.5cm,width=8.0cm]{2504f15a.ps} \\includegraphics[height=5.5cm,width=8.0cm]{2504f15b.ps} \\caption{Summary of the inhomogeneity amplitude observed in the subsystems. The y-axis represents the maximum deviation of $[Fe/S]\\pm2\\sigma$ in each subclump. Subsystems with molecules are respresented with filled symbols.} \\label{fig:inhomog} \\end{figure*} Relative abundance ratios for each sub-clump considered in the six systems analyzed in this work are given in Table~\\ref{tab:res}. In the same way, results are summarized in Fig.~\\ref{fig:depletion_pattern}. It is apparent that, within each subsystem, large departures from the mean ratio are rare: the scatter is small if we take the observational and fitting uncertainties into account . Moreover, when we compare depletion values from one sub-clump to another in the same system, differences are small. A distinct depletion pattern is observed only for some molecular components .This is remarkably summarized by Fig.~\\ref{fig:correl}, where we plot the [Fe/S] ratio versus the [Si/S] ratio in all subclumps. First, as already emphasized by Petitjean et al. (2002) and Ledoux et al. (2002b), the sequence seen in this figure is a dust-depletion sequence. Indeed, there is a correlation between the two quantities which is expected if the depletion is due to the presence of dust. Secondly, the values measured in different clumps of the same system are gathered at the same place in the figure. The only exception is the H$_2$ component at $-$480~km~s$^{-1}$ in Q~0013$-$004 (see above). Thirdly, most of the depletion pattern is similar to that of the gas observed in the Galactic halo. Finally, it seems that silicon is overabundant by about 0.2~dex even relative to sulfur. In all this, however, it must be recalled that we do not have access to the absolute metallicity in the subclumps, because we are not able to disentangle the H~{\\sc i} absorptions of the different subclumps. In the left hand panel of Fig.~\\ref{fig:inhomog}, we plot the scatter, measured as $\\sigma$, of the ratio [Fe/S] in all the subclumps considered. In this case, we considered the subclumps as a whole, not isolating the molecular component, as our aim was to see if there is a relation between the inhomogeneity of a system and its molecular content. The mean value of $\\sigma$ over the subclumps is 0.3, which means that inhomogeneities are less than a factor of 2. Only a few subclumps where H$_2$ is detected have larger $\\sigma$. This is expected because we have seen that depletion is larger over the specific small velocity ranges over which H$_2$ is detected. In the right hand panel of Fig.~\\ref{fig:inhomog}, we plot the different scatter values for each subclump as a function of the total [Fe/S] ratio. This figure confirms that (i) larger [Fe/S] ratios are observed in subclumps where H$_2$ is detected with one exception in a subclump of Q~0013$-$0004, and that (ii) the scatter is larger for subclumps where H$_2$ is detected. \\subsection{The presence of \\hdos~ molecules}\\label{subsec:molecules} \\cite{Ledouxsurvey} have systematically searched for molecular hydrogen in high redshift DLAs, with a $\\sim$20\\% detection rate over the whole sample. The observed molecular fraction is often much smaller than in the ISM of the Galactic disc (Rachford et al. 2002) and is closer to what is observed in the magellanic clouds (Tumlinson et al. 2002). Here, we confirm what was already noticed by Ledoux et al. (2002b) and \\citet{ppjq0013} that, although the presence of molecules sometimes reveals gas with larger depletion into dust grains than average, this is not always the case. In most of the systems, the depletion factor is only a factor of two larger in the components with H$_2$ compared to the overall system. There are a few exceptions, the most noticeable being the molecular component at $-$480~km~s$^{-1}$ toward Q~0013$-$014, in which depletion is as large as in the cold gas of the Galactic disc. \\subsection{Consequences}\\label{subsec:overall} Variations in the relative abundance pattern within DLAs are expected from different nucleosynthesis histories and from its depletion onto dust. The magnitude of variation in the nucleosynthesis pattern may depend on the history of star formation and the level of enrichment. For example, stars in the Milky Way, Magellanic clouds, and local dwarf galaxies with metallicities varying from solar to 1/100 of solar show a dispersion in [Si/Fe] of the order of 0.3~dex (e.g. Shetrone et al. 2001, Venn et al. 2001). Peculiar nucleosynthesis histories may be reflected in the variation of abundance ratios from one subclump to the other. Although we observed differences in the relative metal abundances of different sub-clumps, they are not large. This may indicate that sub-clumps in DLAs have the same origin and history and could be part of the same object. This contrasts with the large differences in absolute metallicities that have been observed in LLS with similar velocity differences (e.g. D'Odorico \\& Petitjean 2001). That depletion onto dust depends on the local physical conditions should induce a large scatter in the observed pixel-to-pixel relative abundance ratios. The fact that only small scatter was observed may reveal that the gas in DLAs is neither very dense nor cold but rather diffuse and warm. At least the filling factor of highly depleted gas is small. All this implies uniform physical conditions and homogeneous and efficient mixing. One can speculate that this is only possible if DLAs are small objects with dimensions on the order of one kilo-parsec. This is difficult to ascertain as direct detection of high-redshift DLAs have not been very successful till now (e.g. Kulkarni et al. 2000, M\\o ller et al. 2004). It is, however, very important to pursue these observations in order to better constrain the nature and physical properties of these objects." }, "0601/astro-ph0601387_arXiv.txt": { "abstract": "We present the on-going observational program of a VIMOS Integral Field Unit survey of the central regions of massive, gravitational lensing galaxy clusters at redshift $z \\simeq 0.2$. We have observed six clusters using the low-resolution blue grism ($R \\simeq 200$), and the spectroscopic survey is complemented by a wealth of photometric data, including {\\it Hubble Space Telescope} optical data and near infrared VLT data. The principal scientific aims of this project are: the study of the high-$z$ lensed galaxies, the transformation and evolution of galaxies in cluster cores and the use of multiple images to constrain cosmography. We briefly report here on the first results from this project on the clusters Abell 2667 and Abell 68. ", "introduction": "\\label{sec:1} Because of their intense gravitational field, massive (i.e., $M > 10^{14} M_{\\odot}$) clusters of galaxies act as Gravitational Telescopes (GTs) and are therefore an important tool to investigate the high-redshift Universe (see, e.g., \\cite{gt}). In order to fully exploit the scientific potential of the GTs we have started an extensive integral field spectroscopy (IFS) survey of massive galaxy clusters. Targets have been selected among well-known gravitational lensing clusters between redshift $\\sim 0.2$ and 0.3, for which complementary {\\it Hubble Space Telescope} (HST) data are available. All the clusters are X-ray bright sources. Our sample partially overlaps with the one analyzed in \\cite{smith05}. IFS is the ideal tool in order to obtain spatially complete spectroscopic information of compact sky regions such as cluster cores. Moreover, cluster cores are the regions where strong lensing phenomena are observed (i.e., giant arcs and multiply imaged sources): for clusters in our sample, strongly lensed galaxies are within $\\theta \\simeq 1$ arcmin from the cluster center. VIMOS-IFU \\cite{lefevre03} is thus the natural choice for such observational program, since, at present, it provides the largest f.o.v. for among integral field spectrographs mounted on the 8-10m telescopes. All the clusters in our sample have been observed using the low-resolution blue (LR-B) grism, with a spectral resolution $R \\sim 200$. Taking into account the lower efficiency at the end of the spectra and the zero order contaminations, the final useful spectral range is limited between $\\simeq 3900$ \\AA \\, and 6800 \\AA. This spectral range is suitable both for detecting high-redshift source (e.g., Ly$\\alpha$ emitters in the redshift range $2.2 < z < 5.5$ or [OII] emitters out to $z\\sim 0.8$) and to sample the rest-frame 4000 \\AA \\, break for the cluster galaxy population. With a fiber size 0.66 arcsec, the IFU f.o.v. covers a contigous region of $54\\times 54$ arcsec$^2$, sampled by 6400 fibers. A subset of the clusters in the sample has also been observed using a higher resolution grism ($R\\simeq3000$) covering the $\\lambda= 6300-8600 \\, $ \\AA\\ range. These observations are useful to probe higher redshift Ly$\\alpha$ emitters at $4.20^{o}$). Covering the remaining 34 percent of our selected region, we detected 60 objects, including 41 known PNe and 19 new PNe. \\ha $+$\\nitrogen\\ images as well as low resolution spectra of the new PNe were taken which confirmed the photoionized nature of the emission. About 84 percent of the detected objects have angular sizes $\\leq$ 20--25 arcsec, while all show \\ha/\\oiii\\ ratios less than 1. Four (4) of our new PNe are found to be associated with IRAS sources. Radio sources for fifteen (15) of our new PNe (including those of paper I) were identified. All new PNe display low N/O and He abundances implying old progenitor stars, which is one of the characteristics of the Galactic bulge PNe. In a forthcoming paper (paper III), the use of the photoionization model CLOYDY will provide a deeper insight to the physical parameters of the new PNe. The information presented in papers I and II, along with the results that will be obtained from the CLOUDY model will offer a valuable tool to studies of the dynamics and kinematics of the Galactic bulge. \\begin{figure*} \\centering \\scalebox{0.90}{\\includegraphics{Fig1.eps}} \\caption[]{Optical imaging survey grid in equatorial coordinates. Galactic coordinates are also included (dash lines) to permit an accurate drawing of the selected Bulge region (bold solid lines). The dark filled and light filled rectangles represent the remaining observed field and the observed fields in the year 2000, respectively.} \\label{fig01} \\end{figure*} \\begin{figure*} \\centering \\scalebox{0.90}{\\includegraphics{Fig2a.eps}} \\caption[]{\\ha $+$\\nitrogen~images of all new PNe taken with the 1.3 m telescope. The arrows point to their position. The images have a size of 150\\arcsec on both sides. North is at the top, East to the left.} \\label{fig02a} \\end{figure*} \\begin{figure*} \\centering \\addtocounter{figure}{-1} \\scalebox{0.90}{\\includegraphics{Fig2b.eps}} \\caption[]{continued} \\label{fig02b} \\end{figure*} \\begin{figure*} \\centering \\scalebox{0.90}{\\includegraphics{Fig3a.eps}} \\caption[]{Observed spectra of our new PNe taken with the 1.3 m telescope. They cover the range of 4750\\AA\\ to 6815\\AA\\ and the emission line fluxes (in units of $10^{-16}$ erg s$^{-1}$ cm$^{-2}$ arcsec$^{-2}$ \\AA$^{-1}$) are corrected for atmospheric extinction. Line fluxes corrected for interstellar extinction are given in Table 2.} \\label{fig03a} \\end{figure*} \\begin{figure*} \\centering \\addtocounter{figure}{-1} \\scalebox{0.90}{\\includegraphics{Fig3b.eps}} \\caption[]{continued} \\label{fig03b} \\end{figure*} \\begin{figure*} \\centering \\addtocounter{figure}{-1} \\scalebox{0.65}{\\includegraphics{Fig3c.eps}} \\caption[]{continued} \\label{fig03c} \\end{figure*} \\begin{figure*} \\centering \\scalebox{0.65}{\\includegraphics{Fig4.eps}} \\caption[]{(a) IRAS colour--colour diagram of 13 of our objects (stars), overlaid on the colour--colour diagram of Acker's catalogued PNe with good (circles) and not good (rectangles) quality fluxes and (b) Diagnostic diagram (Garcia et al. 1991), where the positions of the new PNe are shown with a triangle ($\\bigtriangleup$). For comparison, the position of (i) the new PNe presented in paper I are also shown with a cross (X) and (ii) other objects (supernova remnants - SNR, H {\\sc ii} regions and Herbig Haro objects - HH) are shown, too.} \\label{fig04} \\end{figure*} \\begin{figure*} \\centering \\scalebox{0.80}{\\includegraphics{Fig5.eps}} \\caption[]{(a) Enlargement of the \\ha $+$\\nitrogen~image of PTB34 in high contrast in order to show its outer halo (b) Image of the same PN overlaid with the different scale contours as derived from the method described in the text. North is in the top, east to the left in both figures.} \\label{fig05} \\end{figure*}" }, "0601/astro-ph0601452_arXiv.txt": { "abstract": "{Deep X--ray surveys are providing crucial information on the evolution of AGN and galaxies. We review some of the latest results based on the X--ray spectral analysis of the sources detected in the Chandra Deep Field South, namely: i) constraints on obscured accretion; ii) constraints on the missing fraction of the X--ray background; iii) the redshift distribution of Compton--thick sources and TypeII QSO; iv) the detection of star formation activity in high--z galaxies through stacking techniques; v) the detection of large scale structure in the AGN distribution and its effect on nuclear activity. Such observational findings are consistent with a scenario where nuclear activity and star formation processes develop together in an anti--hierarchical fashion. ", "introduction": "In the last years deep X--ray surveys with the {\\sl Chandra} and {\\sl XMM--Newton} satellites (Brandt et al. 2001; Rosati et al. 2002; Alexander et al. 2003; Hasinger et al. 2001), paralleled by multiwavelength campaigns (see, e.g., GOODS, Giavalisco et al. 2004), provided several crucial information on the evolution of the AGN and galaxy populations. The bold result from the two deepest X--ray fields, the Chandra Deep Field North (CDFN, observed for 2 Ms) and the Chandra Deep Field South (CDFS, observed for about 1Ms) is constituted by the resolution of the X--ray background (XRB) into single sources, mostly AGN, at a level between 80\\% and 90\\% (see Bauer et al. 2004 and the recently revised estimate by Hickox \\& Markevitch 2005), providing an almost complete census of the accretion history of matter onto supermassive black holes through the cosmic epochs. However, the most interesting outcomes go well beyond the demographic characterization of the extragalactic X--ray sky. Indeed, the physical and evolutionary properties of the AGN population are now revealing how they formed and how they are linked to their host galaxies. For the first time, the luminosity function of AGN has been measured up to high redshift. A striking feature is the {\\sl downsizing}, or {\\sl anti--hierarchical} behaviour, of the nuclear activity: the space density of the brightest Seyfert I and QSO is peaking at $z\\geq 2$, while the less luminous Seyfert II and I peak at $z\\leq 1$ (Ueda et al. 2003; Hasinger et al. 2005; La Franca et al. 2005). An analogous behaviour is presently observed in the cosmic star formation history: at low redshift star formation is mostly observed in small objects (see, e.g., Kauffmann et al. 2004), while at redshift 2 or higher, star formation activity is observed also in massive galaxies (with $M_* \\sim 10^{11} M_\\odot$, see, Daddi et al. 2004a). The global picture, as outlined by the present data, requires a tight link between the formation of the massive spheroids and the central black holes, as witnessed by the relation between black hole and stellar masses or between black hole mass and the velocity dispersion of the bulge (Kormendy \\& Richstone 1995; Magorrian et al. 1998; Ferrarese \\& Merritt 2000). The anti--hierarchical behaviour in both star formation and AGN activity (which reflects in an anti--hierarchical supermassive black holes growth, see Merloni 2004; Marconi et al. 2004; for an alternative view see Hopkins et al. 2005), is envisaged by theoretical models where energy feedback is invoked to self--regulate both processes (see Fabian 1999; Granato et al. 2004). In these Proceedings, we will describe a few observational results obtained from the latest analysis of the X--ray and optical data in the Chandra Deep Field South and North and which, in our view, are consistent with this picture. In detail, these results concern the following issues: \\begin{itemize} \\item the physical properties of AGN from the X--ray spectral analysis of faint X--ray sources; \\item the missing fraction of the XRB; \\item the distribution of obscured QSO and Compton--thick sources and their relation with the cosmic mass accretion history; \\item star formation in high--z galaxies measured in the X--ray band thanks to stacking techniques; \\item the effects of large scale structure onto nuclear activity. \\end{itemize} ", "conclusions": "We presented few selected topics which we find particularly relevant among the latest results from deep X--ray surveys. We can summarize our conclusions as follows: \\begin{itemize} \\item a population of strongly absorbed, possibly Compton--thick AGN at $z\\sim 1$ is still missing to the census of the X--ray sky; the detailed X--ray spectral analysis of faint sources shows that we are detecting some of them, and help us in obtaining a complete reconstruction of the cosmic accretion history onto supermassive black holes; \\item we find several absorbed sources (the so--called TypeII QSO) among the population of bright AGN, possibily witnessing the rapid growth of the super massive black holes associated to strong star formation events; \\item thanks to stacking techniques, we detected the X--ray emission associated to massive star forming galaxies at redshift as high as $z\\sim 2$, therefore peering with X--rays in the epoch of massive galaxy formation; \\item investigation of large scale structure in the X--ray detected AGN distribution provides tantalizing hints of its effect on nuclear activity. Studies of spatial correlation of X--ray sources require larger fields of view to kill the cosmic variance and therefore evaluate properly the evolution of the AGN clustering properties. \\end{itemize} Such observational findings are providing crucial information on the evolution of AGN and galaxies. Present--day data are consistent with a scenario where nuclear activity and star formation processes develop together in an anti--hierarchical fashion." }, "0601/astro-ph0601208_arXiv.txt": { "abstract": "\\begin{center} \\end{center} Born-Infeld electromagnetic waves interacting with a static magnetic background are studied in an expanding universe. The non-linear character of Born-Infeld electrodynamics modifies the relation between the energy flux and the distance to the source, which gains a new dependence on the redshift that is governed by the background field. We compute the luminosity distance as a function of the redshift and compare with Maxwellian curves for supernovae type Ia. ", "introduction": "The discovery of an unexpected diminution in the observed energy fluxes coming from supernovae type Ia \\cite{Riess-Perlmutter}, which are thought of as standard candles, has been interpreted in the context of the standard cosmological model as evidence for an accelerating universe dominated by something called dark energy. This fact is one of the most puzzling and deepest problems in cosmology and fundamental physics today. Although the cosmological constant seems to be the simplest explanation for the phenomenon, several dynamical scenarios have been tried out (see, for instance, \\cite{padmanabhan} and references therein). It is worthwhile to emphasize that the evidence for an accelerating universe mainly relies on energy flux measurements for type Ia supernovae at different values of cosmological redshifts. They provide the most direct and consistent way to determine the recent expansion history of the universe. Nevertheless the relation between the cosmological redshift and the energy flux for a point-like source involves not only the evolution of the universe during the light journey but some assumptions about the nature of the light itself. Customarily, one accepts the linear Maxwell theory to describe the light propagation, where light propagates without interacting with other electric or magnetic fields. However, in the context of non-linear electrodynamics the interaction between the light emitted from such distant sources and cosmological magnetic backgrounds modifies the relation between the redshift and the flux of energy. If this kind of effect were not correctly interpreted it could lead to an erroneous conclusion about the expansion history of the universe. Concretely, an effect coming from non-linear electrodynamics could explain the curves of luminosity distance vs redshift for type Ia supernovae without invoking dark energy. This remark drives us to study the propagation of non-linear electromagnetic waves in an expanding universe with a magnetic background. We will benefit from recently obtained results for Born-Infeld electromagnetic plane waves propagating in a magnetic uniform background in Minkowski space-time \\cite{9abf}. Born-Infeld electrodynamics \\cite{1born,2born} is a non-linear theory for the electromagnetic field $F_{\\mu\\nu}$ governed by the Lagrangian \\begin{equation} {\\cal L}=\\sqrt{-g}\\ \\frac{b^2}{4\\:\\pi}\\,\\left(1-\\sqrt{1+\\frac{2S}{b^2}- \\frac{P^2}{b^4}}\\right) \\end{equation} where $b$ is a new fundamental constant and $S$ and $P$ are the scalar and pseudoscalar field invariants \\begin{equation} S=\\frac{1}{4}F_{\\mu\\nu}F^{\\mu\\nu}\\ \\ \\ \\ \\ \\ \\ \\ \\ \\ P=\\frac{1}{4}\\:^{*}F_{\\mu\\nu}F^{\\mu\\nu} \\end{equation} (in Minkowski space-time it is $2S=|{\\bf{\\mathbb B}}|^2-|{\\bf{\\mathbb E}}|^2$ and $P= {\\bf{\\mathbb E}}\\cdot{\\bf{\\mathbb B}}$, $\\bf{\\mathbb E}$ and $\\bf{\\mathbb B}$ being the electric and magnetic fields respectively). Born-Infeld electrodynamics goes to Maxwell electromagnetism when $b\\rightarrow \\infty$. In particular, the Born-Infeld field of a point-like charge reaches the finite value $b$ at the charge position and becomes Coulombian far from the charge. Born-Infeld theory is the only non-linear spin 1 field theory displaying causal propagation and absence of birefringence \\cite{deser,5boi}. Nowadays, Born-Infeld theory is reborn in the context of superstrings because Born-Infeld-like Lagrangians emerge in the low energy limit of string theories \\cite{strings}. Born-Infeld-like Lagrangians have been also proposed to describe a matter dynamics able to drive the universe to an accelerated expansion \\cite{varios}. In spite of this revival of Born-Infeld's ideas, there is no experimental evidence for Born-Infeld effects in electrodynamics: the value of $b$ remains unknown (see an upper bound for $b$ in \\cite{Jack}). In the next section we will summarize the recently obtained results on Born-Infeld waves propagating in Minkowski space-time in the presence of a uniform magnetic background \\cite{9abf}. These results --properly adapted-- will be used in section III to understand the Born-Infeld energy flux coming from a point-like source in a spatially flat Friedman-Robertson-Walker (FRW) expanding universe. In section IV we will reformulate the relation between the luminosity distance $d_L$ and the redshift $z$ within the framework of Born-Infeld electrodynamics. We will show that the presence of magnetic backgrounds modify the curves $d_L$ vs $z$, and we will analyze the consequences for the measurements of the luminosity distance of supernovae type Ia. ", "conclusions": "In this paper we have solved the Born-Infeld equations for electromagnetic plane waves propagating in a background magnetic field. In the absence of a background field, the Born-Infeld plane waves are equal to the Maxwell ones. On the contrary, in the presence of a background magnetic field ${\\bf B}$ the non-linear effects modify both the phase and the amplitude of the wave with corrections that depend on the combination $\\vert{\\bf B}\\vert^2\\, a^{-4}\\, b^{-2}$, where $a$ is the scale factor of the universe. It is remarkable that Born-Infeld electrodynamics depends on $a$ and $b$ only through the combination $a^4 b^2$. This means that the Maxwellian approximation ($b\\rightarrow\\infty$) also corresponds to the limit $a\\rightarrow\\infty$. So, although the electromagnetic field is presently well described by Maxwell equations for a wide range of phenomena, the non-linear Born-Infeld electrodynamics could have an influence in the past when the scale factor was smaller. Therefore the expanding universe is a good laboratory to test Born-Infeld electrodynamics; many non-linear aspects of its equations could be relevant when highly redshifted objects are observed. In this work we have begun the search for this kind of effects. We found that the influence of Born-Infeld electrodynamics on the luminosity distance (\\ref{Hod5}) exhibits interesting features that could be experimentally established by means of more precise supernova observations and a better knowledge of the cosmological background fields. Firstly, the experimental data for $d_L$ vs $z$ could be fitted without invoking dark energy, although there is no observational evidence of the background field that would be required. Secondly, the shape of the curve $d_L$ vs $z$ predicted by the standard cosmology ($\\Omega_m=0.3$, $\\Omega_\\Lambda=0.7$, $b\\rightarrow \\infty$) for high redshifts differs appreciably from the one predicted by Born-Infeld electrodynamics, which opens the possibility of detecting non-linear electrodynamics effects in a future." }, "0601/astro-ph0601514_arXiv.txt": { "abstract": "Name{ABSTRACT}% \\def\\abstract{% \\endmode \\bigskip\\bigskip \\centerline{\\AbstractName}% \\medskip \\bgroup \\let\\endmode=\\endabstract \\narrower\\narrower \\singlespaced \\everyabstract}% \\def\\everyabstract{}% \\def\\endabstract{\\smallskip\\egroup} \\def\\pacs#1{\\medskip\\centerline{PACS numbers: #1}\\smallskip} \\def\\submit#1{\\bigskip\\centerline{Submitted to {\\sl #1}}} \\def\\submitted#1{\\submit{#1}}% \\def\\toappear#1{\\bigskip\\raggedcenter To appear in {\\sl #1} \\endraggedcenter} \\def\\disclaimer#1{\\footnote{}\\bgroup\\tenrm\\singlespaced This manuscript has been authored under contract number #1 \\@disclaimer\\par} \\def\\disclaimers#1{\\footnote{}\\bgroup\\tenrm\\singlespaced This manuscript has been authored under contract numbers #1 \\@disclaimer\\par} \\def\\@disclaimer{% with the U.S. Department of Energy. Accordingly, the U.S. Government retains a non-exclusive, royalty-free license to publish or reproduce the published form of this contribution, or allow others to do so, for U.S. Government purposes. \\egroup} \\catcode`@=11 \\chardef\\other=12 \\def\\center{% \\flushenv \\advance\\leftskip \\z@ plus 1fil \\advance\\rightskip \\z@ plus 1fil \\obeylines\\@eatpar}% \\def\\flushright{% \\flushenv \\advance\\leftskip \\z@ plus 1fil \\obeylines\\@eatpar}% \\def\\flushleft{% \\flushenv \\advance\\rightskip \\z@ plus 1fil \\obeylines\\@eatpar}% \\def\\flushenv{% \\vskip \\z@ \\bgroup \\def\\flushhmode{F}% \\parindent=\\z@ \\parfillskip=\\z@}% \\def\\endcenter{\\endflushenv} \\def\\endflushleft{\\endflushenv} \\def\\endflushright{\\endflushenv} \\def\\@eatpar{\\futurelet\\next\\@testpar} \\def\\@testpar{\\ifx\\next\\par\\let\\@next=\\@@eatpar\\else\\let\\@next=\\relax\\fi\\@next} \\long\\def\\@@eatpar#1{\\relax} \\def\\raggedcenter{% \\flushenv \\advance\\leftskip\\z@ plus4em \\advance\\rightskip\\z@ plus 4em \\spaceskip=.3333em \\xspaceskip=.5em \\pretolerance=9999 \\tolerance=9999 \\hyphenpenalty=9999 \\exhyphenpenalty=9999 \\@eatpar}% \\def\\endraggedcenter{\\endflushenv}% \\def\\hcenter{\\hflushenv \\advance\\leftskip \\z@ plus 1fil \\advance\\rightskip \\z@ plus 1fil \\obeylines\\@eatpar}% \\def\\hflushright{\\hflushenv \\advance\\leftskip \\z@ plus 1fil \\obeylines\\@eatpar}% \\def\\hflushleft{\\hflushenv \\advance\\rightskip \\z@ plus 1fil \\obeylines\\@eatpar}% \\def\\hflushenv{% \\def\\par{\\endgraf\\indent}% \\hbox to \\z@ \\bgroup\\hss\\vtop \\flushenv\\def\\flushhmode{T}}% \\def\\endflushenv{% \\ifhmode\\endgraf\\fi \\if T\\flushhmode \\egroup\\hss\\fi \\egroup}% \\def\\flushhmode{U} \\def\\endhcenter{\\endflushenv} \\def\\endhflushleft{\\endflushenv} \\def\\endhflushright{\\endflushenv} \\newskip\\EnvTopskip \\EnvTopskip=\\medskipamount \\newskip\\EnvBottomskip \\EnvBottomskip=\\medskipamount \\newskip\\EnvLeftskip \\EnvLeftskip=2\\parindent \\newskip\\EnvRightskip \\EnvRightskip=\\parindent \\newskip\\EnvDelt@skip \\EnvDelt@skip=0pt \\newcount\\@envDepth \\@envDepth=\\z@ \\def\\beginEnv#1{% \\begingroup \\def\\@envname{#1}% \\ifvmode\\def\\@isVmode{T}% \\else\\def\\@isVmode{F}\\vskip 0pt\\fi \\ifnum\\@envDepth=\\@ne\\parindent=\\z@\\fi \\advance\\@envDepth by \\@ne \\EnvDelt@skip=\\baselineskip \\advance\\EnvDelt@skip by-\\normalbaselineskip \\@setenvmargins\\EnvLeftskip\\EnvRightskip \\setenvskip{\\EnvTopskip}% \\vskip\\skip@\\penalty-500 } \\def\\endEnv#1{% \\ifnum\\@envDepth<1 \\emsg{> Tried to close ``#1'' environment, but no environment open!}% \\begingroup \\else \\def\\test{#1}% \\ifx\\test\\@envname\\else \\emsg{> Miss-matched environments!}% \\emsg{> Should be closing ``\\@envname'' instead of ``\\test''}% \\fi \\fi \\vskip 0pt \\setenvskip\\EnvBottomskip \\vskip\\skip@\\penalty-500 \\xdef\\@envtemp{\\@isVmode}% \\endgroup \\if F\\@envtemp\\vskip-\\parskip\\par\\noindent\\fi } \\def\\setenvskip#1{\\skip@=#1 \\divide\\skip@ by \\@envDepth} \\def\\@setenvmargins#1#2{% \\advance \\leftskip by #1 \\advance \\displaywidth by -#1 \\advance \\rightskip by #2 \\advance \\displaywidth by -#2 \\advance \\displayindent by #1}% \\def\\itemize{\\beginEnv{itemize}% \\let\\itm=\\itemizeitem \\vskip-\\parskip } \\def\\itemizeitem{% \\par\\noindent \\hbox to 0pt{\\hss\\itemmark\\space}}% \\def\\enditemize{\\endEnv{itemize}}% \\def\\itemmark{$\\bullet$}% \\newcount\\enumDepth \\enumDepth=\\z@ \\newcount\\enumcnt \\def\\enumerate{\\beginEnv{enumerate}% \\global\\advance\\enumDepth by \\@ne \\setenumlead \\enumcnt=\\z@ \\let\\itm=\\enumerateitem \\if F\\@isVmode\\vskip-\\parskip\\fi } \\def\\enumerateitem{% \\par\\noindent \\advance\\enumcnt by \\@ne \\edef\\lab@l{\\enumlead \\enumcur}% \\hbox to \\z@{\\hss \\lab@l \\enummark \\hskip .5em\\relax}% \\ignorespaces}% \\def\\endenumerate{% \\global\\advance\\enumDepth by -\\@ne \\endEnv{enumerate}}% \\def\\enumPoints{% \\def\\setenumlead{\\ifnum\\enumDepth>1 \\edef\\enumlead{\\enumlead\\enumcur.}% \\else\\def\\enumlead{}\\fi}% \\def\\enumcur{\\number\\enumcnt}% } \\def\\enumpoints{\\enumPoints}% \\def\\enumOutline{% \\def\\setenumlead{\\def\\enumlead{}}% \\def\\enumcur{\\ifcase\\enumDepth \\or\\uppercase{\\XA\\romannumeral\\number\\enumcnt}% \\or\\LetterN{\\the\\enumcnt}% \\or\\XA\\romannumeral\\number\\enumcnt \\or\\letterN{\\the\\enumcnt}% \\or{\\the\\enumcnt}% \\else $\\bullet$\\space\\fi}% } \\def\\enumoutline{\\enumOutline}% \\def\\enumNumOutline{% \\def\\setenumlead{\\def\\enumlead{}}% \\def\\enumcur{\\ifcase\\enumDepth \\or{\\XA\\number\\enumcnt}% \\or\\letterN{\\the\\enumcnt}% \\or{\\XA\\romannumeral\\number\\enumcnt}% \\else $\\bullet$\\space\\fi}% } \\def\\enumnumoutline{\\enumNumOutline}% \\def\\LetterN#1{\\count@=#1 \\advance\\count@ 64 \\XA\\char\\count@} \\def\\letterN#1{\\count@=#1 \\advance\\count@ 96 \\XA\\char\\count@} \\def\\enummark{.}% \\def\\enumlead{}% \\enumpoints \\newbox\\@desbox \\newbox\\@desline \\newdimen\\@glodeswd \\newcount\\@deslines \\newif\\ifsingleline \\singlelinefalse \\def\\description#1{\\beginEnv{description}% \\setbox\\@desbox=\\hbox{#1}% \\@glodeswd=\\wd\\@desbox \\@setenvmargins{\\@glodeswd}{0pt}% \\let\\itm=\\descriptionitem \\if F\\@isVmode\\vskip-\\parskip\\fi }% \\def\\descriptionitem#1{% \\goodbreak\\noindent \\setbox\\@desline=\\vtop\\bgroup \\hfuzz=100cm\\hsize=\\@glodeswd \\rightskip=\\z@ \\leftskip=\\z@ \\raggedright \\noindent{#1}\\par \\global\\@deslines=\\prevgraf \\egroup \\ifsingleline \\ifnum\\@deslines>1 \\@deslineitm{#1}% \\else \\setbox\\@desline=\\hbox{#1}% \\ifdim \\wd\\@desline>\\wd\\@desbox \\@deslineitm{#1}% \\else\\@desitm\\fi \\fi \\else \\@desitm \\fi \\ignorespaces} \\def\\@desitm{% \\noindent \\hbox to \\z@{\\hskip-\\@glodeswd \\hbox to \\@glodeswd{\\vtop to \\z@{\\box\\@desline\\vss}% \\hss}\\hss}}% \\def\\@deslineitm#1{% \\hbox{\\hskip-\\@glodeswd {#1}\\hss}% \\vskip-\\parskip\\nobreak\\noindent } \\def\\enddescription{\\ifhmode\\par\\fi \\@setenvmargins{-\\wd\\@desbox}{0pt}% \\endEnv{description}} \\def\\example{\\beginEnv{example}% \\parskip=\\z@ \\parindent=\\z@ \\baselineskip=\\normalbaselineskip }% \\def\\endexample{\\endEnv{example}% \\noindent}% \\let\\blockquote=\\example \\let\\endblockquote=\\endexample \\def\\Listing{% \\beginEnv{Listing}% \\vskip\\EnvDelt@skip \\baselineskip=\\normalbaselineskip \\parskip=\\z@ \\parindent=\\z@ \\def\\\\##1{\\char92##1}% \\catcode`\\{=\\other \\catcode`\\}=\\other \\catcode`\\(=\\other \\catcode`\\)=\\other \\catcode`\\\"=\\other \\catcode`\\|=\\other \\catcode`\\%=\\other \\catcode`\\&=\\other \\catcode`\\-=\\other \\catcode`\\==\\other \\catcode`\\$=\\other \\catcode`\\#=\\other \\catcode`\\_=\\other \\catcode`\\^=\\other \\catcode`\\~=\\other \\obeylines \\tt\\Listingtabs \\everyListing}% \\def\\endListing{\\endEnv{Listing}}% \\def\\everyListing{\\relax} \\def\\ListCodeFile#1{% \\Listing \\rightskip=\\z@ plus 5cm \\catcode`\\\\=\\other \\input #1\\relax \\endListing} {\\catcode`\\^^I=\\active\\catcode`\\ =\\active \\gdef\\Listingtabs{\\catcode`\\^^I=\\active\\let^^I\\@listingtab \\catcode`\\ =\\active\\let \\@listingspace}% }% \\def\\@listingspace{\\hskip 0.5em\\relax}% \\def\\@listingtab{\\hskip 4em\\relax}% \\def\\TeXexample{\\beginEnv{TeXexample}% \\vskip\\EnvDelt@skip \\parskip=\\z@ \\parindent=\\z@ \\baselineskip=\\normalbaselineskip \\def\\par{\\leavevmode\\endgraf}% \\obeylines \\catcode`|=\\z@ \\ttverbatim \\@eatpar}% \\def\\endTeXexample{% \\vskip 0pt \\endgroup \\endEnv{TeXexample}}% \\def\\ttverbatim{\\begingroup \\catcode`\\(=\\other \\catcode`\\)=\\other \\catcode`\\\"=\\other \\catcode`\\[=\\other \\catcode`\\]=\\other \\catcode`\\~=\\other \\let\\do=\\uncatcode \\dospecials \\obeyspaces\\obeylines \\def\\n{\\vskip\\baselineskip}% \\tt}% \\def\\uncatcode#1{\\catcode`#1=\\other}% {\\obeyspaces\\gdef {\\ }}% \\def\\TeXquoteon{\\catcode`\\|=\\active}% \\let\\TeXquoteson=\\TeXquoteon \\def\\TeXquoteoff{\\catcode`\\|=\\other}% \\let\\TeXquotesoff=\\TeXquoteoff {\\TeXquoteon\\obeylines% \\gdef|{\\ifmmode\\vert\\else% \\ttverbatim\\spaceskip=\\ttglue% \\let^^M=\\ \\relax% \\let|=\\endgroup\\fi}% } \\def\\ttvert{\\hbox{\\tt\\char`\\|}} \\outer\\def\\begintt{$$\\let\\par=\\endgraf \\ttverbatim \\parskip=0pt \\catcode`\\|=0 \\rightskip=-5pc \\ttfinish} {\\catcode`\\|=0 |catcode`|\\=\\other% |obeylines% |gdef|ttfinish#1^^M#2\\endtt{#1|vbox{#2}|endgroup$$}% } \\def\\beginlines{\\par\\begingroup\\nobreak\\medskip\\parindent=0pt \\hrule\\kern1pt\\nobreak \\obeylines \\everypar{\\strut}} \\def\\endlines{\\kern1pt\\hrule\\endgroup\\medbreak\\noindent} \\def\\beginproclaim#1#2#3#4#5{\\medbreak\\vskip-\\parskip \\global\\XA\\advance\\csname #2\\endcsname by \\@ne \\edef\\lab@l{\\@chaptID\\@sectID \\number\\csname #2\\endcsname}% \\tag{#4#5}{\\lab@l}% \\noindent{\\bf #1 \\lab@l.\\space}% \\begingroup #3}% \\def\\endproclaim{% \\par\\endgroup\\ifdim\\lastskip<\\medskipamount \\removelastskip\\penalty55\\medskip\\fi}% \\newcount\\theoremnum \\theoremnum=\\z@ \\def\\theorem#1{\\beginproclaim{Theorem}{theoremnum}{\\sl}{Thm.}{#1}} \\let\\endtheorem=\\endproclaim \\def\\Theorem#1{Theorem~\\use{Thm.#1}} \\newcount\\lemmanum \\lemmanum=\\z@ \\def\\lemma#1{\\beginproclaim{Lemma}{lemmanum}{\\sl}{Lem.}{#1}} \\let\\endlemma=\\endproclaim \\def\\Lemma#1{Lemma~\\use{Lem.#1}} \\newcount\\corollarynum \\corollarynum=\\z@ \\def\\corollary#1{\\beginproclaim{Corollary}{corollarynum}{\\sl}{Cor.}{#1}} \\let\\endcorollary=\\endproclaim \\def\\Corollary#1{Corollary~\\use{Cor.#1}} \\newcount\\definitionnum \\definitionnum=\\z@ \\def\\definition#1{\\beginproclaim{Definition}{definitionnum}{\\rm}{Def.}{#1}} \\let\\enddefinition=\\endproclaim \\def\\Definition#1{Definition~\\use{Def.#1}} \\def\\proof{\\medbreak\\vskip-\\parskip\\noindent{\\it Proof. }} \\def\\blackslug{% \\setbox0\\hbox{(}% \\vrule width.5em height\\ht0 depth\\dp0}% \\def\\QED{\\blackslug}% \\def\\endproof{\\quad\\blackslug\\par\\medskip} \\catcode`@=11 \\def\\paper{% \\auxswitchtrue \\refswitchtrue \\texsis \\def\\titlepage{% \\bgroup \\let\\title=\\Title \\let\\endmode=\\relax \\pageno=1}% \\def\\endtitlepage{% \\endmode \\goodbreak\\bigskip \\egroup}% \\autoparens \\quoteon } \\def\\Tbf{\\fourteenpoint\\bf}% \\def\\tbf{\\twelvepoint\\bf}% \\def\\preprint{% \\auxswitchtrue \\refswitchtrue \\texsis \\def\\titlepage{% \\bgroup \\pageno=1 \\let\\title=\\Title \\let\\endmode=\\relax \\banner}% \\def\\endtitlepage{% \\endmode \\vfil\\eject \\egroup}% \\autoparens \\quoteon } \\def\\Manuscript{% \\preprint \\showsectIDfalse \\showchaptIDfalse \\def{% \\endmode \\bigskip\\bigskip\\medskip \\bgroup\\singlespaced \\let\\endmode=\\endabstract \\narrower\\narrower \\everyabstract}% \\def\\endabstract{% \\medskip\\egroup\\bigskip}% \\def\\FootText{--\\ \\tenrm\\folio\\ --}% \\def\\Tbf{\\sixteenpoint\\bf}% \\def\\tbf{\\fourteenpoint\\bf}% \\twelvepoint \\doublespaced \\autoparens \\quoteon }% \\autoload\\thesis{thesis.txs} \\autoload\\UTthesis{thesis.txs} \\autoload\\YaleThesis{thesis.txs} \\def\\Letter{% \\ContentsSwitchfalse \\refswitchfalse \\auxswitchfalse \\texsis \\singlespaced \\LetterFormat}% \\def\\letter{\\Letter}% \\def\\Memo{% \\ContentsSwitchfalse \\refswitchfalse \\auxswitchfalse \\texsis \\singlespaced \\MemoFormat}% \\def\\memo{\\Memo}% \\def\\Referee{% \\ContentsSwitchfalse \\auxswitchfalse \\refswitchfalse \\texsis \\RefReptFormat}% \\def\\referee{\\Referee}% \\def\\Landscape{% \\texsis \\hsize=9in \\vsize=6.5in \\voffset=.5in \\nopagenumbers \\LandscapeSpecial } \\def\\landscape{\\Landscape}% \\def\\LandscapeSpecial{\\special{papersize=11in,8.5in}} \\def\\slides{% \\quoteon \\autoparens \\ATlock \\pageno=1 \\twentyfourpoint \\doublespaced \\raggedright\\tolerance=2000 \\hyphenpenalty=500 \\raggedbottom \\nopagenumbers \\hoffset=-.25in \\hsize=7.0in \\voffset=-.25in \\vsize=9.0in \\parindent=30pt \\def\\bl{\\vskip\\normalbaselineskip}% \\def\\np{\\vfill\\eject}% \\def\\nospace{\\nulldelimiterspace=0pt \\mathsurround=0pt}% \\def\\big##1{{\\hbox{$\\left##1 \\vbox to2ex{}\\right.\\nospace$}}}% \\def\\Big##1{{\\hbox{$\\left##1 \\vbox to2.5ex{}\\right.\\nospace$}}}% \\def\\bigg##1{{\\hbox{$\\left##1 \\vbox to3ex{}\\right.\\nospace$}}}% \\def\\Bigg##1{{\\hbox{$\\left##1 \\vbox to4ex{}\\right.\\nospace$}}}% } \\autoload\\twinout{twin.txs}% \\def\\twinprint{% \\preprint \\let\\t@tl@=\\title \\def\\title{\\vskip-1.5in\\t@tl@}% \\let\\endt@tlep@ge=\\endtitlepage \\def\\endtitlepage{\\endt@tlep@ge \\twinformat}% } \\def\\twinformat{% \\tenpoint\\doublespaced \\def\\Tbf{\\twelvebf}\\def\\tbf{\\tenbf}% \\headlineoffset=0pt \\twinout}% \\catcode`\\@=11 \\let\\NX=\\noexpand\\let\\XA=\\expandafter \\offparens \\newcount\\tabnum \\tabnum=\\z@ \\newcount\\fignum \\fignum=\\z@ \\newif\\ifRomanTables \\RomanTablesfalse \\newif\\ifCaptionList \\CaptionListfalse \\newif\\ifFigsLast \\FigsLastfalse \\newif\\ifTabsLast \\TabsLastfalse \\def\\FiguresLast{\\FigsLasttrue}\\def\\FiguresNow{\\FigsLastfalse} \\def\\TablesLast{\\TabsLasttrue}\\def\\TablesNow{\\TabsLastfalse} \\long\\def\\figure{\\@figure\\topinsert} \\long\\def\\topfigure{\\@figure\\topinsert}% \\long\\def\\midfigure{\\@figure\\midinsert} \\long\\def\\fullfigure{\\@figure\\pageinsert} \\long\\def\\bottomfigure{\\@figure\\bottominsert} \\long\\def\\heavyfigure{\\@figure\\heavyinsert} \\long\\def\\widefigure{\\@figure\\widetopinsert} \\long\\def\\widetopfigure{\\@figure\\widetopinsert} \\long\\def\\widefullfigure{\\@figure\\widepageinsert} \\def\\FigureName{Figure}% \\def\\TableName{Table}% \\def\\@figure#1#2{% \\vskip 0pt \\begingroup \\def\\CaptionName{\\FigureName}% \\def\\@prefix{Fg.}% \\let\\@count=\\fignum \\let\\@FigInsert=#1\\relax \\def\\@arg{#2}\\ifx\\@arg\\empty\\def\\@ID{}% \\else\\LabelParse #2;;\\endlist\\fi \\ifFigsLast \\emsg{\\CaptionName\\space\\@ID. {#2} [storing in \\jobname.fg]}% \\@fgwrite{\\@comment> \\CaptionName\\space\\@ID.\\space{#2}}% \\@fgwrite{\\string\\@FigureItem{\\CaptionName}{\\@ID}{\\NX#1}}% \\seeCR\\let\\@next=\\@copyfig \\else \\emsg{\\CaptionName\\ \\@ID.\\ {#2}}% \\let\\endfigure=\\@endfigure \\setbox\\@capbox\\vbox to 0pt{}% \\def\\@whereCap{N}% \\let\\@next=\\@findcap \\ifx\\@FigInsert\\midinsert\\goodbreak\\fi \\@FigInsert \\fi \\@next} \\def\\@endfigure{\\relax \\if B\\@whereCap\\relax \\vskip\\normalbaselineskip \\centerline{\\box\\@capbox}% \\fi \\endinsert \\endgroup}% \\def\\endfigure{\\emsg{> \\string\\endfigure before \\string\\figure!}} \\def\\figuresize#1{\\vglue #1}% \\def\\@copyfig#1#2\\endfigure{\\endgroup \\ifx#1\\par\\@fgNXwrite{#2\\@endfigure}\\else\\@fgNXwrite{#1#2\\@endfigure}\\fi} \\def\\@FGinit{\\@FileInit\\fgout=\\jobname.fg[Figures]\\gdef\\@FGinit{\\relax}} \\def\\@fgwrite#1{\\@FGinit\\immediate\\write\\fgout{#1}} \\long\\def\\@fgNXwrite#1{\\@FGinit\\unexpandedwrite\\fgout{#1}} \\def\\PrintFigures{\\ifFigsLast\\@PrintFigures\\fi} \\def\\@PrintFigures{% \\@fgwrite{\\@comment>>> EOF \\jobname.fg <<<}% \\immediate\\closeout\\fgout \\begingroup \\FigsLastfalse \\vbox to 0pt{\\hbox to 0pt{\\ \\hss}\\vss}% \\offparens \\catcode`@=11 \\emsg{[Getting figures from file \\jobname.fg]}% \\Input\\jobname.fg \\relax \\endgroup}% \\def\\@FigureItem#1#2#3{% \\begingroup #3\\relax \\def\\@ID{#2}% \\def\\CaptionName{#1}% \\setbox\\@capbox\\vbox to 0pt{}\\def\\@whereCap{N}% \\@findcap}% \\long\\def\\table{\\@table\\topinsert} \\long\\def\\toptable{\\@table\\topinsert}% \\long\\def\\midtable{\\@table\\midinsert} \\long\\def\\fulltable{\\@table\\pageinsert} \\long\\def\\bottomtable{\\@table\\bottominsert} \\long\\def\\heavytable{\\@table\\heavyinsert} \\long\\def\\widetable{\\@table\\widetopinsert} \\long\\def\\widetoptable{\\@table\\widetopinsert} \\long\\def\\widefulltable{\\@table\\widepageinsert} \\def\\@table#1#2{% \\vskip 0pt \\begingroup \\def\\CaptionName{\\TableName}% \\def\\@prefix{Tb.}% \\let\\@count=\\tabnum \\let\\@FigInsert=#1\\relax \\def\\@arg{#2}\\ifx\\@arg\\empty\\def\\@ID{}% \\else\\ifRomanTables \\global\\advance\\@count by\\@ne \\edef\\@ID{\\uppercase\\expandafter {\\romannumeral\\the\\@count}}% \\tag{\\@prefix#2}{\\@ID}% \\else \\LabelParse #2;;\\endlist\\fi \\fi \\ifTabsLast \\emsg{\\CaptionName\\space\\@ID. {#2} [storing in \\jobname.tb]}% \\@tbwrite{\\@comment> \\CaptionName\\space\\@ID.\\space{#2}}% \\@tbwrite{\\string\\@FigureItem{\\CaptionName}{\\@ID}{\\NX#1}}% \\seeCR\\let\\@next=\\@copytab \\else \\emsg{\\CaptionName\\ \\@ID.\\ {#2}}% \\let\\endtable=\\@endfigure \\setbox\\@capbox\\vbox to 0pt{}% \\def\\@whereCap{N}% \\let\\@next=\\@findcap \\ifx\\@FigInsert\\midinsert\\goodbreak\\fi \\@FigInsert \\fi \\@next} \\def\\endtable{\\emsg{> \\string\\endtable before \\string\\table!}} \\def\\@copytab#1#2\\endtable{\\endgroup \\ifx#1\\par\\@tbNXwrite{#2\\@endfigure}\\else\\@tbNXwrite{#1#2\\@endfigure}\\fi} \\def\\@TBinit{\\@FileInit\\tbout=\\jobname.tb[Tables]\\gdef\\@TBinit{\\relax}} \\def\\@tbwrite#1{\\@TBinit\\immediate\\write\\tbout{#1}} \\long\\def\\@tbNXwrite#1{\\@TBinit\\unexpandedwrite\\tbout{#1}} \\def\\PrintTables{\\ifTabsLast\\@PrintTables\\fi} \\def\\@PrintTables{% \\@tbwrite{\\@comment>>> EOF \\jobname.tb <<<}% \\immediate\\closeout\\tbout \\begingroup \\TabsLastfalse \\catcode`@=11 \\offparens \\emsg{[Getting tables from file \\jobname.tb]}% \\Input\\jobname.tb \\relax \\endgroup}% \\newbox\\@capbox \\newcount\\@caplines \\def\\CaptionName{}% \\def\\@ID{}% \\def\\captionspacing{\\normalbaselines}% \\def\\@inCaption{F}% \\long\\def\\caption#1{% \\def\\lab@l{\\@ID}% \\global\\setbox\\@capbox=\\vbox\\bgroup \\def\\@inCaption{T}% \\captionspacing\\seeCR \\dimen@=20\\parindent \\ifdim\\colwidth>\\dimen@\\narrower\\narrower\\fi \\noindent{\\bf \\linkname{\\@TagName}{\\CaptionName~\\@ID}:\\space}% #1\\relax \\vskip 0pt \\global\\@caplines=\\prevgraf \\egroup \\ifnum\\@caplines=\\@ne \\global\\setbox\\@capbox=\\vbox{\\noindent\\seeCR \\hfil{\\bf \\linkname{\\@TagName}{\\CaptionName~\\@ID}:\\space}% #1\\hfil}\\fi \\if N\\@whereCap\\def\\@whereCap{B}\\fi \\if T\\@whereCap \\centerline{\\box\\@capbox}% \\vskip\\baselineskip\\medskip \\fi}% \\def\\Caption{\\begingroup\\seeCR\\@Caption}% \\long\\def\\@Caption#1\\endCaption{\\endgroup \\ifCaptionList \\incaplist{#1}\\fi \\caption{#1}}% \\def\\endCaption{\\emsg{> \\string\\endCaption\\ called before \\string\\Caption.}} \\long\\def\\@findcap#1{% \\ifx #1\\Caption \\def\\@whereCap{T}\\fi \\ifx #1\\caption \\def\\@whereCap{T}\\fi #1}% \\def\\@whereCap{N}% \\def\\ListCaptions{\\@ListCaps\\caplist=\\jobname.cap[List of Captions] {\\let\\FIGLitem=\\CAPLitem}} \\def\\ListFigureCaptions{% \\@ListCaps\\figlist=\\jobname.fgl[List of Figure Captions] {\\let\\FIGLitem=\\CAPLitem}} \\def\\ListTableCaptions{% \\@ListCaps\\tablelist=\\jobname.tbl[List of Table Captions] {\\let\\FIGLitem=\\CAPLitem}} \\def\\CAPLitem#1#2#3\\@endFIGLitem#4{% \\bigskip \\begingroup \\raggedright\\tolerance=1700 \\hangindent=1.41\\parindent\\hangafter=1 \\noindent #1\\ #2 #3 \\hskip 0pt plus 10pt \\vskip 0pt \\endgroup}% \\def\\infiglist{\\begingroup\\seeCR \\@infiglist\\figlist} \\def\\intablelist{\\begingroup\\seeCR \\@infiglist\\tablelist} \\def\\incaplist{\\begingroup\\seeCR \\@infiglist\\caplist} \\def\\FigListWrite#1#2{% \\ifx#1\\figlist\\relax \\FigListInit\\fi \\ifx#1\\tablelist\\relax \\TabListInit\\fi \\ifx#1\\caplist\\relax \\CapListInit\\fi \\edef\\@line@{{#2}}% \\write#1\\@line@}% \\def\\FigListInit{\\@FileInit\\figlist=\\jobname.fgl[List of Figures]% \\gdef\\FigListInit{\\relax}} \\def\\TabListInit{\\@FileInit\\tablelist=\\jobname.tbl[List of Tables]% \\gdef\\TabListInit{\\relax}} \\def\\CapListInit{\\@FileInit\\caplist=\\jobname.cap[List of Captions]% \\gdef\\CapListInit{\\relax}} \\def\\FigListWriteNX#1#2{\\writeNX#1{#2}} \\def\\@infiglist#1#2{% \\FigListWrite#1{\\@comment > \\CaptionName\\space\\@ID:}% \\FigListWrite#1{\\string\\FIGLitem{\\CaptionName} {\\@ID.\\space}}% \\@copycap#1#2\\endlist \\FigListWrite#1{{\\NX\\folio}}% \\endgroup}% \\def\\@copycap#1#2#3\\endlist{% \\ifx#2\\space\\writeNX#1{#3\\@endFIGLitem}% \\else\\writeNX#1{#2#3\\@endFIGLitem}\\fi} \\def\\ListFigures{\\@ListCaps\\figlist=\\jobname.fgl[List of Figures]{}} \\def\\ListTables{\\@ListCaps\\tablelist=\\jobname.tbl[List of Tables]{}} \\def\\@ListCaps#1=#2[#3]#4{% \\immediate\\closeout#1 \\openin#1=#2 \\relax \\ifeof#1\\closein#1 \\else\\closein#1\\emsg{[Getting #3]}% \\begingroup \\showsectIDtrue \\ATunlock\\quoteoff\\offparens #4 \\input #2 \\relax \\endgroup \\fi} \\long\\def\\FIGLitem#1#2#3\\@endFIGLitem#4{% \\medskip \\begingroup \\raggedright\\tolerance=1700 \\ifx\\TOCmargin\\undefined\\skip0=\\parindent \\else\\skip0=\\TOCmargin\\fi \\advance\\rightskip by \\skip0 \\parfillskip=-\\skip0 \\hangindent=1.41\\parindent\\hangafter=1 \\noindent \\ifshowsectID #1\\ \\fi #2 #3 \\hskip 0pt plus 10pt \\leaddots \\hbox to 2em{\\hss\\linkto{page.#4}{#4}}% \\vskip 0pt \\endgroup} \\def\\Fig#1{\\linkto{Fg.#1}{Fig.~\\use{Fg.#1}}} \\def\\Figs#1{\\linkto{Fg.#1}{Figs.~\\use{Fg.#1}}} \\def\\Fg#1{\\linkto{Fg.#1}{\\use{Fg.#1}}} \\def\\Tab#1{\\linkto{Tb.#1}{Table~\\use{Tb.#1}}} \\def\\Tbl#1{\\linkto{Tb.#1}{Table~\\use{Tb.#1}}} \\def\\Tb#1{\\linkto{Tb.#1}{\\use{Tb.#1}}} \\autoload\\Tablebody{Tablebod.txs}\\autoload\\Tablebodyleft{Tablebod.txs} \\autoload\\tablebody{Tablebod.txs} \\autoload\\epsffile{epsf.tex} \\autoload\\epsfbox{epsf.tex} \\autoload\\epsfxsize{epsf.tex} \\autoload\\epsfysize{epsf.tex} \\autoload\\epsfverbosetrue{epsf.tex}\\autoload\\epsfverbosefalse{epsf.tex} \\obsolete\\topFigure\\figure \\obsolete\\midFigure\\midfigure \\obsolete\\fullFigure\\fullfigure \\obsolete\\TOPFIGURE\\figure \\obsolete\\MIDFIGURE\\midfigure \\obsolete\\FULLFIGURE\\fullfigure \\obsolete\\endFigure\\endfigure \\obsolete\\ENDFIGURE\\endfigure \\obsolete\\topTable\\toptable \\obsolete\\midTable\\midtable \\obsolete\\fullTable\\fulltable \\obsolete\\TOPTABLE\\toptable \\obsolete\\MIDTABLE\\midtable \\obsolete\\FULLTABLE\\fulltable \\obsolete\\endTable\\endtable \\obsolete\\ENDTABLE\\endtable \\def\\FIG{\\@obsolete\\FIG\\Fig\\Fig}% \\def\\TBL{\\@obsolete\\TBL\\Tbl\\Tbl}% \\catcode`@=11 \\catcode`\\|=12 \\catcode`\\&=4 \\newcount\\ncols \\ncols=\\z@ \\newcount\\nrows \\nrows=\\z@ \\newcount\\curcol \\curcol=\\z@ \\let\\currow=\\nrows \\newdimen\\thinsize \\thinsize=0.6pt \\newdimen\\thicksize \\thicksize=1.5pt \\newdimen\\tablewidth \\tablewidth=-\\maxdimen \\newdimen\\parasize \\parasize=4in \\newif\\iftableinfo \\tableinfotrue \\newif\\ifcentertables \\centertablestrue \\def\\centeredtables{\\centertablestrue}% \\def\\noncenteredtables{\\centertablesfalse}% \\def\\nocenteredtables{\\centertablesfalse}% \\let\\plaincr=\\cr \\let\\plainspan=\\span \\let\\plaintab=& \\def\\ampersand{\\char`\\&}% \\let\\lparen=( \\let\\NX=\\noexpand \\def\\ruledtable{\\relax \\@BeginRuledTable \\@RuledTable}% \\def\\@BeginRuledTable{% \\ncols=0\\nrows=0 \\begingroup \\offinterlineskip \\def~{\\phantom{0}}% \\def\\span{\\plainspan\\omit\\relax\\colcount\\plainspan}% \\let\\cr=\\crrule \\let\\CR=\\crthick \\let\\nr=\\crnorule \\let\\|=\\Vb \\def\\hfill{\\hskip0pt plus1fill\\hbox{}}% \\ifx\\tablestrut\\undefined\\relax \\else\\let\\tstrut=\\tablestrut\\fi \\catcode`\\|=13 \\catcode`\\&=13\\relax \\TableActive \\curcol=1 \\ifdim\\tablewidth>-\\maxdimen\\relax \\edef\\@Halign{\\NX\\halign to \\NX\\tablewidth\\NX\\bgroup\\TablePreamble}% \\tabskip=0pt plus 1fil \\else \\edef\\@Halign{\\NX\\halign\\NX\\bgroup\\TablePreamble}% \\tabskip=0pt \\fi \\ifcentertables \\ifhmode\\vskip 0pt\\fi \\line\\bgroup\\hss \\else\\hbox\\bgroup \\fi}% \\long\\def\\@RuledTable#1\\endruledtable{% \\vrule width\\thicksize \\vbox{\\@Halign \\thickrule #1\\killspace \\tstrut \\linecount \\plaincr\\thickrule \\egroup}% \\vrule width\\thicksize \\ifcentertables\\hss\\fi\\egroup \\endgroup \\global\\tablewidth=-\\maxdimen \\iftableinfo \\immediate\\write16{[Nrows=\\the\\nrows, Ncols=\\the\\ncols]}% \\fi}% \\def\\TablePreamble{% \\TableItem{####}% \\plaintab\\plaintab \\TableItem{####}% \\plaincr}% \\def\\@TableItem#1{% \\hfil\\tablespace #1\\killspace \\tablespace\\hfil }% \\def\\@tableright#1{% \\hfil\\tablespace\\relax #1\\killspace \\tablespace\\relax}% \\def\\@tableleft#1{% \\tablespace\\relax #1\\killspace \\tablespace\\hfil}% \\let\\TableItem=\\@TableItem \\def\\RightJustifyTables{\\let\\TableItem=\\@tableright}% \\def\\LeftJustifyTables{\\let\\TableItem=\\@tableleft}% \\def\\NoJustifyTables{\\let\\TableItem=\\@TableItem}% \\def\\LooseTables{\\let\\tablespace=\\quad}% \\def\\TightTables{\\let\\tablespace=\\space}% \\LooseTables \\def\\TrailingSpaces{\\let\\killspace=\\relax}% \\def\\NoTrailingSpaces{\\let\\killspace=\\unskip}% \\TrailingSpaces \\def\\setRuledStrut{% \\dimen@=\\baselineskip \\advance\\dimen@ by-\\normalbaselineskip \\ifdim\\dimen@<.5ex \\dimen@=.5ex\\fi \\setbox0=\\hbox{\\lparen}% \\dimen1=\\dimen@ \\advance\\dimen1 by \\ht0% \\dimen2=\\dimen@ \\advance\\dimen2 by \\dp0% \\def\\tstrut{\\vrule height\\dimen1 depth\\dimen2 width\\z@}% }% \\def\\tstrut{\\vrule height 3.1ex depth 1.2ex width 0pt}% \\def\\bigitem#1{% \\dimen@=\\baselineskip \\advance\\dimen@ by-\\normalbaselineskip \\ifdim\\dimen@<.5ex \\dimen@=.5ex\\fi \\setbox0=\\hbox{#1}% \\dimen1=\\dimen@ \\advance\\dimen1 by \\ht0 \\dimen2=\\dimen@ \\advance\\dimen2 by \\dp0 \\vrule height\\dimen1 depth\\dimen2 width\\z@ \\copy0}% \\def\\vctr#1{\\hfil\\vbox to 0pt{\\vss\\hbox{#1}\\vss}\\hfil}% \\def\\nextcolumn#1{% \\plaintab\\omit#1\\relax\\colcount \\plaintab}% \\def\\tab{% \\nextcolumn{\\relax}}% \\let\\novb=\\tab \\def\\vb{% \\nextcolumn{\\vrule width\\thinsize}}% \\def\\Vb{% \\nextcolumn{\\vrule width\\thicksize}}% \\def\\dbl{% \\nextcolumn{\\vrule width\\thinsize \\hskip 2\\thinsize \\vrule width\\thinsize}}% {\\catcode`\\|=13 \\let|0 \\catcode`\\&=13 \\let&0 \\gdef\\TableActive{\\let|=\\vb \\let&=\\tab}% }% \\def\\crrule{\\killspace \\tstrut \\linecount \\plaincr\\tablerule }% \\def\\crthick{\\killspace \\tstrut \\linecount \\plaincr\\thickrule }% \\def\\crnorule{\\killspace \\tstrut \\linecount \\plaincr }% \\def\\crpart{\\killspace \\linecount \\plaincr}% \\def\\tablerule{\\noalign{\\hrule height\\thinsize depth 0pt}}% \\def\\thickrule{\\noalign{\\hrule height\\thicksize depth 0pt}}% \\def\\cskip{\\omit\\relax}% \\def\\crule{\\omit\\leaders\\hrule height\\thinsize depth0pt\\hfill}% \\def\\Crule{\\omit\\leaders\\hrule height\\thicksize depth0pt\\hfill}% \\def\\linecount{% \\global\\advance\\nrows by1 \\ifnum\\ncols>0 \\ifnum\\curcol=\\ncols\\relax\\else \\immediate\\write16 {\\NX\\ruledtable warning: Ncols=\\the\\curcol\\space for Nrow=\\the\\nrows}% \\fi\\fi \\global\\ncols=\\curcol \\global\\curcol=1}% \\def\\colcount{\\relax \\global\\advance\\curcol by 1\\relax}% \\long\\def\\para#1{% \\vtop{\\hsize=\\parasize \\normalbaselines \\noindent #1\\relax \\vrule width 0pt depth 1.1ex}% }% \\def\\begintable{\\relax \\@BeginRuledTable \\@begintable}% \\long\\def\\@begintable#1\\endtable{% \\@RuledTable#1\\endruledtable}% \\def\\E#1{\\hbox{$\\times 10^{#1}$}} \\def\\square{\\hbox{{$\\sqcup$}\\llap{$\\sqcap$}}}% \\def\\grad{\\nabla}% \\def\\del{\\partial}% \\def\\frac#1#2{{#1\\over#2}} \\def\\smallfrac#1#2{{\\scriptstyle {#1 \\over #2}}} \\def\\half{\\ifinner {\\scriptstyle {1 \\over 2}}% \\else {\\textstyle {1 \\over 2}}\\fi} \\def\\bra#1{\\langle#1\\vert}% \\def\\ket#1{\\vert#1\\/\\rangle}% \\def\\vev#1{\\langle{#1}\\rangle}% \\def\\simge{% \\mathrel{\\rlap{\\raise 0.511ex \\hbox{$>$}}{\\lower 0.511ex \\hbox{$\\sim$}}}} \\def\\simle{% \\mathrel{\\rlap{\\raise 0.511ex \\hbox{$<$}}{\\lower 0.511ex \\hbox{$\\sim$}}}} \\def\\gtsim{\\simge}% \\def\\ltsim{\\simle}% \\def\\therefore{% \\setbox0=\\hbox{$.\\kern.2em.$}\\dimen0=\\wd0 \\mathrel{\\rlap{\\raise.25ex\\hbox to\\dimen0{\\hfil$\\cdotp$\\hfil}}% \\copy0}} \\def\\|{\\ifmmode\\Vert\\else \\char`\\|\\fi} \\def\\sterling{{\\hbox{\\it\\char'44}}} \\def\\degrees{\\hbox{$^\\circ$}}% \\def\\degree{\\degrees}% \\def\\real{\\mathop{\\rm Re}\\nolimits}% \\def\\imag{\\mathop{\\rm Im}\\nolimits}% \\def\\tr{\\mathop{\\rm tr}\\nolimits}% \\def\\Tr{\\mathop{\\rm Tr}\\nolimits}% \\def\\Det{\\mathop{\\rm Det}\\nolimits}% \\def\\mod{\\mathop{\\rm mod}\\nolimits}% \\def\\wrt{\\mathop{\\rm wrt}\\nolimits}% \\def\\diag{\\mathop{\\rm diag}\\nolimits}% \\def\\TeV{{\\rm TeV}}% \\def\\GeV{{\\rm GeV}}% \\def\\MeV{{\\rm MeV}}% \\def\\keV{{\\rm keV}}% \\def\\eV{{\\rm eV}}% \\def\\Ry{{\\rm Ry}}% \\def\\mb{{\\rm mb}}% \\def\\mub{\\hbox{\\rm $\\mu$b}}% \\def\\nb{{\\rm nb}}% \\def\\pb{{\\rm pb}}% \\def\\fb{{\\rm fb}}% \\def\\cmsec{{\\rm cm^{-2}s^{-1}}}% \\def\\units#1{\\hbox{\\rm #1}} \\let\\unit=\\units \\def\\dimensions#1#2{\\hbox{$[\\hbox{\\rm #1}]^{#2}$}} \\def\\parenbar#1{{\\null\\! \\mathop{\\smash#1}\\limits ^{\\hbox{\\fiverm(--)}}% \\!\\null}}% \\def\\nunubar{\\parenbar{\\nu}} \\def\\ppbar{\\parenbar{p}} \\def\\buildchar#1#2#3{{\\null\\! \\mathop{\\vphantom{#1}\\smash#1}\\limits ^{#2}_{#3}% \\!\\null}}% \\def\\overcirc#1{\\buildchar{#1}{\\circ}{}} \\def\\sun{{\\hbox{$\\odot$}}}\\def\\earth{{\\hbox{$\\oplus$}}} \\def\\slashchar#1{\\setbox0=\\hbox{$#1$}% \\dimen0=\\wd0 \\setbox1=\\hbox{/} \\dimen1=\\wd1 \\ifdim\\dimen0>\\dimen1 \\rlap{\\hbox to \\dimen0{\\hfil/\\hfil}}% #1 \\else \\rlap{\\hbox to \\dimen1{\\hfil$#1$\\hfil}}% / \\fi}% \\def\\subrightarrow#1{% \\setbox0=\\hbox{% $\\displaystyle\\mathop{}% \\limits_{#1}$}% \\dimen0=\\wd0 \\advance \\dimen0 by .5em \\mathrel{% \\mathop{\\hbox to \\dimen0{\\rightarrowfill}}% \\limits_{#1}}}% \\newdimen\\vbigd@men \\def\\vbigl{\\mathopen\\vbig} \\def\\vbigm{\\mathrel\\vbig} \\def\\vbigr{\\mathclose\\vbig} \\def\\vbig#1#2{{\\vbigd@men=#2\\divide\\vbigd@men by 2 \\hbox{$\\left#1\\vbox to \\vbigd@men{}\\right.\\n@space$}}} \\def\\Leftcases#1{\\smash{\\vbigl\\{{#1}}} \\def\\Rightcases#1{\\smash{\\vbigr\\}{#1}}} \\def\\doublecolumns{\\relax}\\def\\enddoublecolumns{\\relax} \\def\\leftcolrule{\\relax}\\def\\rightcolrule{\\relax} \\def\\longequation{\\relax}\\def\\endlongequation{\\relax} \\def\\newcolumn{\\relax} \\def\\widetopinsert{\\topinsert}\\def\\widepageinsert{\\pageinsert} \\def\\forceleft{\\relax}\\def\\forceright{\\relax} \\def\\SetDoubleColumns#1{% \\imsg{The double column macros are not a part of mTeXsis.} \\imsg{If you want to use double column mode, get TXSdcol.tex} \\imsg{and add \\string\\input\\space TXSdcol.tex to your .tex file.} } \\def\\addTOC#1#2#3{\\relax}\\def\\Contents{\\relax} % \\newif\\ifContents % \\def\\ContentsSwitchtrue{\\Contentstrue}\\def\\ContentsSwitchfalse{\\Contentsfalse} \\def\\obsolete#1#2{\\let#1=#2\\relax #2} % \\let\\Input=\\input % \\newdimen\\colwidth \\colwidth=\\hsize % \\def\\ORGANIZATION{}% \\newhelp\\@utohelp{% loadstyle: The definition of the macro named above is actually contained^^J% in a style file, and so it cannot be used with mTeXsis. If you really^^J% need to load the definition from that file, you should do so explicitly^^J% at the begining of your manuscript file, with % '\\string\\input\\space stylefilename.txs'^^J} \\Ignore \\def\\loadstyle#1#2{% \\newlinechar=10 % \\errhelp=\\@utohelp % \\emsg{> Whoops! Trying to load \\string#1\\space from style file #2.}% \\errmessage{You cannot use macro definitions from style files in mTeXsis}} \\endIgnore \\hbadness=10000 % \\overfullrule=0pt % \\vbadness=10000 % \\ATunlock \\SetDate % \\ReadAUX % \\def\\fmtname{TeXsis}\\def\\fmtversion{2.17}% \\def\\revdate{1 January 1998}% \\def\\imsg#1{\\emsg{\\@comment #1}}% \\imsg{=========================================================== \\@comment} \\imsg{This is mTeXsis, the core macros from TeXsis.} \\imsg{You can get the complete TeXsis package (and avoid this annoying} \\imsg{advertisement) from ftp://lifshitz.ph.utexas.edu/texsis, } \\imsg{or from a CTAN server near you (in macros/texsis).} \\imsg{See the README and INSTALL files there for more information.} \\imsg{============================================================ \\@comment} \\emsg{m\\fmtname\\space version \\fmtversion\\space (\\revdate) loaded.}% \\ATlock % \\texsis % \\quoteoff \\global\\mathchardef\\DELTA=\"7001 \\global\\mathchardef\\LAMBDA=\"7003 \\global\\mathchardef\\XI=\"7004 \\global\\mathchardef\\SIGMA=\"7006 \\global\\mathchardef\\UPSILON=\"7007 \\global\\mathchardef\\OMEGA=\"700A \\global\\mathchardef\\Delta=\"7101 \\global\\mathchardef\\Lambda=\"7103 \\global\\mathchardef\\Xi=\"7104 \\global\\mathchardef\\Sigma=\"7106 \\global\\mathchardef\\Upsilon=\"7107 \\global\\mathchardef\\Omega=\"710A \\catcode`\\@=11 \\font\\tenmsa=msam10 \\font\\sevenmsa=msam7 \\font\\fivemsa=msam5 \\font\\tenmsb=msbm10 \\font\\sevenmsb=msbm7 \\font\\fivemsb=msbm5 \\newfam\\msafam \\newfam\\msbfam \\textfont\\msafam=\\tenmsa \\scriptfont\\msafam=\\sevenmsa \\scriptscriptfont\\msafam=\\fivemsa \\textfont\\msbfam=\\tenmsb \\scriptfont\\msbfam=\\sevenmsb \\scriptscriptfont\\msbfam=\\fivemsb \\def\\hexnumber@#1{\\ifnum#1<10 \\number#1\\else \\ifnum#1=10 A\\else\\ifnum#1=11 B\\else\\ifnum#1=12 C\\else \\ifnum#1=13 D\\else\\ifnum#1=14 E\\else\\ifnum#1=15 F\\fi\\fi\\fi\\fi\\fi\\fi\\fi} \\def\\msa@{\\hexnumber@\\msafam} \\def\\msb@{\\hexnumber@\\msbfam} \\global\\mathchardef\\boxdot=\"2\\msa@00 \\global\\mathchardef\\boxplus=\"2\\msa@01 \\global\\mathchardef\\boxtimes=\"2\\msa@02 \\global\\mathchardef\\square=\"0\\msa@03 \\global\\mathchardef\\blacksquare=\"0\\msa@04 \\global\\mathchardef\\centerdot=\"2\\msa@05 \\global\\mathchardef\\lozenge=\"0\\msa@06 \\global\\mathchardef\\blacklozenge=\"0\\msa@07 \\global\\mathchardef\\circlearrowright=\"3\\msa@08 \\global\\mathchardef\\circlearrowleft=\"3\\msa@09 \\global\\mathchardef\\rightleftharpoons=\"3\\msa@0A \\global\\mathchardef\\leftrightharpoons=\"3\\msa@0B \\global\\mathchardef\\boxminus=\"2\\msa@0C \\global\\mathchardef\\Vdash=\"3\\msa@0D \\global\\mathchardef\\Vvdash=\"3\\msa@0E \\global\\mathchardef\\vDash=\"3\\msa@0F \\global\\mathchardef\\twoheadrightarrow=\"3\\msa@10 \\global\\mathchardef\\twoheadleftarrow=\"3\\msa@11 \\global\\mathchardef\\leftleftarrows=\"3\\msa@12 \\global\\mathchardef\\rightrightarrows=\"3\\msa@13 \\global\\mathchardef\\upuparrows=\"3\\msa@14 \\global\\mathchardef\\downdownarrows=\"3\\msa@15 \\global\\mathchardef\\upharpoonright=\"3\\msa@16 \\let\\restriction=\\upharpoonright \\global\\mathchardef\\downharpoonright=\"3\\msa@17 \\global\\mathchardef\\upharpoonleft=\"3\\msa@18 \\global\\mathchardef\\downharpoonleft=\"3\\msa@19 \\global\\mathchardef\\rightarrowtail=\"3\\msa@1A \\global\\mathchardef\\leftarrowtail=\"3\\msa@1B \\global\\mathchardef\\leftrightarrows=\"3\\msa@1C \\global\\mathchardef\\rightleftarrows=\"3\\msa@1D \\global\\mathchardef\\Lsh=\"3\\msa@1E \\global\\mathchardef\\Rsh=\"3\\msa@1F \\global\\mathchardef\\rightsquigarrow=\"3\\msa@20 \\global\\mathchardef\\leftrightsquigarrow=\"3\\msa@21 \\global\\mathchardef\\looparrowleft=\"3\\msa@22 \\global\\mathchardef\\looparrowright=\"3\\msa@23 \\global\\mathchardef\\circeq=\"3\\msa@24 \\global\\mathchardef\\succsim=\"3\\msa@25 \\global\\mathchardef\\gtrsim=\"3\\msa@26 \\global\\mathchardef\\gtrapprox=\"3\\msa@27 \\global\\mathchardef\\multimap=\"3\\msa@28 \\global\\mathchardef\\therefore=\"3\\msa@29 \\global\\mathchardef\\because=\"3\\msa@2A \\global\\mathchardef\\doteqdot=\"3\\msa@2B \\let\\Doteq=\\doteqdot \\global\\mathchardef\\triangleq=\"3\\msa@2C \\global\\mathchardef\\precsim=\"3\\msa@2D \\global\\mathchardef\\lesssim=\"3\\msa@2E \\global\\mathchardef\\lessapprox=\"3\\msa@2F \\global\\mathchardef\\eqslantless=\"3\\msa@30 \\global\\mathchardef\\eqslantgtr=\"3\\msa@31 \\global\\mathchardef\\curlyeqprec=\"3\\msa@32 \\global\\mathchardef\\curlyeqsucc=\"3\\msa@33 \\global\\mathchardef\\preccurlyeq=\"3\\msa@34 \\global\\mathchardef\\leqq=\"3\\msa@35 \\global\\mathchardef\\leqslant=\"3\\msa@36 \\global\\mathchardef\\lessgtr=\"3\\msa@37 \\global\\mathchardef\\backprime=\"0\\msa@38 \\global\\mathchardef\\risingdotseq=\"3\\msa@3A \\global\\mathchardef\\fallingdotseq=\"3\\msa@3B \\global\\mathchardef\\succcurlyeq=\"3\\msa@3C \\global\\mathchardef\\geqq=\"3\\msa@3D \\global\\mathchardef\\geqslant=\"3\\msa@3E \\global\\mathchardef\\gtrless=\"3\\msa@3F \\global\\mathchardef\\sqsubset=\"3\\msa@40 \\global\\mathchardef\\sqsupset=\"3\\msa@41 \\global\\mathchardef\\trianglerighteq=\"3\\msa@44 \\global\\mathchardef\\trianglelefteq=\"3\\msa@45 \\global\\mathchardef\\bigstar=\"0\\msa@46 \\global\\mathchardef\\between=\"3\\msa@47 \\global\\mathchardef\\blacktriangledown=\"0\\msa@48 \\global\\mathchardef\\blacktriangleright=\"3\\msa@49 \\global\\mathchardef\\blacktriangleleft=\"3\\msa@4A \\global\\mathchardef\\blacktriangle=\"0\\msa@4E \\global\\mathchardef\\triangledown=\"0\\msa@4F \\global\\mathchardef\\eqcirc=\"3\\msa@50 \\global\\mathchardef\\lesseqgtr=\"3\\msa@51 \\global\\mathchardef\\gtreqless=\"3\\msa@52 \\global\\mathchardef\\lesseqqgtr=\"3\\msa@53 \\global\\mathchardef\\gtreqqless=\"3\\msa@54 \\global\\mathchardef\\Rrightarrow=\"3\\msa@56 \\global\\mathchardef\\Lleftarrow=\"3\\msa@57 \\global\\mathchardef\\veebar=\"2\\msa@59 \\global\\mathchardef\\barwedge=\"2\\msa@5A \\global\\mathchardef\\doublebarwedge=\"2\\msa@5B \\global\\mathchardef\\angle=\"0\\msa@5C \\global\\mathchardef\\measuredangle=\"0\\msa@5D \\global\\mathchardef\\sphericalangle=\"0\\msa@5E \\global\\mathchardef\\varpropto=\"3\\msa@5F \\global\\mathchardef\\smallsmile=\"3\\msa@60 \\global\\mathchardef\\smallfrown=\"3\\msa@61 \\global\\mathchardef\\Subset=\"3\\msa@62 \\global\\mathchardef\\Supset=\"3\\msa@63 \\global\\mathchardef\\Cup=\"2\\msa@64 \\let\\doublecup=\\Cup \\global\\mathchardef\\Cap=\"2\\msa@65 \\let\\doublecap=\\Cap \\global\\mathchardef\\curlywedge=\"2\\msa@66 \\global\\mathchardef\\curlyvee=\"2\\msa@67 \\global\\mathchardef\\leftthreetimes=\"2\\msa@68 \\global\\mathchardef\\rightthreetimes=\"2\\msa@69 \\global\\mathchardef\\subseteqq=\"3\\msa@6A \\global\\mathchardef\\supseteqq=\"3\\msa@6B \\global\\mathchardef\\bumpeq=\"3\\msa@6C \\global\\mathchardef\\Bumpeq=\"3\\msa@6D \\global\\mathchardef\\lll=\"3\\msa@6E \\let\\llless=\\lll \\global\\mathchardef\\ggg=\"3\\msa@6F \\let\\gggtr=\\ggg \\global\\mathchardef\\circledS=\"0\\msa@73 \\global\\mathchardef\\pitchfork=\"3\\msa@74 \\global\\mathchardef\\dotplus=\"2\\msa@75 \\global\\mathchardef\\backsim=\"3\\msa@76 \\global\\mathchardef\\backsimeq=\"3\\msa@77 \\global\\mathchardef\\complement=\"0\\msa@7B \\global\\mathchardef\\intercal=\"2\\msa@7C \\global\\mathchardef\\circledcirc=\"2\\msa@7D \\global\\mathchardef\\circledast=\"2\\msa@7E \\global\\mathchardef\\circleddash=\"2\\msa@7F \\def\\ulcorner{\\delimiter\"4\\msa@70\\msa@70 } \\def\\urcorner{\\delimiter\"5\\msa@71\\msa@71 } \\def\\llcorner{\\delimiter\"4\\msa@78\\msa@78 } \\def\\lrcorner{\\delimiter\"5\\msa@79\\msa@79 } \\def\\yen{\\mathhexbox\\msa@55 } \\def\\checkmark{\\mathhexbox\\msa@58 } \\def\\circledR{\\mathhexbox\\msa@72 } \\def\\maltese{\\mathhexbox\\msa@7A } \\global\\mathchardef\\lvertneqq=\"3\\msb@00 \\global\\mathchardef\\gvertneqq=\"3\\msb@01 \\global\\mathchardef\\nleq=\"3\\msb@02 \\global\\mathchardef\\ngeq=\"3\\msb@03 \\global\\mathchardef\\nless=\"3\\msb@04 \\global\\mathchardef\\ngtr=\"3\\msb@05 \\global\\mathchardef\\nprec=\"3\\msb@06 \\global\\mathchardef\\nsucc=\"3\\msb@07 \\global\\mathchardef\\lneqq=\"3\\msb@08 \\global\\mathchardef\\gneqq=\"3\\msb@09 \\global\\mathchardef\\nleqslant=\"3\\msb@0A \\global\\mathchardef\\ngeqslant=\"3\\msb@0B \\global\\mathchardef\\lneq=\"3\\msb@0C \\global\\mathchardef\\gneq=\"3\\msb@0D \\global\\mathchardef\\npreceq=\"3\\msb@0E \\global\\mathchardef\\nsucceq=\"3\\msb@0F \\global\\mathchardef\\precnsim=\"3\\msb@10 \\global\\mathchardef\\succnsim=\"3\\msb@11 \\global\\mathchardef\\lnsim=\"3\\msb@12 \\global\\mathchardef\\gnsim=\"3\\msb@13 \\global\\mathchardef\\nleqq=\"3\\msb@14 \\global\\mathchardef\\ngeqq=\"3\\msb@15 \\global\\mathchardef\\precneqq=\"3\\msb@16 \\global\\mathchardef\\succneqq=\"3\\msb@17 \\global\\mathchardef\\precnapprox=\"3\\msb@18 \\global\\mathchardef\\succnapprox=\"3\\msb@19 \\global\\mathchardef\\lnapprox=\"3\\msb@1A \\global\\mathchardef\\gnapprox=\"3\\msb@1B \\global\\mathchardef\\nsim=\"3\\msb@1C \\global\\mathchardef\\napprox=\"3\\msb@1D \\global\\mathchardef\\nsubseteqq=\"3\\msb@22 \\global\\mathchardef\\nsupseteqq=\"3\\msb@23 \\global\\mathchardef\\subsetneqq=\"3\\msb@24 \\global\\mathchardef\\supsetneqq=\"3\\msb@25 \\global\\mathchardef\\subsetneq=\"3\\msb@28 \\global\\mathchardef\\supsetneq=\"3\\msb@29 \\global\\mathchardef\\nsubseteq=\"3\\msb@2A \\global\\mathchardef\\nsupseteq=\"3\\msb@2B \\global\\mathchardef\\nparallel=\"3\\msb@2C \\global\\mathchardef\\nmid=\"3\\msb@2D \\global\\mathchardef\\nshortmid=\"3\\msb@2E \\global\\mathchardef\\nshortparallel=\"3\\msb@2F \\global\\mathchardef\\nvdash=\"3\\msb@30 \\global\\mathchardef\\nVdash=\"3\\msb@31 \\global\\mathchardef\\nvDash=\"3\\msb@32 \\global\\mathchardef\\nVDash=\"3\\msb@33 \\global\\mathchardef\\ntrianglerighteq=\"3\\msb@34 \\global\\mathchardef\\ntrianglelefteq=\"3\\msb@35 \\global\\mathchardef\\ntriangleleft=\"3\\msb@36 \\global\\mathchardef\\ntriangleright=\"3\\msb@37 \\global\\mathchardef\\nleftarrow=\"3\\msb@38 \\global\\mathchardef\\nrightarrow=\"3\\msb@39 \\global\\mathchardef\\nLeftarrow=\"3\\msb@3A \\global\\mathchardef\\nRightarrow=\"3\\msb@3B \\global\\mathchardef\\nLeftrightarrow=\"3\\msb@3C \\global\\mathchardef\\nleftrightarrow=\"3\\msb@3D \\global\\mathchardef\\divideontimes=\"2\\msb@3E \\global\\mathchardef\\varnothing=\"0\\msb@3F \\global\\mathchardef\\nexists=\"0\\msb@40 \\global\\mathchardef\\mho=\"0\\msb@66 \\global\\mathchardef\\thorn=\"0\\msb@67 \\global\\mathchardef\\beth=\"0\\msb@69 \\global\\mathchardef\\gimel=\"0\\msb@6A \\global\\mathchardef\\daleth=\"0\\msb@6B \\global\\mathchardef\\lessdot=\"3\\msb@6C \\global\\mathchardef\\gtrdot=\"3\\msb@6D \\global\\mathchardef\\ltimes=\"2\\msb@6E \\global\\mathchardef\\rtimes=\"2\\msb@6F \\global\\mathchardef\\shortmid=\"3\\msb@70 \\global\\mathchardef\\shortparallel=\"3\\msb@71 \\global\\mathchardef\\smallsetminus=\"2\\msb@72 \\global\\mathchardef\\thicksim=\"3\\msb@73 \\global\\mathchardef\\thickapprox=\"3\\msb@74 \\global\\mathchardef\\approxeq=\"3\\msb@75 \\global\\mathchardef\\succapprox=\"3\\msb@76 \\global\\mathchardef\\precapprox=\"3\\msb@77 \\global\\mathchardef\\curvearrowleft=\"3\\msb@78 \\global\\mathchardef\\curvearrowright=\"3\\msb@79 \\global\\mathchardef\\digamma=\"0\\msb@7A \\global\\mathchardef\\varkappa=\"0\\msb@7B \\global\\mathchardef\\hslash=\"0\\msb@7D \\global\\mathchardef\\hbar=\"0\\msb@7E \\global\\mathchardef\\backepsilon=\"3\\msb@7F \\def\\Bbb{\\ifmmode\\let\\next\\Bbb@\\else \\def\\next{\\errmessage{Use \\string\\Bbb\\space only in math mode}}\\fi\\next} \\def\\Bbb@#1{{\\Bbb@@{#1}}} \\def\\Bbb@@#1{\\fam\\msbfam#1} \\catcode`\\@=12 \\ATunlock % \\def\\refFormat{\\catcode`\\^^M=10} \\def\\Refs#1#2{Refs.~\\use{Ref.#1},\\use{Ref.#2}} \\def\\Refsand#1#2{Refs.~\\use{Ref.#1} and~\\use{Ref.#2}} \\def\\Refsrange#1#2{Refs.~\\use{Ref.#1}--\\use{Ref.#2}} \\superrefsfalse % \\def\\Chapter#1{Chapter~\\use{Chap.#1}} \\def ", "introduction": " ", "conclusions": "" }, "0601/astro-ph0601500_arXiv.txt": { "abstract": "{ The dynamics of the explosive burning process is highly sensitive to the flame speed model in numerical simulations of type Ia supernovae. Based upon the hypothesis that the effective flame speed is determined by the unresolved turbulent velocity fluctuations, we employ a new subgrid scale model which includes a localised treatment of the energy transfer through the turbulence cascade in combination with semi-statistical closures for the dissipation and non-local transport of turbulence energy. In addition, subgrid scale buoyancy effects are included. In the limit of negligible energy transfer and transport, the dynamical model reduces to the Sharp-Wheeler relation. According to our findings, the Sharp-Wheeler relation is insuffcient to account for the complicated turbulent dynamics of flames in thermonuclear supernovae. The application of a co-moving grid technique enables us to achieve very high spatial resolution in the burning region. Turbulence is produced mostly at the flame surface and in the interior ash regions. Consequently, there is a pronounced anisotropy in the vicinity of the flame fronts. The localised subgrid scale model predicts significantly enhanced energy generation and less unburnt carbon and oxygen at low velocities compared to earlier simulations. ", "introduction": "For supernovae of type Ia, \\citet{HoyFow60} proposed a thermonuclear runaway initiated in C+O white dwarfs close to the Chandrasekhar limit as the cause of the explosion. Since the original proposal, there has been vivid controversy of how such an explosion might come about and what the exact physical mechanism could be. Today the computational facilities to process three-dimensional large-eddy simulations (LES) of the explosion event are available. Remarkably, these powerful means have not aided in arriving at a consensus yet. The disagreement stems from some crucial questions. Firstly, what is the appropriate flame speed model? Secondly, does the explosion completely proceed as a deflagration, or does a transition to a delayed detonation set in at some point? The deflagration to detonation transition (DDT) proposed by \\citet{Khok91a} and \\citet{WoosWeav94} appears to resolve the drawbacks of the pure deflagration model. In particular, the energy output obtained from simulations with artificial DDT is closer to the observed one, and less carbon and oxygen is left behind \\citep{GamKhok04,GamKhok05,GolNie05}. For the theoretical understanding of thermonuclear supernovae, however, the lack of a convincing explanation for the initiation of the transition is unsatisfactory \\citep{KhokOr97,NieWoos97,Nie99,ZingWoos05}. In the aforementioned numerical models, a DDT is artifically triggered at more or less arbitrary instants of time. As for the flame speed model, the controversy is whether subgrid scale (SGS) turbulence is mostly driven by Rayleigh-Taylor instabilities or dominated by the energy transfer through the turbulence cascade. The former point of view holds that the magnitude of SGS velocity fluctuations $v^{\\prime}$ is basically given by the Sharp-Wheeler relation \\citep{DavTay50,Sharp84} \\begin{equation} \\label{eq:vel_sharp} v_{\\mathrm{RT}}(l)=0.5\\sqrt{l g_{\\mathrm{eff}}} \\end{equation} where $v_{\\mathrm{RT}}(l)$ is the asymptotic rise velocity of a perturbation of size $l$ due to buoyancy. The effective gravity $g_{\\mathrm{eff}}$ is determined by the density contrast at the interface between burned and unburned material. Setting the turbulent flame speed equal to $v_{\\mathrm{RT}}(\\Delta)$, where $\\Delta$ is the resolution of the numerical grid, has been used in some simulations of type Ia supernovae \\citep{GamKhok03,CaldPle03,GamKhok04}. However, simple scaling arguments disfavour this proposition \\citep{NieHille95a,NieKer97}. Assuming that non-linear interactions between turbulent eddies of different size $l$ set up a Kolmogorov spectrum, the root mean square turbulent velocity fluctuations obey the scaling law $v^{\\prime}(l)\\propto l^{1/3}$. Since the Sharp-Wheeler relation implies $v_{\\mathrm{RT}}(l)\\propto l^{1/2}$, we have $v_{\\mathrm{RT}}(l)/v^{\\prime}(l)~\\propto l^{1/6}\\rightarrow 0$ towards decreasing length scales. Consequently, \\citet{NieHille95a} adopted a SGS model based on the dynamical equation for the turbulence energy $k_{\\mathrm{sgs}}$, i.e. the kinetic energy of unresolved velocity fluctuations \\citep{Schumann75}. The major weakness of their approach arises from the fairly tentative closures which were formulated \\emph{ad hoc} for LES of stellar convection \\citep{Clement93}. Various refutations of the scaling argument have been put forward. To begin with, the spectrum of turbulence energy might be different from the Kolmogorov spectrum. However, recent direct numerical simulations support the hypothesis of a Kolmogorov spectrum in buoyancy-driven turbulent combustion \\citep{ZingWoos05}. A more serious concern is that there might be not enough time to reach the state of developed turbulence with a Kolmogorov spectrum in the transient scenario of a supernova explosion. This question is difficult to settle \\emph{a priori}. For this reason, we took an unbiased point of view and accommodated buoyancy effects in the form of an Archimedian force term in the SGS turbulence energy model. In contrast to the previously used SGS turbulence energy model, the new localised model which is thoroughly discussed in paper I neither presumes isotropy nor a certain turbulence energy spectrum function. This is possible by virtue of a dynamical procedure for the determination of the SGS eddy-viscosity $\\nu_{\\mathrm{sgs}}= C_{\\nu}\\Delta k_{\\mathrm{sgs}}^{1/2}$ which was adapted from \\cite{KimMen99}. Furthermore, we apply the co-expanding grid introduced by \\cite{Roep05}. The growth of the cutoff length $\\Delta$ due to the grid expansion poses a challenge for the SGS model because of the partitioning between resolved energy and SGS energy changes in time. We will show that this rescaling effect can be taken into account by utilising the dynamical procedure for the calculation of eddy-viscosity parameter $C_{\\nu}$. The rescaling algorithm as well as the computation of the Archimedian force is explained in Sect.~\\ref{sc:flame_speed_model}, followed by the discussion of results from three-dimension numerical simulations in Sect.~\\ref{sc:num_simul}. It is demonstrated that the newly proposed SGS model substantially alters the predictions of the deflagration model. In particular, we will analyse the significance of SGS buoyancy affects. ", "conclusions": "We applied the SGS turbulence energy model to the large eddy simulation of turbulent deflagration in thermonuclear supernova explosions. The novel features of this model are a localised closure for the rate of energy transfer, an additional Archimedian force term which accounts for buoyancy effects on unresolved scales and the rescaling of the SGS turbulence energy due to the shift of the cutoff length in simulations with a co-expanding grid. We found that the production of turbulence is largely confined to the regions near the flame fronts and in the interior ash regions. Consequently, there is pronounced anisotropy at the flame surface which can be tackled by the localised SGS model only. The Archimedian force contributes noticeably to the turbulent flame speed, particularly once the flame surface has grown substantially. However, the dominating effect is the energy transfer through the turbulence cascade. In the late stage of the explosion, sustained turbulence energy comes from the rescaling, while the major dynamical contribution is SGS dissipation. Furthermore, it appears that numerical grids with more than $N=256^{3}$ cells in one octant are necessary in order to sufficiently resolve the turbulent dynamics in the burning regions and to obtain converged results. An investigation of probability density functions over the flame fronts (see Fig.~\\ref{fg:pdf_speed}) reveals that the Rayleigh-Taylor velocity scale $v_{\\mathrm{RT}}(\\Delta_{\\mathrm{eff}})$ given by the Sharp-Wheeler relation~(\\ref{eq:vel_sharp}) is not negligible compared to the SGS turbulence velocity $q_{\\mathrm{sgs}}$, once the regime of fully turbulent burning has been entered. This reflects the slow decrease of the ratio $v_{\\mathrm{RT}}(\\Delta_{\\mathrm{eff}})/q_{\\mathrm{sgs}}$ with the numerical cutoff scale according to the scaling argument discussed in the introduction. The underlying scaling relations do not necessarily apply to transient and inhomogenous flows as in the supernova explosion scenario. The PDF for $q_{\\mathrm{sgs}}$ shows a considerably wider spread than the sharply peaked PDF for $v_{\\mathrm{RT}}(\\Delta_{\\mathrm{eff}})$. We interpret this observation as a consequence of the additional physics in the localised SGS model, which also encompasses turbulent energy transfer (i.e. interaction between resolved and subgrid scales) and turbulent transport (i.e. non-local interactions among subgrid scales). The relation between the Sharp-Wheeler and the turbulence energy models may be analogous to the relation between the mixing length and Reynolds stress models of convection. The final kinetic energy in the simulation with the highest resolution is about $6\\cdot 10^{50}\\,\\mathrm{erg}$. The produced mass of iron group elements, $0.58M_{\\sun}$, falls within the range deduced from observations of type Ia supernovae \\citep{Leib00}. However, some observed events are substantially more energetic. Regarding the numerical simulations, the explosion energy is very sensitive to the initial conditions and the localised SGS model appears to increase the sensitivity even further. Using initial conditions that are different to the highly artificial centrally ignited flame is in progress. A persistent difficulty is the large amount of left over carbon and oxygen (see Fig.~\\ref{fg:densty_150} and~\\ref{fg:mass_distrb_500}). It is not clear yet to what extent this problem can be solved with the aid of the localised SGS model in simulations with more realistic ignition scenarios. It appears more likely that a different mode of burning is required in the late explosion phase. The currently implemented numerical burning extinction at a density threshold of $10^{7}\\,\\mathrm{g\\,cm^{-3}}$ is mostly arbitrary and should be replaced by a physical criterion motivated by the properties of distributed burning at low densities. This might turn out to be an alternative to the DDT scenario." }, "0601/physics0601198_arXiv.txt": { "abstract": "This paper provides a detailed study of turbulent statistics and scale-by-scale budgets in turbulent Rayleigh-B\\'enard convection. It aims at testing the applicability of \\cite{k41} and \\cite{bolgiano59} theories in the case of turbulent convection and at improving the understanding of the underlying inertial range scalings, for which a general agreement is still lacking. Particular emphasis is laid on anisotropic and inhomogeneous effects, which are often observed in turbulent convection between two differentially heated plates. For this purpose, the SO(3) decomposition of structure functions \\citep*{arad99a} and a method of description of inhomogeneities proposed by \\cite{danaila01} are used to derive inhomogeneous and anisotropic generalizations of Kolmogorov and Yaglom equations applying to Rayleigh-B\\'enard convection, which can be extended easily to other types of anisotropic and/or inhomogeneous flows. The various contributions to these equations are computed in and off the central plane of a convection cell using data produced by a Direct Numerical Simulation of turbulent Boussinesq convection at $\\ray=10^6$ and $\\pr=1$ with aspect ratio $A=5$. The analysis of the isotropic part of the Kolmogorov equation demonstrates that the shape of the third-order velocity structure function is significantly influenced by buoyancy forcing and large-scale inhomogeneities, while the isotropic part of the mixed third-order structure function $\\moy{(\\Delta\\theta)^2\\Delta\\vec{u}}$ appearing in Yaglom equation exhibits a clear scaling exponent 1 in a small range of scales. The magnitudes of the various low $\\ell$ degree anisotropic components of the equations are also estimated and are shown to be comparable to their isotropic counterparts at moderate to large scales. The analysis of anisotropies notably reveals that computing reduced structure functions (structure functions computed at fixed depth for correlation vectors $\\vec{r}$ lying in specific planes only) in order to reveal scaling exponents predicted by isotropic theories is misleading in the case of fully three-dimensional turbulence in the bulk of a convection cell, since such quantities involve linear combinations of different $\\ell$ components which are not negligible in the flow. This observation also indicates that using single points measurements together with the Taylor hypothesis in the particular direction of a mean flow to test the predictions of asymptotic dimensional isotropic theories of turbulence or to calculate intermittency corrections to these theories may lead to significant biases for mildly anisotropic three-dimensional flows. A qualitative analysis is finally used to show that the influence of buoyancy forcing at scales smaller than the Bolgiano scale is likely to remain important up to $\\ray=10^9$, thus preventing Kolmogorov scalings from showing up in convective flows at lower Rayleigh numbers. ", "introduction": "The quest for inertial range scaling laws in turbulent convection has been very active in recent years (\\textit{e.~g.} \\cite{chilla93,benzi94,calzavarini02,verzicco03,ching04}). Their determination is expected to give some important insight into the thermal and mechanical processes at work in the flow. To this end, turbulent thermal convection is investigated in convection cells heated from below using laboratory and numerical experiments, within the framework of the Boussinesq approximation. Such a flow exhibits two essential properties: it is both strongly anisotropic and inhomogeneous. Anisotropy comes from gravity, while inhomogeneity results from the presence of top and bottom horizontal boundaries in convection cells. As a consequence, inertial range scalings of convective turbulence, when they can ever be observed, depend strongly on the vertical coordinate. Accordingly, asymptotic theories of turbulence constructed under the assumptions of homogeneity and isotropy may be partially irrelevant to understand the observed properties of turbulent convection. Neglecting intermittency effects, the classical picture regarding scaling laws in this flow is that Bolgiano-Obukhov turbulence (\\cite{bolgiano59,obukhov59}, hereafter BO59 theory) should be present on correlation lengths $r$ larger than the so-called Bolgiano length. A dimensional estimate of this length \\citep{chilla93} is given by \\begin{equation} \\label{bolglength} L_B=\\frac{\\nuss^{1/2}d}{(\\ray\\pr)^{1/4}}~, \\end{equation} where $d$ is the depth of the convective layer, $\\nuss$ is the Nusselt number, $\\ray$ is the Rayleigh number and $\\pr$ is the Prandtl number. Instead, homogeneous and isotropic Kolmogorov turbulence (\\cite{k41}, hereafter K41 theory) should be observed for $r1$ or $\\pr<1$), thus leading to the same result at all Rayleigh numbers. For the present aspect ratio and Prandtl number, $L_B$ should therefore remain close to 1 at all Rayleigh numbers. On the contrary, the ratio between the dissipation scale and the effective Bolgiano scale is expected to increase with increasing Rayleigh number, which should in principle help to observe K41 scalings if $\\eta\\ll r\\ll L_B$ can be achieved \\citep*{grossmannlvov93}. One should however derive more precise conditions of applicability of K41 in that case. Assuming K41 to be valid for $r\\leq L_B$, the buoyancy term, which equals the $-4/3\\moy{\\varepsilon} r$ term at $L_B$, would scale like $r^{5/3}$ in that range. This approximation looks rather crude for $r$ close to $L_B$, but as the buoyancy term should scale like $r^{9/5}$ above $L_B$ according to BO59, the actual local scaling exponent close to $L_B$ may not be very different from this 5/3 value. The important point here is that this exponent be larger than 1 (which is the exponent of the $-4/3\\moy{\\varepsilon} r$ term). According to these scalings, non-dissipative scales as small as $3\\times10^{-2} L_B$ should be available in order for the buoyancy term to become ten times smaller than the $-4/3\\moy{\\varepsilon} r$ term, which roughly corresponds to the conditions of the present simulation, in which not definite K41 plateau can be observed (figure~\\ref{figS3l=0}). In order to find a range of scales in which the buoyancy term would become at least one hundred times smaller than the $-4/3\\moy{\\varepsilon}r$ term, one should then look for a regime where $\\eta<10^{-3} L_B$. The required Rayleigh number can be evaluated by using $\\eta\\sim \\ray^{-9/28}$, proposed by \\cite{grossmannlohse93}. This relation can be calibrated using the results of the present simulation at $\\ray=10^6$, for which $\\eta=0.016$. One finally finds that $\\ray$ has to exceed several times $10^9$ to obtain a flow with $\\eta< 10^{-3} L_B$. Even in this situation, definite scalings are expected only in the subrange $10\\eta10$days) LMC models imply that the photosphere is disengaged from the HIF at minimum light, similar to the Galactic models, but there are some indications that the photosphere is located near the HIF for the short period ($P<10$ days) LMC models. We also use the updated LMC data to derive empirical PC and AC relations at these phases. Our numerical models are broadly consistent with our theory and the observed data, though we discuss some caveats in the paper. We apply the idea of the HIF-photosphere interaction to explain recent suggestions that the LMC period-luminosity (PL) and PC relations are non-linear with a break at a period close to 10 days. Our empirical LMC PC and PL relations are also found to be non-linear with the $F$-test. Our explanation relies on the properties of the Saha ionization equation, the HIF-photosphere interaction and the way this interaction changes with the phase of pulsation and metallicity to produce the observed changes in the LMC PC and PL relations. ", "introduction": "\\citet{cod47} found that the Galactic Cepheids follow a spectral type that is independent of their pulsational periods at maximum light and gets later as the periods increase at minimum light. \\citet[][hereafter SKM]{sim93} used radiative hydrodynamical models to explain these observational phenomena as being due to the location of the hydrogen ionization front (HIF) relative to the photosphere. Their results agreed very well with Code's observation. SKM further used the Stefan-Boltzmann law applied at the maximum and minimum light, together with the fact that radial variation is small in the optical \\citep{cox80}, to derive: \\begin{eqnarray} \\log T_{max} - \\log T_{min} = {1\\over{10}}(V_{min} - V_{max}), \\end{eqnarray} \\ni where $T_{max/min}$ are the effective temperature at the maximum/minimum light, respectively. If $T_{max}$ is independent of the pulsation period $P$ (in days), then equation (1) predicts there is a relation between the $V$-band amplitude and the temperature (or the colour) at minimum light, and vice versa. In other words, if the period-colour (PC) relation at maximum (or minimum) light is flat, then there is an amplitude-colour (AC) relation at minimum (or maximum) light. Equation (1) has shown to be valid theoretically and observationally for the classical Cepheids and RR Lyrae variables \\citep{kan04,kan05}. For the RR Lyrae variables, \\citet{kan95} and \\citet{kan96} used linear and non-linear hydrodynamic models of RRab stars in the Galaxy to explain why RRab stars follow a flat PC relation at {\\it minimum} light. Later, \\citet{kan05} used MACHO RRab stars in the LMC to prove that LMC RRab stars follow a relation such that higher amplitude stars are driven to cooler temperatures at maximum light. Similar studies were also carried out for Cepheid variables, as in SKM, \\citet{kan96}, \\citet[][hereafter Paper I]{kan04} and \\citet[][hereafter Paper II]{kan04a}. In contrast to the RR Lyrae variables, Cepheids show a flat PC relation at the {\\it maximum} light, and there is a AC relation at the minimum light. Therefore, the PC relation and the AC relation are intimately connected. All these studies are in accord with the predictions of equation (1). In Paper I, the Galactic, Large Magellanic Cloud (LMC) and Small Magellanic Cloud (SMC) Cepheids were analyzed in terms of the PC and AC relations at the phase of maximum, mean and minimum light. One of the motivations for this paper originates from recent studies on the non-linear LMC PC relation \\citep[as well as the period-luminosity, PL, relation. See Paper I;][]{tam02a,san04,nge05}: the optical data are more consistent with two lines of differing slopes which are continuous or almost continuous at a period close to 10 days. Paper I also applied the the $F$-test \\citep{wei80} to the PC and AC relations at maximum, mean and minimum $V$-band light for the Galactic, LMC and SMC Cepheids. The $F$-test results implied that the LMC PC relations are broken or non-linear, in the sense described above, across a period of 10 days, at mean and minimum light, but only marginally so at maximum light. The results for the Galactic and SMC Cepheids are similar, in a sense that at mean and minimum light the PC relations do not show any non-linearity and the PC(max) relation exhibited marginal evidence of non-linearity. For the AC relation, Cepheids in all three galaxies supported the existence of two AC relations at maximum, mean and minimum light. In addition, the Cepheids in these three galaxies also exhibited evidence of the PC-AC connection, as implied by equation (1), which give further evidence of the HIF-photosphere interactions as outlined in SKM. To further investigate the connection between equation (1) and the HIF-photosphere interaction, and also to explain Code's observations with modern stellar pulsation codes, Galactic Cepheid models were constructed in Paper II. In contrast to SKM's purely radiative models, the stellar pulsation codes used in Paper II included the treatment of turbulent convection as outlined in \\citet{yec98}. One of the results from Paper II was that the general forms of the theoretical PC and AC relation matched the observed relations well. The properties of the PC and AC relations for the Galactic Cepheids with $\\log(P)>0.8$ can be explained with the HIF-photosphere interaction. This interaction, to a large extent, is independent of the pulsation codes used, the adopted ML relations, and the detailed input physics. The aim of this paper is to extend the investigation of the connections between PC-AC relations and the HIF-photosphere interactions in theoretical pulsation models of LMC Cepheids, in addition to the Galactic models presented in Paper II. In Section 2, we describe the basic physics of the HIF-photosphere interaction. The updated observational data, after applying various selection criteria, that used in this paper are described in Section 3. In Section 4, the new empirical PC and AC relations based on the data used are presented. In Section 5, we outline our methods and model calculations, and the results are presented in Section 6. Examples of the HIF-photosphere interaction in astrophysical applications are given in Section 7. Our conclusions \\& discussion are presented in Section 8. Throughout the paper, short and long period Cepheid are referred to Cepheids with period less and greater than 10 days, respectively. ", "conclusions": "In this paper, we have confronted updated PC and AC relations at maximum, mean and minimum light for LMC Cepheids observed by the OGLE team, and additional Cepheids from the literature, with theoretical, full amplitude pulsation models of LMC Cepheids. The observed PC and AC relations provide compelling evidence of a non-linearity or break at a period of 10 days. We also constructed theoretical Cepheid pulsation models appropriate for the LMC using the Florida pulsation codes \\citep{yec98} to study the HIF-photosphere interaction. The empirical results presented in this paper, as well as in other papers such as \\citet{san04} and \\citet{nge05}, provide strong empirical evidence that the PC and PL relations for the LMC Cepheids are non-linear, in the sense described in previous sections. Issues such as extinction and a lack of long period Cepheids that may cause the non-linear LMC PL and PC relations have been addressed and argued against in Paper I, \\citet{san04} and \\citet{nge05}, and will not be repeated here. Other arguments against the non-linear LMC PL relation include the results presented in \\citet{per04}, as the authors found no evidence for a non-linear PL relation in the LMC at $JHK$-bands. However, \\citet{nge05} treated the data of \\citet{per04} extensively and found, in a statistically rigorous way, that the reason why \\citet{per04} found linear $JHK$ PL relations, is due to the small number of short period Cepheids ($\\sim18$) in their sample. \\citet{nge05} also reduce the number of OGLE/MACHO LMC Cepheids and show how the $F$-test can produce a non-significant result when the number of short/long period Cepheids become small. Instead, using the 2MASS data that are cross-correlated with MACHO Cepheids, \\citet{nge05} have found that the LMC $JH$-band PL relations are non-linear\\footnote{Non-linear PL relations are relatively easier to see in $V$-band than in $J$-band, as shown in \\citet{nge05}.} and the $K$-band PL relation starts to become linear. \\citet{nge05} also discussed why this is the case. Another argument against the non-linear PL relation is that the PL relation should be universal, as found in \\citet{gie05}. We argue that their results are based on a handful of Cepheids ($\\sim15$) and on short periods Cepheids in a cluster whose membership to the LMC is in question. Their shallower Galactic PL relation based on the revised infra-red surface brightness method also contradicts the steeper Galactic PL relation based on independent methods from open cluster main-sequence fitting \\citep{tam03,nge04,san04}. It is worthwhile to point out that our sample selection does not affect the detection on non-linear LMC PL relation at mean light. Since the mean magnitudes of a Cepheid light curve is less affected by our constrains on selecting the Cepheids with good light curves, we can use the published (reddening corrected) mean magnitudes to test the non-linear LMC PL relation. The anonymous referee kindly provided a large sample of LMC Cepheids that combined the published mean $V$-band magnitudes from the OGLE, \\citet{seb02} and \\citet{cal91} datasets. There are a total of 115 long period Cepheids in this sample and the $F$-test still return a significant detection of the non-linear LMC PL relation. The OGLE+\\citet{seb02} combined data also give very similar results. Similar tests have also been done in \\citet{nge05} by using the MACHO data alone and the MACHO+\\citet{seb02} combined data. The non-linear LMC PL relation is still present from the $F$-test results on these two datasets. Therefore we believe our sample selection does not affect the detection of the non-linear LMC PL relation. The detection of non-linear LMC PL relation from totally independent OGLE and MACHO data, using totally independent reddening estimates, suggested that this non-linearity is real and our paper is the first attempt to theoretically explain this non-linearity in terms of the HIF-photosphere interaction. Due to small number of LMC models, it is impossible to derive the theoretical PC and AC relations with a small error on the slope and compare directly to the empirical relations. However, these LMC models can be qualitatively compared to the observations by converting some physical quantities to the observable quantities and vice versa, such as the temperature-colour conversion. Hence we compared our model light curves to the observations in terms of theoretical PC and AC relations at the phases of maximum, mean and minimum light and also in terms of the Fourier parameters from theoretical light curves with observations. The theoretical quantities from the models generally agree with the observations, but it was found out that these models tend to have smaller amplitudes and (hence) the temperature is cooler at maximum light than the real Cepheids. Though our models have some drawbacks in this comparison, our main interest is in comparing the interaction of the photosphere and HIF as a function of phase with similar results presented in Paper II for Galactic Cepheid pulsation models. The aim is {\\it not} to compare our models rigorously with observations but rather to study models which match observations reasonably well in the context of the theoretical framework described in previous sections and in Paper I \\& II. Nevertheless we argued that the qualitative nature of the photosphere-HIF interaction is not seriously affected by these problems. Our postulate is that at certain phases, this interaction can affect the PC relation due to the properties of the Saha ionization equation: specifically for reasonably low densities in Cepheid envelopes, hydrogen ionizes at a temperature that is almost independent of period. Consequently, when the photosphere is located at the base of the HIF, the photospheric temperature and hence the colour is almost independent of period. However, when this engagement occurs, but the density is greater, then the temperature at which hydrogen ionizes again becomes sensitive to global surroundings and hence on period. When the photosphere is not engaged with the HIF in this way, its temperature is again dependent on period and global stellar parameters. For Galactic Cepheids, this HIF-photosphere interaction occurs mainly at maximum light for Cepheids with $\\log P > 0.8$ (Paper II). At minimum light, there is a strong correlation between the HIF-photosphere distance and period leading to a definite AC relation at minimum light for Galactic Cepheids (SKM, Paper I \\& II). In this paper, we have found tentative evidence that, for short period LMC models which match observations in the period-color plane, the HIF-photosphere interaction occurs at most phases but at densities which are too high to produce a flat PC relation. Why would these short period LMC Cepheids be different in this regard to short period Galactic Cepheids? One possibility could be that this is partly because these LMC Cepheids are hotter than their Galactic counterparts \\citep{kan04,san04}. The HIF-photosphere are disengaged for most of the pulsation cycle for long period LMC Cepheids. This happens because as the period increases, so does the $L/M$ ratio which pushes the HIF further inside the mass distribution. When the HIF-photosphere are disengaged in this way, the photospheric temperature is more dependent on density and hence on period. The change is sudden because the HIF-photosphere are either engaged or they are not. This can lead to a sudden change in the PC relation at 10 days as shown by the observations \\citep{tam02a,kan04,san04,nge05}. However, at maximum light the HIF-photosphere are engaged at low densities for long period LMC Cepheids leading to the observed flat PC relation for these stars. Taken together with equation (1), this theoretical scenario is consistent with the observed PC-AC behavior described in Paper I and in this study. The anonymous referee has noted that these suggestions about photospheric density can be tested by spectroscopic means. We now enumerate some caveats to our argument that could be addressed in future papers. \\begin{enumerate} \\item Since the SMC PC relation at mean light is linear (e.g., Paper I), how do SMC (i.e., metal-poor) models fit into the theoretical scenario outlined in this paper and Paper II, if at all? This is a difficult question and its full answer is beyond the scope of this paper. However, as the metallicity decreases, we do note that the SMC has a different ML relation to the LMC and Galaxy and so does the temperatures associated with the instability strip. These will change the relative location of the HIF and photosphere \\citep{kan95,kan96} and possibly alter the phase at which they interact. Further the amplitudes for SMC Cepheids are smaller due to the lower metallicity \\citep{pac00}. This will also affect the HIF-photosphere interaction. One difference which can be consistent with this is the fact that the PC relation at maximum light in the SMC is not flat (see Paper I) but it is the case for the Galaxy and LMC PC relations. This indicates that at maximum light, there is less interaction between the HIF and photosphere at low densities. This leads to an observed linear PC relation at mean light for the SMC Cepheids. These will be investigated further in a future paper in this series. \\item Could the well-known Hertzsprung progression play any part in causing the observed changes in the Galactic and LMC PC relations? \\item It may also be that higher order overtones becoming unstable or stable, though with the fundamental mode still being dominant, may also have an impact on the PC relation in some as yet unknown way (Paper II). \\item The behavior of short period LMC Cepheids still needs to be understood, for example, what causes the difference between the bottom left panels of Figures \\ref{c9deltalmc} and \\ref{c9delta}? That is, why is it that for short period Galactic/LMC Cepheids, the HIF-photosphere are disengaged/engaged? Our experience suggests that constructing short period full amplitude fundamental mode Cepheids requires more care than the long period case because the first overtone has a non-negligible growth rate. Because of this we feel a thorough study of these short period Cepheids merits a separate paper. \\item Would more advanced pulsation codes which, for example, can match the observed amplitudes and which contain a more accurate model of time dependent turbulent convection, yield similar results, especially for Figure \\ref{c9delta}? Could such codes fare better in modeling short period LMC Cepheids? \\end{enumerate}" }, "0601/astro-ph0601393_arXiv.txt": { "abstract": "We present $K_{\\rm s}$-band surface photometry of NGC 7690 (Hubble type Sab) and NGC 4593 (SBb). We find that, in both galaxies, a major part of the ``bulge'' is as flat as the disk and has approximately the same color as the inner disk. In other words, the ``bulges'' of these galaxies have disk-like properties. We conclude that these are examples of ``pseudobulges'' -- that is, products of secular dynamical evolution. Nonaxisymmetries such as bars and oval disks transport disk gas toward the center. There, star formation builds dense stellar components that look like -- and often are mistaken for -- merger-built bulges but that were constructed slowly out of disk material. These pseudobulges can most easily be recognized when, as in the present galaxies, they retain disk-like properties. NGC 7690 and NGC 4593 therefore contribute to the growing evidence that secular processes help to shape galaxies. NGC 4593 contains a nuclear ring of dust that is morphologically similar to nuclear rings of star formation that are seen in many barred and oval galaxies. The nuclear dust ring is connected to nearly radial dust lanes in the galaxy's bar. Such dust lanes are a signature of gas inflow. We suggest that gas is currently accumulating in the dust ring and hypothesize that the gas ring will starburst in the future. The observations of NGC 4593 therefore suggest that major starburst events that contribute to pseudobulge growth can be episodic. ", "introduction": "\\pretolerance=15000 \\tolerance=15000 Internal secular evolution of galaxies is the dynamical redistribution of energy and angular momentum that causes galaxies to evolve slowly between rapid (collapse-timescale) transformation events that are caused by galaxy mergers. Driving agents include nonaxisymmetries in the gravitational potential such as bars, oval disks, and global spiral structure. Kormendy (1993) and Kormendy \\& Kennicutt (2004, hereafter KK) review the growing evidence that secular processes have shaped the structure of many galaxies. The fundamental way that self-gravitating disks evolve~-- provided that there is an efficient driving agent -- is by spreading (Lynden-Bell \\& Kalnajs 1972; Lynden-Bell \\& Pringle 1974; Tremaine 1989; see Kormendy \\& Fisher 2005 for a review in the present context). In general, it is energetically favorable to shrink the inner parts by expanding the outer parts. In barred galaxies, one well known consequence is the production of ``inner rings'' around the end of the bar and ``outer rings'' at about 2.2 bar radii. The most general consequence of secular evolution, and the one that is of interest in this paper, is that some disk gas is driven to small radii where it reaches high densities, feeds starbursts, and builds ``pseudobulges''. Because of their high stellar densities and steep density gradients, pseudobulges superficially resemble -- and often are mistaken for -- bulges. Following Sandage (1961) and Sandage \\& Bedke (1994), Renzini (1999) adopts this definition of a bulge: ``It appears legitimate to look at bulges as ellipticals that happen to have a prominent disk around them [and] ellipticals as bulges that for some reason have missed the opportunity to acquire or maintain a prominent disk.'' We adopt the same definition. Our paradigm of galaxy formation then is that bulges and ellipticals both formed via galaxy mergers (e.{\\ts}g., Toomre 1977; Steinmetz \\& Navarro 2002, 2003), a conclusion that is well supported by observations (see Schweizer 1990 for a review). Pseudobulges are therefore fundamentally different from bulges -- they were built slowly out of the disk. Two well developed examples, one in the unbarred galaxy NGC 7690 and one in the barred galaxy NGC 4593, are the subjects of this paper. Hierarchical clustering and galaxy merging are well known. Secular evolution is less studied and less well known. We are therefore still in the ``proof of concept'' phase in which it is useful to illustrate clearcut examples of the results of secular evolution. This paper continues a series (see the above reviews and Kormendy \\& Cornell 2004) in which we illustrate the variety of disk-like features that define pseudobulges. ", "conclusions": "NGC 7690 (Sab) and NGC 4593 (SBb) provide clean examples of relatively early-type galaxies whose ``bulges'' are more disk-like than any elliptical galaxy. In particular, elliptical galaxies are never flatter than axial ratio $\\simeq 0.4$ (Sandage et al.~1970; Binney \\& de Vaucouleurs 1981; Tremblay \\& Merritt 1995), whereas part (NGC 7690) or essentially all (NGC 4593) of the bulges of the present galaxies are as flat as their outer disks. We conclude that both galaxies contain pseudobulges -- that is, high-density, central components that were made out of disk gas by secular evolution. In NGC 7690, blue colors imply that star formation and hence pseudobulge growth are still in progress. In NGC 4593, gas appears currently to be accumulating in a ring that plausibly will form stars in the future. Our results are examples of the general conclusion (see KK for a review) that secular dynamical evolution occurs naturally and often in disk galaxies, whether (NGC 4593) or not (NGC 7690) an engine for the evolution is readily recognized. To further investigate secular evolution, a desirable next step would be to quantify bar strengths by measuring~bar~torques, the ratio of the bar-induced, tangential force to the mean radial force as a function of radius (Buta \\& Block 2001; Laurikainen \\& Salo 2002; Block et al.~2001, 2004; Laurikainen et al.~2004). This ratio can be as high as 0.6 in strong bars, emphasizing how efficient bars can be in redistributing angular momentum. It would particularly be worthwhile to look for correlations between maximum bar torques and quantifiable consequences of secular evolution (e.{\\ts}g., ring-to-disk and pseudobulge-to-disk mass ratios) in a statistically representative sample of galaxies." }, "0601/astro-ph0601116_arXiv.txt": { "abstract": "Variable A in M33 is a member of a rare class of highly luminous, evolved stars near the upper luminosity boundary that show sudden and dramatic shifts in apparent temperature due to the formation of optically thick winds in high mass loss episodes. Recent optical and infrared spectroscopy and imaging reveal that its ``eruption'' begun in $\\sim$1950 has ended, {\\it lasting $\\approx$ 45 yrs}. Our current observations show major changes in its wind from a cool, dense envelope to a much warmer state surrounded by low density gas with rare emission lines of Ca II, [Ca II] and K I. Its spectral energy distribution has unexpectedly changed, especially at the long wavelengths, with a significant decrease in its apparent flux, while the star remains optically obscured. We conclude that much of its radiation is now escaping out of our line of sight. We attribute this to the changing structure and distribution of its circumstellar ejecta corresponding to the altered state of its wind as the star recovers from a high mass loss event. ", "introduction": "Variable A in M33 is one of the highly luminous and unstable stars that define the upper luminosity limit in the Hertzsprung-Russell Diagram for evolved cool stars (see Humphreys and Davidson 1994). It was one of the original Hubble -- Sandage variables (Hubble and Sandage 1953) and at its maximum light in 1950 one of the visually brightest stars in M33. Its historical light curve shows a rapid decline from maximum by more than three magnitudes in less than a year followed by a brief recovery and a second decline. It had an intermediate F-type spectrum at maximum consistent with its observed colors. After its second decline to fainter than 18th magnitude (see light curves in Hubble and Sandage 1953 and Rosino and Bianchini 1973), no further observations were obtained until our optical and near infrared photometry beginning in 1977. In 1985-86 Var A had the spectrum of an M-type supergiant (Humphreys, Jones and Gehrz 1987, hereafter HJG), and its large infrared excess at 10$\\mu$m and spectral energy distribution showed that it was still as luminous ($5 \\times 10^{5} L_{\\odot}$) as it was at its maximum light in the visible. Thus its large photometric and spectral variations had occurred at nearly constant bolometric luminosity. We concluded that its cool M-type spectrum was produced in a pseudo-photosphere or optically thick wind formed during a high mass loss episode. A recent spectrum, twenty years later, reveals another dramatic change. Its spectrum is now that of a much warmer star consistent with the warmer photosphere and colors of 50 years ago. In this paper we discuss its remarkable spectral changes reminiscent of the variability of $\\rho$ Cas (Lobel et al. 2003), but on much longer time scales. Variable A was an obvious target for spectroscopy and imaging with the Spitzer Space Telescope. Its resulting spectral energy distribution (.4{\\micron} to 24{\\micron}) also shows unexpected changes, especially at the long wavelengths, most likely corresponding to the changes in its wind and its impact on its circumstellar medium. In this paper we combine all of the available spectroscopy, multi-wavelength photometry and imaging to reconstruct the changes in its wind and circumstellar material over the past twenty years. In the next section we present our multi-wavelength observations. The current spectrum and its light curve and spectral energy distribution are described in \\S 3 and \\S 4. In \\S 5 we discuss Var A's ``eruption'', its duration, and the changing structure of its wind and circumstellar nebula, and in the last section we review its relationship to other cool hypergiants and the possible origins of their instability. ", "conclusions": "Our current observations reveal some remarkable changes in the spectrum and energy distribution of Variable A. Although, we do not have a more complete spectroscopic and photometric record over the past 20 years, our observations do allow us to put together a picture of a very luminous, unstable star experiencing major changes in the structure of its wind and circumstellar material at the end of a high mass loss event. Variable A still had its cool, dense wind and was in a high mass loss phase 35 years after it rapid decline in 1951. Thus it was in ``eruption'' for at least that long. The photometric record also shows that Var A faded two magnitudes between 1954 and 1986 presumably due to the formation of dust, and may have created an additional two magnitudes of circumstellar extinction since then. Our recent spectrum 20 years later shows that the star's F-type photosphere has returned. When did this transition occur? The star has remained faint even though the spectrum has presumably recovered. The variation in its near-infrared photometry may correspond to changes in the wind accompanying the subsidence of its cool, dense wind phase. The 2MASS observations obtained in Dec. 1997 show that Var A had faded significantly in the near-infrared. However, the JHK photometry from 1992 does not show any change from the earlier data in 1986. Assuming that the near-infrared variation is due to the wind in transition, then Var A's eruption or dense wind stage may have lasted between 41 and 46 years! How long the transition back to a warmer temperature may have taken is uncertain, but we note that the initial formation of the dense wind took approximately a year based on its 1950's light curve or at most two years if we include the brief recovery and second decline. Using the record of $\\rho$ Cas, which shows similar outbursts or shell episodes as an example, the onset and recovery timescales are comparable, and in $\\rho$ Cas they occur very rapidly, in only a couple of months. Thus many of the changes we discuss below may have occurred over only one to two years at most for the spectroscopic changes and perhaps up to five years (1992 to 1997) or so for the onset of the changes in the distribution and structure of the circumstellar material. The transition in the wind from an M-type false-photosphere to a warmer F-type star, could reasonably have occurred on even shorter timescales. Although we do not have a direct measurement of Var A's wind speed, the expansion velocities for the winds and ejecta in IRC+10420 and VY CMa are 40 -- 60 km s$^{-1}$ and 35 km s$^{-1}$ for the envelope expansion during $\\rho$ Cas's recent episiode. Assuming 50 km s$^{-1}$ for the wind speed, Var A's envelope could have made these transitions in as short as 3/4's of a year, consistent with its variations during the early 1950's. To explain Var A's observed energy distribution in 1986, HJG proposed a simple model of a cool, dense false-photosphere with an obscuring torus and reflection nebulae at the poles. This model combined extinction by circumstellar dust with the blueing effect of scattering by dust grains and assumed that most of the flux was radiated in the mid-infrared. Most of the visible light was scattered and not viewed through the intervening material. In this model the visible light would very probably have been more highly polarized than the upper limit of 15\\% reported here. There was no optical polarimetry in 1986, so we cannot check this model or verify that there has been a change. Despite its warmer apparent temperature and the corresponding change in the blue-visual energy distribution, Var A has not brightened in the visual, and comparison with its maximum light implies $\\approx$ 4 magnitudes of circumstellar extinction currently in the line of sight. HJG showed that the visual interstellar extinction for Var A was about 0.6 to 0.8 mag which is typical for stars in M33. With its current $\\bv$ color of 0.8 to 0.9 this suggests that its current true color is 0.6 to 0.4, appropriate for a late F-type star, and in comparison with its colors at maximum light ($\\sim$ .4), implies virtually no circumstellar reddening in the visual. As we mentioned previously, the dust grains must be large enough to provide the neutral extinction in the visual which is also consistent with the scattering requirements for the K I emission. Our polarimetry upper limit of 15\\% suggests that the optical light at present is not due mostly to scattered light. If we adopt 30\\% for the polarization of a star completely blocked by circumstellar dust and visible only in reflected light (Johnson and Jones 1991), then at most half the visible light from Var A is from scattered photospheric radiation. The polarimetry also rules out any current asymmetries such as bipolar lobes although reflection nebulosity in a more spherical distribution could still be present. The most dramatic change in Var A's energy distribution is the apparent decrease in its total flux. Since it is highly unlikely that the total luminosity of the star would have declined by more than a magnitude, the energy must now be escaping in some direction other than along our line of sight. Several massive stars are now known to have asymmetrical winds or bipolar outflows. The most notable is $\\eta$ Car with a latitude dependent wind (Smith et al 2002) that is both faster and denser at the poles, although $\\eta$ Car is a very different kind of star. However, given the evidence from other evolved, massive stars including the cool hypergiants like IRC+10420 and VY CMa, for irregularities and density variations in their circumstellar ejecta, it is likely that Var A's dusty shroud is not uniform or completely opaque in all directions. In the remaining discussion of Var A's wind and circumstellar material we will assume that the missing radiation is escaping through large, low density regions, even holes, in the obscuring material, and based on the timescales discussed above for the duration of the eruption and the recovery, we'll assume that Var A has been in this warmer state for approximately ten years. HJG fit Var A's 1986 mid-infrared flux by a 370$\\arcdeg$ BB which implies a dusty zone or shell with a radius of $\\sim$ 400 AU from the star. The energy distribution had a significant near-infrared flux due to warmer dust closer to the star, so the dusty zone probably extended from 100 to 400 AU or so. Figure 4 shows that Var A slowly faded over 30 years due to increased circumstellar obscuration presumably as the amount and density of the obscuring material increased. HJG had estimated a mass loss rate of $2 \\times 10^{-4}$ M$_{\\odot}$ yr$^{-1}$ from Elitzur's (1981) formulation assuming radiation coupling to the grains. But when Var A's optically thick wind ceased and quickly subsided back to a warmer state, then its mass loss rate would very likely have decreased as well. For example, the mass loss rates of the LBV's during their eruptions are typically 10 to 100 times that during their quiescent stage, and the normal mass loss rates of $\\rho$ Cas and HR 8752, two hypergiants of comparable luminosity and temperature, are $\\sim$ $10^{-5}$ M$_{\\odot}$ yr $^{-1}$. In addition to a less dense wind and a lower mass loss rate, Var A may also have a higher wind speed now. Wind speeds of 100 to 200 km s$^{-1}$ are typical of normal A to F-type supergiants. Following HJG's procedure, using Var A's apparent luminosity inferred from its current 10$\\micron$ flux, and assuming that the optical depth is near one in our line of sight, we find a current mass loss rate of $\\sim$ $6.7 \\times 10^{-5}$ M$_{\\odot}$ yr$^{-1}$ for a 50 km s$^{-1}$ wind. But if the wind speed is higher, the mass loss rate will be $\\sim$ 2 -- 3 $\\times 10^{-5}$ M$_{\\odot}$ yr$^{-1}$. In the approximately 10 years since its transition, Var A's lower density and possibly faster wind would have reached 100 to 300 AU, the region of the dusty zone. Consequently, the dusty material may not be replenished as efficiently as in the previous dense wind, higher mass loss stage. So as the dusty zone has continued to expand during this same period, the lower density regions or gaps will also have enlarged allowing more radiation to escape. If the dusty distribution is flattened, as HJG suggested, this is most likely to occur at the poles. Therefore we propose that large dusty condensations in a flattened distribution, possibly a torus, currently obscure our direct view of Var A. The K I emission lines are produced by resonance scattering in this dusty zone. In addition, a very low density gas or wind responsible for the hydrogen, Ca II, and peculiar [Ca II] emission fills the region between the star's photosphere and the dusty zone. More than half of the radiation is escaping from our line of sight, presumably from low density regions, all or most of which must be out of our line of sight. Furthermore, the polarimetry measurements rule out a strongly bipolar structure with asymmetrically scattered light. This then suggests a unique geometry even if the gaps are restricted to the polar regions. The dusty zone must either be nearly aligned with our line of sight in such a way to also block most of the escaping radiation from being reflected back into our line of sight or else there is very little reflecting nebulosity." }, "0601/astro-ph0601320_arXiv.txt": { "abstract": "N-Body simulations are a very important tool in the study of formation of large scale structures. Much of the progress in understanding the physics of galaxy formation and comparison with observations would not have been possible without N-Body simulations. Given the importance of this tool, it is essential to understand its limitations as ignoring these can easily lead to interesting but unreliable results. In this paper we study the limitations due to the finite size of the simulation volume. We explicitly construct the correction term arising due to a finite box size and study its generic features for clustering of matter and also on mass functions. We show that the correction to mass function is maximum near the scale of non-linearity, as a corollary we show that the correction to the number density of haloes of a given mass changes sign at this scale; the number of haloes at small masses is over estimated in simulations. This over estimate results from a delay in mergers that lead to formation of more massive haloes. The same technique can be used to study corrections to other physical quantities. The corrections are typically small if the scale of non-linearity is much smaller than the box-size. However, there are some cases of physical interest in which the relative correction term is of order unity even though a simulation box much larger than the scale of non-linearity is used. Within the context of the concordance model, our analysis suggests that it is very difficult for present day simulations to resolve mass scales smaller than $10^2$~M$_\\odot$ accurately and the level of difficulty increases as we go to even smaller masses, though this constraint does not apply to multi-scale simulations. ", "introduction": "Large scale structures like galaxies and clusters of galaxies are believed to have formed by gravitational amplification of small perturbations. For an overview and original references, see, e.g., \\citet{1980lssu.book.....P,1999coph.book.....P,2002tagc.book.....P,2002PhR...367....1B}. Initial density perturbations were present at all scales that have been observed \\citep{2003ApJS..148..175S,2003MNRAS.346...78H,2004ApJ...607..655P}. Understanding evolution of density perturbations for such initial conditions is essential for the study of formation of galaxies and large scale structures. The equations that describe the evolution of density perturbations in an expanding universe have been known for a long time \\citep{1974A&A....32..391P} and these are easy to solve when the amplitude of perturbations is small. These equations describe the evolution of density contrast defined as $\\delta(\\mathbf r, t) = (\\rho(\\mathbf r, t) - \\bar\\rho(t))/\\bar\\rho(t)$. Here $\\rho(\\mathbf r, t)$ is the density at point $\\mathbf r$ and time $t$, and $\\bar\\rho$ is the average density in the universe at time $t$. These are densities of non-relativistic matter, the component that clusters at all scales and is believed to drive the formation of large scale structures in the universe. Once density contrast at relevant scales becomes large, i.e., $|\\delta| \\geq 1$, the perturbation becomes non-linear and coupling with perturbations at other scales cannot be ignored. The equation for evolution of density perturbations cannot be solved for generic perturbations in this regime. N-Body simulations \\citep{1998ARA&A..36..599B,1997Prama..49..161B,2005CSci...88.1088B} are often used to study the evolution in this regime. Alternative approaches can be used if one requires only a limited amount of information and in such a case either quasi-linear approximation schemes \\citep{1970A&A.....5...84Z,1989MNRAS.236..385G,1992MNRAS.259..437M,1993ApJ...418..570B,1994MNRAS.266..227B,1995PhR...262....1S,1996ApJ...471....1H,2002PhR...367....1B} or scaling relations \\citep{1977ApJS...34..425D,1991ApJ...374L...1H,1995MNRAS.276L..25J,2000ApJ...531...17Ka,1998ApJ...508L...5M,1994MNRAS.271..976N,1996ApJ...466..604P,1994MNRAS.267.1020P,1996MNRAS.278L..29P,1996MNRAS.280L..19P,2003MNRAS.341.1311S} suffice. In cosmological N-Body simulations, we simulate a representative region of the universe. This is a large but finite volume and periodic boundary conditions are often used. Almost always, the simulation volume is taken to be a cube. Effect of perturbations at scales smaller than the mass resolution of the simulation, and of perturbations at scales larger than the box is ignored. Indeed, even perturbations at scales comparable to the box are under sampled. It has been shown that perturbations at small scales do not influence collapse of perturbations at much larger scales in a significant manner \\citep{1974A&A....32..391P,1985ApJ...297..350P,1991MNRAS.253..295L,1997MNRAS.286.1023B,1998ApJ...497..499C} if we study the evolution of the correlation function or power spectrum at large scales due to gravitational clustering in an expanding universe. This is certainly true if the scales of interest are in the non-linear regime \\citep{1997MNRAS.286.1023B}. Therefore we may assume that ignoring perturbations at scales much smaller than the scales of interest does not affect results of N-Body simulations. However, there may be other effects that are not completely understood at the quantitative level \\citep{2004astro.ph..8429B} even though these have been seen only in somewhat artificial situations. Perturbations at scales larger than the simulation volume can affect the results of N-Body simulations. Use of periodic boundary conditions implies that the average density in the simulation box is same as the average density in the universe, in other words we ignore perturbations at the scale of the simulation volume (and at larger scales). Therefore the size of the simulation volume should be chosen so that the amplitude of fluctuations at the box scale (and at larger scales) is ignorable. If the amplitude of perturbations at larger scales is not ignorable then clearly the simulation will not be a faithful representation of the model being studied. It is not obvious as to when fluctuations at larger scales can be considered ignorable, indeed the answer to this question depends on the physical quantity of interest, the model being studied and the specific length/mass scale of interest as well. The effect of a finite box size has been studied using N-Body simulations and the conclusions in this regard may be summarised as follows. \\begin{itemize} \\item If the amplitude of density perturbations around the box scale is small ($\\delta < 1$) but not much smaller than unity, simulations underestimate the correlation function though the number density of small mass haloes does not change by much \\citep{1994ApJ...436..467G,1994ApJ...436..491G}. In other words, the formation of small haloes is not disturbed but their distribution is affected by non-inclusion of long wave modes. \\item In the same situation, the number density of massive haloes drops significantly \\citep{1994ApJ...436..467G,1994ApJ...436..491G,2005MNRAS.358.1076B}. \\item Effects of a finite box size modify values of physical quantities like the correlation function even at scales much smaller than the simulation volume \\citep{2005MNRAS.358.1076B}. \\item The void spectrum is also affected by finite size of the simulation volume if perturbations at large scales are not ignorable \\citep{1992ApJ...393..415K}. \\item It has been shown that properties of a given halo can change significantly as the contribution of perturbations at large scales is removed to the initial conditions but the distribution of most internal properties remain unchanged \\citep{2005astro.ph.12281P}. \\end{itemize} In some cases, one may be able to devise a method to ``correct'' for the effects of a finite box-size \\citep{1994A&A...281..301C}, but such methods cannot be generalised to all statistical measures or physical quantities. The effects of perturbations at scales larger than the box size can be added using MAP (Mode Adding Procedure) after a simulation has been run \\citep{1996ApJ...472...14T}. This method makes use of the fact that if the box size is chosen to be large enough then the contribution of larger scales can be incorporated by adding displacements due to the larger scales independently of the evolution of the system in an N-Body simulation. The motivation for development of such a tool is to enhance the range of scales over which results of an N-Body simulation can be used by improving the description at scales comparable to the box size. Such an approach ignores the coupling of large scale modes with small scale modes and this again brings up the issue of what is a large enough scale for a given model such that the effects of mode coupling can be ignored. Large scales contribute to displacements and velocities, and variations in density due to these scales modify the rate of growth for small scales perturbations \\citep{1997MNRAS.286...38C}. Effects of a finite box size modify values of physical quantities even at scales much smaller than the simulation volume \\citep{2005MNRAS.358.1076B} (BR05, hereafter). In BR05, we suggested use of the fraction of mass in collapsed haloes as an indicator of the effect of a finite box size. We found that if the simulation volume is not large enough, the fraction of mass in collapsed haloes is underestimated. As the collapsed fraction is less sensitive to box-size as compared to measures of clustering, several other statistical indicators of clustering to deviate significantly from expected values in such simulations. A workaround for this problem was suggested in the form of an ensemble of simulations to take the effect of convergence due to long wave modes into account \\citep{2005ApJ...634..728S}, the effects of shear due to long wave modes are ignored here. It has also been shown that the distribution of most internal properties of haloes, e.g., concentration, triaxiality and angular momentum do not change considerably with the box size even though properties of a given halo may change by a significant amount \\citep{2005astro.ph.12281P}. There is a clear need to develop a formalism for estimating the effect of perturbations at large scales on a variety of physical quantities. Without such a formalism, we cannot decide in an objective manner whether a simulation box size is sufficiently large or not. In this paper we generalise the approach suggested in BR05 and write down an explicit correction term for a number of statistical indicators of clustering. This approach allows us to study generic properties of the expected correction term in any given case, apart of course from allowing us to evaluate the magnitude of the correction as compared to the expected value of the given statistical indicator. We apply this technique to mass functions in this paper. ", "conclusions": "Conclusions of this work may be summarised as follows. \\begin{itemize} \\item We have presented a formalism that can be used to estimate the deviations of cosmological N-Body simulations from the models being simulated due to the use of a finite box size. These deviations/errors are independent of the specific method used for doing simulations. \\item For a given model, the deviations can be expressed as a function of the scale $r$ of interest and $\\lbx$, the box size of simulations. \\item We have applied the formalism to study deviations in {\\it rms}\\/ fluctuations in mass in the initial conditions. \\item We find that the errors are small except for models where the slope of the power spectrum is close to $-3$ at scales of interest. \\item The errors in case of the $\\Lambda$CDM model are significant if the scale of interest is smaller than a kpc even if simulations as large as the Millennium simulation \\citep{2005Natur.435..629S} are used. \\item We have studied errors in mass function in the Press-Schechter model, as well as other models. \\item The main error due to a finite box size is that the number of high mass haloes is under estimated. \\item The number of low mass haloes is over predicted in simulations if the box size effects are important. This happens as low mass haloes do not merge to form the (missing) high mass haloes. \\item We have verified these trends using N-Body simulations. \\end{itemize} We note that it is extremely important to understand the sources of errors in N-Body simulations and the magnitude of errors in quantities of physical interest. N-Body simulations are used to make predictions for a number of observational projects and also serve as a test bed for methods. In this era of ``precision cosmology'', it will be tragic if simulations prove to be a weak link. We would like to note that our results apply equally to all methods of doing cosmological N-Body simulations, save those where techniques like MAP are used to include the effects of scales larger than the simulation volume. The method for estimating errors due to a finite box-size described in this paper can be used for several physical quantities. In this paper we have used the method to study errors in clustering properties and mass functions. We are studying the effect of finite box size on velocity fields and related quantities, the results will be presented in a later publication." }, "0601/astro-ph0601099_arXiv.txt": { "abstract": "We revisit the constraint on the primordial helium mass fraction $Y_p$ from observations of cosmic microwave background (CMB) alone. By minimizing $\\chi^2$ of recent CMB experiments over 6 other cosmological parameters, we obtained rather weak constraints as $0.17\\le Y_p \\le 0.52$ at 1$\\sigma$ C.L. for a particular data set. We also study the future constraint on cosmological parameters when we take account of the prediction of the standard big bang nucleosynthesis (BBN) theory as a prior on the helium mass fraction where $Y_p$ can be fixed for a given energy density of baryon. We discuss the implications of the prediction of the standard BBN on the analysis of CMB. \\vspace{1cm} ", "introduction": "\\setcounter{equation}{0} Recent precise cosmological observations such as WMAP \\cite{Bennett:2003bz} push us toward the era of so-called precision cosmology. In particular, the combination of the data from cosmic microwave background (CMB), large scale structure, type Ia supernovae and so on can severely constrain the cosmological parameters such as the energy density of baryon, cold dark matter and dark energy, the equation of state for dark energy, the Hubble parameter, the amplitude and the scale dependence of primordial fluctuation. Among the various cosmological parameters, the primordial helium mass fraction $Y_p$ is the one which has been mainly discussed in the context of big bang nucleosynthesis (BBN) but not that of CMB so far. One of the reason is that the primordial helium abundance has not been considered to be well constrained by observations of CMB since its effects on the CMB power spectrum is expected to be too small to be measured. However, since now we have very precise measurements of CMB, we may have a chance to constrain the primordial helium mass fraction from CMB observations. Since the primordial helium mass fraction can affect the number density of free electron in the course of the recombination history, the effects of $Y_p$ can be imprinted on the power spectrum of CMB. Recently, some works along this line have been done by two different groups \\cite{Trotta:2003xg,Huey:2003ef}, which have discussed the constraints on $Y_p$ from current observations of CMB. In fact they claim different bounds on the primordial helium mass fraction, especially in terms of its uncertainty: the author of Ref.~\\cite{Trotta:2003xg} obtained $ 0.160 \\le Y_p \\le 0.501$, on the other hand the authors of Ref.~\\cite{Huey:2003ef} got $ Y_p =0.250^{+0.010}_{-0.014}$ at 1$\\sigma$ confidence level. It should be noticed that the latter bound is much more severe than that of the former. If the helium mass fraction is severely constrained by CMB data, it means that the CMB power spectrum is sensitive to the values of $Y_p$. In such a case, the prior on $Y_p$ should be important to constrain other cosmological parameters too and the usual fixing of $Y_p=0.24$ in CMB power spectrum calculations might not be a good assumption. Especially, analyses like Refs.~\\cite{Cyburt:2003fe,Cuoco:2003cu,Coc:2003ce,Cyburt:2004cq,Serpico:2004gx} predict light element abundances including $^4$He from the baryon density which is obtained from the CMB data sets with the analysis fixing the value of $Y_p$. Such procedure is only valid when $Y_p$ is not severely constrained by CMB. Thus it is very important to check the CMB bound on $Y_p$. One of the main purpose of the present paper is that we revisit the constraint on $Y_p$ from observations of CMB alone with a different analysis method from Markov chain Monte Carlo (MCMC) technique which is widely used for the determination of cosmological parameters and adopted in Refs.~\\cite{Trotta:2003xg,Huey:2003ef}. In this paper, we calculate $\\chi^2$ minimum as a function of $Y_p$ and derive constraints on $Y_p$. We adopt the Brent method of the successive parabolic interpolation to minimize $\\chi^2$ varying 6 other cosmological parameters of the $\\Lambda$CDM model with the power-law adiabatic primordial fluctuation. We obtain the constraint on $Y_p$ by this method and compare it with previously obtained results. We also study the constraint on $Y_p$ from future CMB experiment. A particular emphasis is placed on investigating the role of the standard BBN theory. Since the primordial helium is synthesized in BBN, once the baryon-to-photon ratio is given, the value of $Y_p$ is fixed theoretically. Thus, using this relation between the baryon density and helium abundance, we do not have to regard $Y_p$ as an independent free parameter when we analyze CMB data. We study how the standard BBN assumption on $Y_p$ affects the determination of other cosmological parameters in the future Planck experiment using the Fisher matrix analysis. The structure of this paper is as follows. In the next section, we briefly discuss the effects of the helium mass fraction on the CMB power spectrum, in particular its effects on the change of the structure of the acoustic peaks. Then we study the constraint on the primordial helium mass fraction from current observations of CMB using the data from WMAP, CBI, ACBAR and BOOMERANG. In section 4, we discuss the expected constraint on $Y_p$ from future CMB observation of Planck and also study how the standard BBN assumption on $Y_p$ can affect the constraints on cosmological parameters. The final section is devoted to the summary of this paper. ", "conclusions": "We revisited the constraint on the primordial helium mass fraction $Y_p$ from current observations of CMB. Some authors have already studied the constraint \\cite{Trotta:2003xg,Huey:2003ef}, however their results were different especially in terms of the uncertainty. One of the main purpose of the present paper is to study the constraints on $Y_p$ from current observations adopting a different analysis method. Instead of MCMC method which was adopted by the authors of Refs.~\\cite{Trotta:2003xg,Huey:2003ef} to obtain the constraint, here we adopted a $\\chi^2$ minimization by a nested grid search. We did not obtain a severe constraint in agreement with Ref.~\\cite{Trotta:2003xg}. Using the data from WMAP, CBI and ACBAR as well as recent BOOMERANG data, we get 1$\\sigma$ constraint as $ 0.25 \\le Y_p \\le 0.54$ and $0.17 \\le Y_p \\le 0.52$ for the cases with and without the data from BOOMERANG. It might be interesting to note that usual assumption of $Y_p=0.24$ is not in the 1$\\sigma$ error range of BOOMERANG combined analysis but it is not of high significance at this stage so we can safely assume $Y_p=0.24$ for current CMB data analysis. We also studied the future constraint from CMB on $Y_p$ taking account of the standard BBN prediction as a prior on $Y_p$. Although we cannot obtain a severe constraint at present, observations of CMB can be much more precise in the future. Thus we may have a chance to obtain a precise measurement of $Y_p$ from upcoming CMB experiments. On the other hand, since the primordial helium has been formed during the time of BBN, once the baryon-to-photon ratio is given, the value of $Y_p$ can be evaluated theoretically assuming the standard BBN. Thus, in this case, we do not have to assume $Y_p$ as an independent free parameter when we analyze CMB data. We studied how such BBN theory prior on $Y_p$ affects the determination of other cosmological parameters in the future Planck experiment. We evaluated the uncertainties for the case with $Y_p$ being an independent free parameter and $Y_p$ being fixed for a given $\\omega_b$ using the BBN relation. We showed that the BBN prior improves the constraints on other cosmological parameters by a factor of $\\mathcal{O}(1)$ and also it induces some correlations among the parameters which appear in the BBN relation. As shown in Fig.~\\ref{fig:future}, as far as we consider the standard scenario of cosmology, the helium mass fraction can be fixed for CMB analysis even in the future experiments since we can expect that the constraint from Planck is much weaker than the uncertainty of the theoretical calculation of the standard BBN. However, it is worthwhile to do CMB analysis treating $Y_p$ as a free parameter and measure the helium mass fraction independently from the baryon density since it provides a consistency test for the standard BBN theory (of course, measurements of primordial light element abundances by astrophysical means provide further consistency tests). By checking the robustness of the consistency from various observations, the golden age of precision cosmology can push us toward the accurate understanding of the universe. \\bigskip \\noindent {\\bf Acknowledgment:} We acknowledge the use of CMBFAST \\cite{Seljak:1996is} package for our numerical calculations. The work of T.T. is supported by Grand-in-Aid for JSPS fellows." }, "0601/astro-ph0601266_arXiv.txt": { "abstract": "We investigate the possibility that part of the dark matter is not made out of the usual cold dark matter (CDM) dustlike particles, but is under the form of a fluid of strings with barotropic factor $w_s = -1/3$ of cosmic origin. To this aim, we split the dark matter density parameter in two terms and investigate the dynamics of a spatially flat universe filled with baryons, CDM, fluid of strings and dark energy, modeling this latter as a cosmological constant or a negative pressure fluid with a constant equation of state $w < 0$. To test the viability of the models and to constrain their parameters, we use the Type Ia Supernovae Hubble diagram and the data on the gas mass fraction in galaxy clusters. We also discuss the weak field limit of a model comprising a significant fraction of dark matter in the form of a fluid of strings and show that this mechanism makes it possible to reduce the need for the elusive and up to now undetected CDM. We finally find that a model comprising both a cosmological constant and a fluid of strings fits very well the data and eliminates the need of phantom dark energy thus representing a viable candidate to alleviate some of the problems plaguing the dark side of the universe. ", "introduction": "Soon after the discovery of cosmic acceleration from the Hubble diagram of the high redshift Type Ia Supernovae (SNeIa) \\cite{SNeIa,Riess04}, a strong debate arose in the scientific community about the origin of this unexpected result. An impressive flow of theoretical proposals have appeared, while the observational results were constantly providing more and more evidences substantiating the emergence of a new cosmological scenario. The anisotropy spectrum of the cosmic microwave background radiation (CMBR) \\cite{CMBR,WMAP,VSA}, the matter power spectrum determined from the clustering properties of the large scale distribution of galaxies \\cite{LSS} and the data on the Ly$\\alpha$ emitting regions \\cite{Lyalpha} all provide indications that the universe have to be described as a spatially flat manifold where matter and its fluctuations are isotropically distributed and represent only about $30\\%$ of the overall content. In order to fill the gap and drive the acceleration, a dominant contribute from a homogeneously distributed negative pressure fluid has been invoked. Usually referred to as {\\it dark energy}, the nature and the nurture of this mysteryous component represent a new and fascinating conundrum for theoreticians. While the models proposed to explain this puzzle increase day by day, the most simple answer is still the old Einstein cosmological constant $\\Lambda$. Although being the best fit to a wide set of different astrophysical observations \\cite{Teg03,Sel04}, it is nevertheless plagued by two evident shortcomings, namely the {\\it cosmological constant problem} and the {\\it coincidence problem}. A possible way to overcome these problems invokes replacing $\\Lambda$ with a scalar field (dubbed {\\it quintessence}) evolving down a suitably chosen self interaction potential \\cite{QuintFirst}. Although solving the cosmological constant problem, quintessence does not eliminate the coincidence one since too severe constraints on the potential seem to be needed thus leading to the {\\it fine tuning problem} \\cite{QuintRev}. The ignorance of the fundamental physical properties of both dark energy and dark matter has motivated a completely different approach to the problem of cosmic acceleration relying on modification of the matter equation of state (EoS). Referred to as {\\it unified dark energy} (UDE) models, these proposals resort to a single fluid with exotic EoS as the only candidate to both dark matter and dark energy thus automatically solving the coincidence problem. The EoS is then tuned such that the fluid behaves as dark matter at high energy density and quintessence (or $\\Lambda$) at the low energy limit. Interesting examples are the Chaplygin gas \\cite{Chaplygin}, the tachyonic field \\cite{tachyon} and the Hobbit model \\cite{Hobbit}. It is worth noting that observations only tell us that the universe is accelerating, but they are not direct evidences for new fluids or modifications of the usual matter properties. It is indeed possible to consider cosmic acceleration as the first signal of the breakdown of the laws of physics as we know them. As a consequence, one has to to give off the standard Friedman equations in favour of a generalized version of them arising from some more fundamental theory. Interesting examples of this kind are the Cardassian expansion \\cite{Cardassian} and the Dvali\\,-\\,Gabadadze\\,-\\,Porrati (DGP) gravity \\cite{DGP} both related to higher dimensional braneworld theories. In the same framework, one should also give off the Einsteinian general relativity and turn to fourth order theories of gravitation replacing the Ricci scalar curvature $R$ in the gravity Lagrangian with a generic function $f(R)$ that have been formulated both in the metric \\cite{capozcurv,MetricRn,cct} and Palatini approach \\cite{PalRn,lnR,ABF04,ACCF} providing a good fit to the data in both cases \\cite{noiijmpd,CCF04}. Actually, it is worth noting that it has been recently demonstrated that, under quite general conditions, it is possible to find a $f(R)$ theory that predicts the same dynamics of a given quintessence model. Although resorting to modified gravity theories is an interesting and fascinating approach, it is worth exploring other possibilities in the framework of standard general relativity. Indeed, all the approaches we have described are mainly interested in solving the dark energy puzzle, while little is said about the dark matter problem. It is worth remembering that dark matter is usually invoked because of the need of a source of gravitational potential other than the visible matter. Considered from this point of view, it is worth wondering whether dark matter could be replaced by a different mechanism that is able to give the same global effect. Moreover, such a mechanism must not alter the delicate balance between dark matter and dark energy that is needed to explain observations. Indeed, if we abruptly reduce the dark matter content of the universe without altering neither the dark energy term nor the background fundamental properties, we are not able to fit the available astrophysical data. Therefore, it is mandatory to test any proposed mechanism both at galactic and cosmological scales. In a series of interesting papers \\cite{letelier}, Letelier investigated the consequences of changing the properties of the right hand side of the Einstein equations adopting a fluid of strings as source term rather than the usual dust matter. Since such strings are not observed at the present time, it seems meaningful to extend the concept of dust clouds and perfect fluid referred to point particles to the case of strings. In particular, Letelier was able to find exact solutions for the case of a spherically symmetric fluid of strings. It is worth noting that such strings could be of cosmological origin \\cite{Vil84} and have thus to be included in the energy budget when investigating the dynamics of the universe. It is important to stress, however, that the strings we are referring to have {\\it finite lenght} so that the results obtained for a network of cosmic strings of infinite length cannot be extended to the strings considered by Letelier. In this sense, the fluid of finite length strings we are considering represents a generalization of the dust matter. While in this latter case, the matter particles are considered as pointlike, in the case of a fluid of strings\\footnote{Hereafter, by {\\it fluid of strings} we mean a {\\it fluid of finite length strings}.} the elementary constituents are one dimensional objects with finite length. A fluid of strings has a profound impact at galactic scales. Indeed, assuming that the string transverse pressure was proportional to its energy density, Soleng \\cite{soleng} has demonstrated that the force law is altered thus offering the possibility of solving the problem of the flatness of spiral galaxy rotation curves \\cite{rc} in a way similar to the MOND proposal \\cite{mond}. Motivated by these considerations, we explore here the possibility that a part (if not all) of the dark matter may be replaced by a fluid of strings whose effective gravitational action may be considered as the source of the gravitational potential needed to flatten the rotation curves. To this aim, we consider cosmic strings as components of such fluid so that its e.o.s. may be simply parametrized by a constant barotropic factor $w_s = -1/3$. Before discussing the impact at galaxy scales, it is preliminarily needed to investigate the effects at cosmological scales. We thus fit different cosmological models, both with and without such a component, to the available astrophysical data in order to test the viability of our proposal and explore if and how the constraints on the model parameters are affected by the presence of a fluid of strings. The paper is organized as follows. The models we discuss are described in Sect.\\,2, while the matching with observations is presented in Sect.\\,3 where we also compare the different models in terms of the information criteria parameters. Sect.\\,4 is devoted to the weak energy limit of models comprising standard matter embedded in a fluid of strings and show how the corresponding modified gravitational potential could help in reducing the need for CDM. A summary of the results and of their implications are presented in the concluding Sect.\\,5 ", "conclusions": "Shedding light on the dark side of the universe is a very difficult, but also very attractive challenge of modern cosmology. The nature and the fundamental properties of the two main ingredients of the cosmic pie, namely the dark energy and the dark matter, are still substantially unknown and it is, indeed, this wide ignorance that justifies and motivates the impressive amount of theoretical models proposed to explain the observed astrophysical evidences. Moving in this framework, we have considered the dark matter as made out not only of massive dustlike CDM particles, but also of a fluid of strings of cosmic origin with an equation of state $w_s = -1/3$. Starting from this idea, we have considered four cosmological models comprising four components, namely dustlike baryons and CDM, fluid of strings and dark energy with constant barotropic factor $w$. Two of these four models ($\\Lambda$SDM and QSDM) have a non vanishing fraction of dark matter in the form of a fluid of strings, while in two models ($\\Lambda$CDM and $\\Lambda$SDM) the energy budget is dominated by the $\\Lambda$ term. Our main results are briefly outlined as follows. \\begin{enumerate} \\item{All the models are able to fit the data on the SNeIa Hubble diagram and the gas mass fraction in galaxy clusters with very good accuracy. In particular, it is remarkable that the total dark matter density parameter $\\Omega_{DM} = \\Omega_{CDM} + \\Omega_s$ is very well constrained and turns out to be the same in all models. When the assumption $w = -1$ is relaxed, the dark energy barotropic factor is constrained to be in the region $w < -1$ so that phantom like models are clearly preferred with a disturbing violation of the strong energy condition. It is worth noting that present data do not require phantom dark energy since they can be equally well fit by models with the cosmological constant $\\Lambda$ driving the accelerated expansion. Discriminating among the different possibilities will need a large sample of high redshift SNeIa such as those that should be available with the SNAP satellite mission \\cite{SNAP}.} \\item{According to both the AIC and BIC, the QCDM model is statistically preferred over the other considered possibilities and this is not an unexpected result. However, this is obtained to the price of admitting phantom dark energy which is affected by serious theoretical difficulties. On the other hand, the $\\Lambda$SDM model is preferred over the popular concordance $\\Lambda$CDM scenario and is only slightly disfavoured with respect to the QCDM one. The good fit to the data and the graceful feature of avoiding to enter the realm of ghosts makes this model a good compromise between observations and theory and we therefore consider it as our final best choice.} \\item{The $\\Lambda$SDM model predicts that a significant fraction ($\\varepsilon \\simeq 59\\%$) of the dark matter is made out by a fluid of strings so that the standard matter density parameter $\\Omega_b + \\Omega_{CDM}$ is only half of the fiducial value in the concordance scenario (0.15 vs 0.30). Since we may assume that the fluid of strings is massless (or nearly so), we should expect a corresponding decrease of the mass of galactic dark haloes. If the gravitational potential is still Newtonian, decreasing the CDM halo mass should lead to lower values of the circular velocity in the outer dark matter dominated regions of galaxies. This is not the case since, in the weak field limit, the $\\Lambda$SDM model gives rise to a modification of the gravitational potential. As a result, the circular velocity due to a mass $M(r)$ is higher than in the classical case so that less massive haloes are necessary to give the observed values of $v_c(r)$. Moreover, a very qualitative calculation suggests that the typical value of the scalelength over which deviations from Newtonian formulae cannot be neglected is sufficiently high that the inner luminous matter dominated rotation curve is unaltered.} \\end{enumerate} These encouraging results motivate further studies of the $\\Lambda$SDM model. To this end, there are two different routes connected to two different features of the model which can be followed. First, because of its scaling with the redshift as $\\rho_s \\propto (1 + z)^2$, that is intermediate between that of CDM and that of $\\Lambda$, a new era dominated by a fluid of strings is predicted in the expansion history of the universe. It is thus worth investigating how this imprints on the CMBR anisotropies in order to see whether the spectrum measured by WMAP is still accurately reproduced. To this regard, it is worth noting that the attempts recently made to constrain the cosmic strings contribution to the CMBR spectrum \\cite{Fraisse} may not be extended to our case since they refer to a network of cosmic strings rather than clouds of finite length strings. Less theoretically demanding, but more observationally ambitious is the possibility to test the proposed scenario on the basis of the transition redshift $z_T$. As Table 1 shows, for the $\\Lambda$SDM model, $z_T$ is significantly higher than in the other cases so that a model independent estimate of this quantity could be a powerful discriminating tool. One of the most peculiar features of the $\\Lambda$SDM model is the modified gravitational potential in Eq.(\\ref{eq: phiext}) leading to the corrected circular velocity in Eq.(\\ref{eq: vcstrings}). Having been obtained in the weak field limit, such correction should be tested at the scale of galaxies and clusters of galaxies thus offering the possibility to test the model at a very different level. To this aim, one should try fitting the rotation curve of spiral galaxies to see whether the problem of their flatness could be solved in this framework. Moreover, it is interesting to check how much the halo mass is reduced and to compare the reduction with respect to the classical Newtonian estimates with the decreasing of $\\Omega_{CDM}$ obtained above. To this aim, low surface brightness (LSB) galaxies are ideal candidates since they are likely dark matter dominated so that systematic uncertainties on the luminous matter modelling have only a minor impact on the fitting procedure. Moreover, the stellar mass\\,-\\,to\\,-\\,light ratio of LSB galaxies is well constrained so that we may fix this quantity thus decreasing the degeneracy among the other parameters. Useful samples of LSB galaxies with detailed measurements of the rotation curve are yet available (see, for instance, \\cite{dBB02}) so that this kind of test may be easily implemented. In this same framework, it is also interesting to consider the velocity dispersion curves in elliptical galaxies where recent studies seem to indicate a dark matter deficit \\cite{Cap}. Changing the gravitational potential does not only alter galaxies rotation curve, but also affects the clustering properties and thus leads to a different matter power spectrum. It is thus interesting to compare the predicted power spectrum with those measured from the SDSS galaxies in order to check the validity of the $\\Lambda$SDM model. A similar comparison has been recently performed by Shirata et al. \\cite{SSYS04} for two phenomenological modifications of the law of gravity. We stress, however, that their approach is purely empirical and, furthermore, assumes that the universe can still be described at large scales with the $\\Lambda$CDM model. Since in order to compute the power spectrum, one also needs the background Hubble parameter, it is important to use an expression for $H(z)$ that is consistent with the proposed modification of gravity. For the $\\Lambda$SDM model considered here, all the ingredients are at disposal so that a coherent calculation can be performed. It is worth noting that such a test is the only one capable of probing the model both at the galactic (through the gravitational potential) and cosmological (because of the use of $H$) scales at the same time. As a concluding general remark, we would like to stress the need for tackling the dark matter and dark energy problem together taking care of what is the effect at the galaxy scale of any modification of the fundamental properties of one of these two components. In our opionion, this could be a valid approach in elucidating the problems connected to the dark side of the universe." }, "0601/hep-ph0601014_arXiv.txt": { "abstract": "#1{\\vskip 7mm \\begin{center}{\\large Abstract}\\par \\smallskip \\begin{minipage}[c]{12cm} \\small #1 \\end{minipage} \\end{center} } \\def\\title#1{\\begin{center}{\\Large\\bf #1}\\end{center}} \\def\\author#1{\\vskip 5mm \\begin{center}{#1}\\end{center}} \\def\\address#1{\\begin{center}{\\it #1}\\end{center}} \\makeatletter \\@ifundefined{lesssim}{\\def\\lesssim{\\mathrel{\\mathpalette\\vereq<}}}{} \\@ifundefined{gtrsim}{\\def\\gtrsim{\\mathrel{\\mathpalette\\vereq>}}}{} \\def\\vereq#1#2{\\lower3pt\\vbox{\\baselineskip1.5pt \\lineskip1.5pt \\ialign{$\\m@th#1\\hfill##\\hfil$\\crcr#2\\crcr\\sim\\crcr}}} \\makeatother \\begin{document} \\title{% PBH and DM from cosmic necklaces } \\author{% Tomohiro Matsuda\\footnote{E-mail:matsuda@sit.ac.jp} } \\address{% $^1$Theoretical Physics Group, Saitama Institute of Technology, Saitama 369-0293, Japan } \\abstract{Cosmic strings in the brane Universe have recently gained a great interest. I think the most interesting story is that future cosmological observations distinguish them from the conventional cosmic strings. If the strings are the higher-dimensional objects that can (at least initially) move along the compactified space, and finally settle down to (quasi-)degenerated vacua in the compactified space, then kinks should appear on the strings, which interpolate between the degenerated vacua. These kinks look like ``beads'' on the strings, which means that the strings turn into necklaces. Moreover, in the case that the compact manifold is not simply connected, the string loop that winds around a non-trivial circle is stable due to the topological reason. Since the existence of degenerated vacua and a non-trivial circle is the common feature of the brane models, it is important to study cosmological constraints on the cosmic necklaces and their stable winding states in the brane Universe.} ", "introduction": "In this talk we will explain the cosmological consequences of the production of Dark Matter(DM) and Primordial Black Hole(PBH) from the loops of the cosmic necklaces. To begin with, I think it is fair to explain why necklaces\\cite{vilenkin_book} are produced in brane models, since in many papers it is discussed that ``only strings are produced in the brane Universe''\\cite{tutorial}. Of course I think this claim is not wrong, however somewhat misleading for non-specialists. To explain what is misleading in the ``standard scenario'', we have a figure in Fig.\\ref{fig:fig0}. In general, the distance between branes may appear in the four-dimensional effective action as a Higgs field of the effective gauge dynamics. At least in this case, it is natural to consider the cosmological defects coming from the spatial variation of the Higgs field, which corresponds to the ``deformation'' of the branes\\cite{matsuda_deformation}. Is the spatial variation of the Higgs field unnatural in the brane Universe? The answer is, of course, no. One should therefore consider at least two different kinds of defects in brane models: One is induced by the brane creation that is due to the spatial variation of the tachyon condensation, while the other is induced by the brane deformation that is due to the spatial variation of the brane distance. Along the line of the above arguments, it is possible to construct Q-ball's counterpart in brane models\\cite{matsuda_Q-balls}, which can be distinguished from the conventional Q-balls by their decay process. We therefore have an expectation that strings can be distinguished from the conventional ones, if one properly considers their characteristic features. Now let us discuss about the validity of the conventional Kibble mechanism. Of course the Kibble mechanism is an excellent idea that explains the nature of the cosmological defect formation. However, if there is oscillation of the brane distance that may be induced by the brane inflation or by a later phase transition that changes the brane distance, the four-dimensional counterpart of the brane distance (i.e. the Higgs field) oscillates in the effective action. In the four-dimensional counterpart, defect production induced by such oscillation is already discussed by many authors, including the production of the sphaleron domain walls which otherwise cannot be produced in the Universe\\cite{matsuda_deformation}. The defect production induced by such oscillation may or may not be explained by the Kibble mechanism, however it should be fair to distinguish it from the ``conventional'' Kibble mechanism. Let us summarize the above discussion about the defect production in the brane Universe. Actually, it is possible to produce all kinds of defects in the brane Universe, however it is impossible to produce defects other than the strings {\\bf simply as the result of the brane creation that is induced by the conventional Kibble mechanism}. One should therefore be careful about the assumption that is made in the manuscript, which may or may not be explicit. The necklaces are produced as the hybrid of the brane creation and the brane deformation. It should be noted that the stable loops of the necklaces that we will discuss in this talk may appear in the four-dimensional gauge dynamics, irrespective of the existence of the branes\\cite{050906x}. The stabilization of the necklace loops is first discussed in Ref.\\cite{matsuda_necklace} for brane models and in Ref.\\cite{050906x} for necklaces embedded in four-dimensional gauge dynamics. In order to produce necklaces in the brane Universe, the motion in the compactified direction is important. I know that in the ``standard scenario'' it is sometimes discussed that the position of the strings are fixed by the potential that is induced by the supersymmetry breaking, and the position is a homogeneous parameter of the Universe because all the decay products (Typically, they are F, D, and $(p,q)$ strings) lie (at least initially) along the same plane of the original hypersurface on which the tachyon condensation took place. However, in this case one may hit upon the idea that the potential for a string cannot be identical to all the other kinds of the strings. One may therefore obtain many kinds of strings that may move independently along different hypersurfaces, with exponentially small intersection ratios. Moreover, I think it is not reasonable(but may be possible) to assume that the string motion is utterly restricted by the potential even in the most energetic epoch just after inflation. Please remember that in general the moving (inflating) brane carries kinetic energy, and the brane annihilation should be an energetic process, although one may admit that there could be exceptional scenarios. I therefore think that the decay products should have kinetic energy, which is enough to climb up the potential hill at least just after brane inflation. \\begin{figure}[h] \\begin{picture}(640,330)(0,0) \\resizebox{16cm}{!}{\\includegraphics{JGRGFig.eps}} \\end{picture} \\caption{We show how necklaces are produced despite the ``standard'' arguments. } \\label{fig:fig0} \\end{figure} ", "conclusions": "" }, "0601/gr-qc0601085_arXiv.txt": { "abstract": "Quantum gravity is expected to be necessary in order to understand situations where classical general relativity breaks down. In particular in cosmology one has to deal with initial singularities, i.e.\\ the fact that the backward evolution of a classical space-time inevitably comes to an end after a finite amount of proper time. This presents a breakdown of the classical picture and requires an extended theory for a meaningful description. Since small length scales and high curvatures are involved, quantum effects must play a role. Not only the singularity itself but also the surrounding space-time is then modified. One particular realization is loop quantum cosmology, an application of loop quantum gravity to homogeneous systems, which removes classical singularities. Its implications can be studied at different levels. Main effects are introduced into effective classical equations which allow to avoid interpretational problems of quantum theory. They give rise to new kinds of early universe phenomenology with applications to inflation and cyclic models. To resolve classical singularities and to understand the structure of geometry around them, the quantum description is necessary. Classical evolution is then replaced by a difference equation for a wave function which allows to extend space-time beyond classical singularities. One main question is how these homogeneous scenarios are related to full loop quantum gravity, which can be dealt with at the level of distributional symmetric states. Finally, the new structure of space-time arising in loop quantum gravity and its application to cosmology sheds new light on more general issues such as time. ", "introduction": "\\label{section:introduction} \\begin{flushright} \\begin{quote} {\\em Die Grenzen meiner Sprache bedeuten die Grenzen meiner Welt.} ({\\em The limits of my language mean the limits of my world.}) \\end{quote} {\\sc Ludwig Wittgenstein} Tractatus logico-philosophicus \\end{flushright} While general relativity is very successful in describing the gravitational interaction and the structure of space and time on large scales \\cite{Will}, quantum gravity is needed for the small-scale behavior. This is usually relevant when curvature, or in physical terms energy densities and tidal forces, becomes large. In cosmology this is the case close to the big bang, and also in the interior of black holes. We are thus able to learn about gravity on small scales by looking at the early history of the universe. Starting with general relativity on large scales and evolving backward in time, the universe becomes smaller and smaller and quantum effects eventually become important. That the classical theory by itself cannot be sufficient to describe the history in a well-defined way is illustrated by singularity theorems \\cite{SingTheo} which also apply in this case: After a finite time of backward evolution the classical universe will collapse into a single point and energy densities diverge. At this point, the theory breaks down and cannot be used to determine what is happening there. Quantum gravity, with its different dynamics on small scales, is expected to solve this problem. The quantum description does not only present a modified dynamical behavior on small scales but also a new conceptual setting. Rather than dealing with a classical space-time manifold, we now have evolution equations for the wave function of a universe. This opens a vast number of problems on various levels from mathematical physics to cosmological observations, and even philosophy. This review is intended to give an overview and summary of the current status of those problems, in particular in the new framework of loop quantum cosmology. ", "conclusions": "" }, "0601/astro-ph0601521_arXiv.txt": { "abstract": "We report the parallax and proper motion of millisecond pulsar J0030+0451, one of thirteen known isolated millisecond pulsars in the disk of the Galaxy. We obtained more than 6 years of monthly data from the 305 m Arecibo telescope at 430 MHz and 1410 MHz. We measure the parallax of PSR J0030+0451 to be $3.3\\pm0.9$\\,mas, corresponding to a distance of $300\\pm90$\\,pc. The Cordes and Lazio (2002) model of galactic electron distribution yields a dispersion measure derived distance of 317 pc which agrees with our measurement. We place the pulsar's transverse space velocity in the range of 8 to 17 km~s$^{-1}$, making this pulsar one of the slowest known. We perform a brief census of velocities of isolated versus binary millisecond pulsars. We find the velocities of the two populations are indistinguishable. However, the scale height of the binary population is twice that of the isolated population and the luminosity functions of the two populations are different. We suggest that the scale height difference may be an artifact of the luminosity difference. ", "introduction": "A pulsar parallax can be combined with a measurement of the pulsar's dispersive delay (the Dispersion Measure or DM) to provide an accurate measure of the free electron density along the line of sight (LOS). PSR J0030+0451 is one of fewer than a dozen pulsars to have its parallax measured via timing \\citep{Kaspi94, Camilo94_1713, Sandhu97, Toscano99, Wolszczan00, Jacoby03, Hotan04, Loehmer04, Splaver05}. Another dozen have been measured via VLBI (e.g. see \\nocite{Brisken02, Chatterjee04} Brisken et al. 2002, Chatterjee et al. 2004, and also this URL\\footnote{ http://www.astro.cornell.edu/$\\sim$shami/psrvlb/parallax.html}). These measurements are important because they give us most of our knowledge about the galactic thermal electron distribution \\citep{Cordes02, Toscano99, Taylor93}. In addition, PSR J0030+0451 presents a rare evolutionary case as an isolated millisecond pulsar (MSP). In the most popular MSP evolutionary model, MSPs are formed via accretion of matter from a companion star. The incoming matter adds to the pulsar's angular momentum, i.e., the pulsar is ``spun up.'' However, isolated MSPs present a conundrum: they were presumably spun up, yet they are without a companion which would have done so. One possible scenario is that the pulsar has ablated its companion \\citep{Ruderman89}. We expect MSPs to have lower velocities than the regular population, because the kick from the supernova progenitor had to be small enough to leave the binary intact. To produce an isolated MSP, the binary must remain intact during and after the supernova, but then after the spin-up phase the companion must leave the system or be evaporated. Several authors have debated whether isolated MSP velocities are lower, higher, or indistinguishable from those of the general population of MSPs. \\cite{McLaughlin04b} suggest we might expect isolated MSPs to have higher velocities. They argue that if isolated MSPs are formed by ablation, we would expect them to form from the tighter binaries which are more susceptible to ablation. The correlation between tight binaries and higher velocities is suggested by \\citet{Tauris96}. \\citet{McLaughlin04b} present the argument for faster velocities for isolated MSPs as a counterpoint to their timing proper motion and scintillation measurements which suggest the opposite, as do the measurements of \\citet{Johnston98} and \\cite{Toscano99}. \\cite{Hobbs05}, however, find the velocities of the populations to be indistinguishable. We present a measurement of the transverse velocity of PSR J0030+0451 which is unusually small, even compared to the isolated MSP population. We reconsider the question of the velocity of isolated MSPs as compared to the binary MSP population. Owing to its small timing residual, $\\sim$1~$\\mu$s, PSR J0030+0451 is a good candidate for membership in the Pulsar Timing Array (PTA), which is a collection of pulsars that will be used for detecting gravitational radiation \\citep{Jaffe03, Jenet04}. For this reason, continued refinement of the timing model of PSR J0030+0451 is important. In fact, the PTA, as it is conceived, is an interferometer, so a variety of baselines will be important to its operation. PSR J0030+0451 may be particularly useful in this regard because it has large angular separation from PSRs B1855+09, J1713+07 and J0437-4715, which are among the most stable and precise pulsars \\citep{Kaspi94, vanStraten01, Lommenthesis}. In \\S2 we present a significant refinement to the timing model previously published \\citep{Lommen00}. We discuss our measurements of parallax and proper motion in \\S\\ref{sec:pm_px}. We include the effects of the solar wind in our analysis, which we discuss in \\S4. In \\S5,\\S6, and \\S7 we discuss the space velocity of J0030+0451, corrections to its measured period derivative, and the implications of its measured distance for the local interstellar medium (LISM). In \\S8 we summarize our conclusions. ", "conclusions": "We have measured the parallax of PSR J0030+0451 to be $3.3\\pm0.5$\\,mas. We have measured its proper motion to be $-5.74\\pm 0.09$\\,~mas~yr$^{-1}$ in the plane of the ecliptic and have established an upper limit on its motion out of the plane at 10 mas~yr$^{-1}$. The is one of the lowest velocities measured for any pulsar and is noteworthy even within the relatively low-velocity millisecond pulsar population. Combining proper motion data from this pulsar with the collection of existing MSP proper motion measurements, we find the statistical properties of the transverse velocities of isolated and binary MSPs are indistinguishable from each other. We do, however, find that the average $z$-height of the isolated MSPs is half that of the binary MSPs. We suggest that a luminosity difference between the two classes of objects, such as that suggested by \\citet{Bailes97}, \\citet{Kramer98}, and \\citet{Hobbs04}, is the simplest way to account for both the observed difference in $z$-height and the similarity of the velocity distributions." }, "0601/astro-ph0601651_arXiv.txt": { "abstract": "We present an analysis of 13 of the best quality Ultraluminous X-ray source (ULX) datasets available from \\xmmn European Photon Imaging Camera (EPIC) observations. We utilise the high signal-to-noise in these ULX spectra to investigate the best descriptions of their spectral shape in the 0.3--10 keV range. Simple models of an absorbed power-law or multicolour disc blackbody prove inadequate at describing the spectra. Better fits are found using a combination of these two components, with both variants of this model - a cool ($\\sim 0.2$ keV) disc blackbody plus hard power-law continuum, and a soft power-law continuum, dominant at low energies, plus a warm ($\\sim 1.7$ keV) disc blackbody - providing good fits to 8/13 ULX spectra. However, by examining the data above 2 keV, we find evidence for curvature in the majority of datasets (8/13 with at least marginal detections), inconsistent with the dominance of a power-law in this regime. In fact, the most successful empirical description of the spectra proved to be a combination of a cool ($\\sim 0.2$ keV) classic blackbody spectrum, plus a warm disc blackbody, that fits acceptably to 10/13 ULXs. The best overall fits are provided by a physically self-consistent accretion disc plus Comptonised corona model ({\\sc diskpn + eqpair}), which fits acceptably to 11/13 ULXs. This model provides a physical explanation for the spectral curvature, namely that it originates in an optically-thick corona, though the accretion disc photons seeding this corona still originate in an apparently cool disc. We note similarities between this fit and models of Galactic black hole binaries at high accretion rates, most notably the model of \\citet{donek05}. In this scenario the inner-disc and corona become energetically-coupled at high accretion rates, resulting in a cooled accretion disc and optically-thick corona. We conclude that this analysis of the best spectral data for ULXs shows it to be plausible that the majority of the population are high accretion rate stellar-mass (perhaps up to 80-$M_{\\odot}$) black holes, though we cannot categorically rule out the presence of larger, $\\sim 1000$-\\Msun intermediate-mass black holes (IMBHs) in individual sources with the current X-ray data. ", "introduction": "\\einstein X-ray observations were the first to reveal point-like, extranuclear sources in some nearby galaxies with luminosities in excess of $10^{39} \\ergsec$ \\citep{fabbiano89}. Subsequently, many of these so-called Ultraluminous X-ray sources (ULXs) have displayed short and long term variability, which suggests they are predominantly accreting objects (see \\citealt{milcol04} and references therein). However, the observed luminosities of most ULXs exceed the Eddington limit for spherical accretion onto a stellar-mass ($\\sim$10-$ M_{\\odot}$) black hole (BH). In fact, their luminosites are intermediate between those of normal stellar mass BH X-ray Binaries (BHBs) and Active Galactic Nuclei (AGN). Therefore, the accretion of matter onto {\\it intermediate-mass\\/} black holes (IMBHs, of $\\sim$$10^2$--$10^4$ $ M_{\\odot}$) provide an attractive explanation for ULXs, and could represent the long sought-after `missing link' between stellar mass BHs and the supermassive BHs in the nuclei of galaxies. However, the large populations of ULXs associated with sites of active star formation (\\eg in the Cartwheel galaxy, \\citealt{gao03}) demand rather too high formation rates of IMBHs if they are to explain the ULX class as a whole \\citep{king04}. An alternative to accreting IMBHs is that ULXs may be a type of stellar-mass BHB with geometrically \\citep{king01} or relativistically \\citep*{koerding02} beamed emission, such that their intrinsic X-ray luminosity does not exceed the Eddington limit. Another possibility is that they are stellar-mass BHBs that can achieve truly super Eddington luminosities via slim \\citep{ebisawa03} or radiation pressure dominated \\citep{begelman02} accretion discs. As ULXs are probably the brightest class of X-ray binary fueled by the accretion of matter onto a BH\\footnote{Although the most likely reservoir of fuel for an ULX is a companion star, others have been suggested, for example the direct accretion of matter from molecular clouds \\citep{krolik04}.}, a knowledge of the properties of Galactic BHBs could be vital in interpreting their characteristics. Traditionally the X-ray spectra of BHBs have been fitted empirically with two components, namely a power-law continuum and a multicolour disc blackbody (MCD) component (\\citealt{mitsuda84}; \\citealt{makishima86}). In the standard picture, the power-law component is thought to represent inverse-Compton scattering of thermal photons from the accretion disc by hot electrons in a surrounding corona. As such, the power-law represents the hard tail of the X-ray emission while the MCD component models the soft X-ray emission from the accretion disc. The MCD model itself has been formulated based on the best known model for accretion onto BHs (\\ie the thin accretion disc model, \\citealt{shakura73}). It has long been recognised that Galactic BHBs demonstrate various X-ray spectral states which are defined by the balance of these two components (\\ie power-law and MCD) at any one time. The three most familiar X-ray bright states are the low/hard (LH), high/soft (HS, also described as `thermal dominated') and the very high (VH, or `steep power-law') states (see \\citealt{mcclintock03} for further details). At lower mass accretion rates, a BHB usually enters the LH state where their X-ray emission is dominated by a hard power-law component ($\\Gamma$$\\sim$1.7), thought to arise from Comptonisation of soft photons by a hot optically thin corona\\footnote{However this is still a topic of debate, with the main alternative for the X-ray power-law emission being synchrotron emission from the radio jet that is associated with this state (\\eg \\citealt{falcke95}; \\citealt*{markoff01}).}. In this state, the disc is either undetected (\\eg \\citealt{belloni99}) or appears truncated at a much larger inner radius and hence cooler than the parameters derived for the soft state (\\citealt{wilms99}, \\citealt{mcclintock01}). The soft X-ray state is generally seen at a higher luminosity (\\ie the HS state) and is best explained as $\\sim$1 keV thermal emission from a multi-temperature accretion disc (\\ie modelled with a MCD component). In this state, the spectrum may also display a hard tail that contributes a small percentage of the total flux. The VH state is in many cases the most luminous state and is characterised by an unbroken power-law spectrum extending out to a few hundred keV or more. The photon index is typically steeper ($\\geq 2.5$) than found in the LH state and generally coincides with the onset of strong X-ray quasi-periodic oscillations (QPOs). A MCD component may also be present in the VH state and the \\exosat era demonstrated that some of the QPOs occur when both disc and power-law components contribute substantial luminosity \\citep{vanderklis95}. The idea of ULXs as analogues to Galactic BHBs in the HS state was supported by \\asca observations, which revealed that their 0.5--10 keV spectra were successfully fitted with the MCD model with relatively high disc temperatures (1.0--1.8 keV, \\citealt{makishima00}). As such, the ULXs were considered to be mass-accreting BHs with the X-ray emission originating in an optically-thick accretion disc. In fact, the use of the MCD model to describe these spectra permits one to obtain an `X-ray--estimated' BH mass, $M_{\\rm XR}$, from the following equation (cf. \\citealt{makishima00} equations (5)--(8)). \\begin{equation} M_{\\rm XR} = {{\\xi\\kappa^2}\\over{8.86\\alpha}} \\frac{D}{\\sqrt{{\\rm cos} i}} \\sqrt{\\frac{f_{\\rm bol}}{2\\sigma T^4_{\\rm in}}} \\; M_{\\odot} \\end{equation} Where $D$ is the distance to the X-ray source, which has an inclination $i$, a full bolometric luminosity (from the MCD model) of $f_{\\rm bol}$ and an observed maximum disc colour temperature $T_{\\rm in}$. In addition $\\sigma$ is the Stefan-Boltzmann constant, $\\kappa$ is the ratio of the colour temperature to the effective temperature (`spectral hardening factor'), and $\\xi$ is a correction factor reflecting the fact that $T_{\\rm in}$ occurs at a radius somewhat larger than $R_{\\rm in}$ (here, we assume that $R_{\\rm in}$ is at the last stable Keplerian orbit). \\cite{makishima00} use values of $\\xi = 0.412$ and $\\kappa = 1.7$, though other work has found different values for the spectral hardening factor (e.g. $\\kappa = 2.6$ for GRO J1655-40, \\citealt{st03}). Finally, $\\alpha$ is a positive parameter with $\\alpha = 1$ corresponding to a Schwarzschild BH. However, the masses inferred from the \\asca data and Equation (1) are far too low to be compatible with the large BH masses suggested by their luminosities (assuming Eddington-limited accretion), for standard accretion discs around Schwarzchild BHs. \\citet{makishima00} suggested that this incompatibility could be explained if the BHs were in the Kerr metric (\\ie rapidly rotating objects), allowing smaller inner disc radii and hence higher disc temperatures. \\chandra observations have provided some support for the \\citet{makishima00} results, with some ULX spectra being consistent with the MCD model (\\eg \\citealt{roberts02}). However, \\chandra also revealed that some ULX spectra showed a preference for a power-law continuum rather than the MCD model (\\eg \\citealt{strickland01}; \\citealt{roberts04}; \\citealt{terashima04}). It has been suggested that this preference for a power-law spectrum could be interpreted in terms of the LH state seen in Galactic BHB candidates, relativistically beamed jets or emission from a Comptonised accretion disc in the VH state. As well as these single component models, \\asca and \\chandra spectroscopy have also suggested the presence of two component spectra for some ULXs, comprising a MCD with a power-law component. For example, previous \\asca analyses hinted at evidence for IMBHs, \\ie cool accretion disc components (see below), but these observations were not sensitive enough to statistically require two component modelling (\\eg \\citealt{colbert99}). Similar results have been obtained by {\\it Chandra}, \\eg ULXs in NGC 5408 \\citep{kaaret03} and NGC 6946 \\citep{roberts03}. Conversely, \\chandra spectra of the Antennae ULXs (\\citealt{zezas02a}; \\citealt{zezas02b}) revealed an accretion disc (MCD) component consistent with the high temperature \\asca results (\\ie $kT_{\\rm in}$$\\sim$1 keV), together with a hard power-law component ($\\Gamma$$\\sim$1.2). It is only recently, using high quality {\\it XMM-Newton}/EPIC spectroscopy of ULXs, that it has been demonstrated that the addition of a soft thermal disc component to a power-law continuum spectrum provides a strong statistical improvement to the best fitting models to ULX data (\\eg \\citealt{miller03}; \\citealt*{miller04a}). These particular observations have provided strong support for the IMBH hypothesis by revealing disc temperatures in these sources up to 10 times lower than commonly measured in stellar mass BHBs, consistent with the expectation for the accretion disc around a $\\sim$1000-\\Msun IMBH\\footnote{It is common for the generic range of masses for IMBHs to be quoted as 20--$10^6$ \\Msun. The lower limit comes from a consideration of the measured masses of BHs in our own Galaxy \\citep{mcclintock03}, and a theoretical limit for the mass of a BH formed from a single massive star \\citep{fryer01}. However, more recent population synthesis analyses show that BHs of up to $\\sim 80$-\\Msun may be formed in young stellar populations \\citep{belczynski04}. Hence, when we refer to IMBHs in this paper we refer specifically to the larger $\\sim$1000-\\Msun IMBHs implied by the cool accretion disc measurements.} (cf. Equation (1)). However, in a few cases, {\\it XMM-Newton}/EPIC observations have revealed a more unusual two-component X-ray spectrum. A detailed analysis of such a source is presented in \\citet*{stobbart04}. In this case, the ULX is the brightest X-ray source in the nearby (1.78 Mpc) Magellanic-type galaxy NGC 55 and, although its X-ray luminosity only marginally exceeds $10^{39} \\ergsec$, it represents one of the highest quality ULX datasets obtained to date. The initial low flux state data were best fitted with an absorbed power-law continuum ($\\Gamma$$\\sim$4), while a subsequent flux increase was almost entirely due to an additional contribution at energies $> 1$ keV, adequately modelled by a MCD component ($kT_{\\rm in}$$\\sim$0.9 keV). Whilst this accretion disc component is reasonable for stellar-mass BHs, the dominance of the power-law continuum at soft X-ray energies is problematic. Such a soft power-law cannot represent Comptonised emission from a hot corona, as one would not expect to see the coronal component extend down below the peak emissivity of the accretion disc, where there would be insufficient photons to seed the corona. Alternative sources of seed photons for the corona are unlikely; for example, the incident photon flux of the secondary star at the inner regions of the accretion disc is too low to provide the seeding (cf. \\citealt{roberts05}). It also seems unlikely that the power-law emission could arise from processes at the base of a jet, as these are typically represented by much harder photon indices than measured here ($\\Gamma$$\\sim$1.5--2; \\citealt{mnw05} and references therein). Indeed, with the possible exception of NGC 5408 X-1 \\citep{kaaret03}, there is no evidence that ULXs do display bright radio jets, though this cannot be excluded by current observations \\citep{kcf05}. The possibility of the soft component resulting from an outflow of material from the accretion disc may also be discounted as this would produce a thermal spectrum rather than a power-law continuum. Although this spectral description has not been seen in Galactic systems, a second case has been reported independently for the nearest persistent extragalactic ULX (M33 X-8) by \\citet{foschini04}. The non-standard model provided the best fit to this ULX with $\\Gamma$$\\sim$2.5 and $kT_{\\rm in}$$\\sim$1.2 keV. However, this source is also at the low luminosity end of the ULX regime with \\lx$\\sim$2$\\times 10^{39}\\ergsec$. A further possible case, in a more luminous ULX, has arisen from the \\xmmn data analysis of NGC 5204 X-1 \\citep{roberts05}. In this case the authors show that there is spectral ambiguity between the non-standard fit ($\\Gamma$$\\sim$3.3, $kT_{\\rm in}$$\\sim$ 2.2 keV) and the IMBH model ($\\Gamma$$\\sim$2.0, $kT_{\\rm in}$$\\sim$ 0.2 keV), with both providing statistically acceptable fits to the data. Even more recently, two additional examples of this spectral form have been uncovered in an \\xmmn survey of ULXs by \\citet{feng05}. Although it is difficult to derive a literal physical interpretation from the non-standard model, it does provide an accurate empirical description of ULX spectra in some cases, and as such it has the potential to provide new insights into the nature of these sources. Therefore in this paper we re-evaluate current data in an attempt to determine the best spectral description for the shape of high quality ULX spectra, and ask what consequences this has for the idea of ULXs as accreting IMBHs. The paper is structured as follows: Sec.~\\ref{sec_sample} -- introduction to the ULX sample; Sec.~\\ref{sec_obs} -- details of the observations and data reduction; Sec.~\\ref{sec_spectra} -- description of the spectral analysis; Sec.~\\ref{sec_lx_kt} -- a comment on the luminosity and inner disc temperature relationship of these ULXs; Sec.~\\ref{sec_discussion} -- a discussion of our results; and finally Sec.~\\ref{sec_conclusions} -- our conclusions. ", "conclusions": "\\label{sec_conclusions} We have conducted a detailed examination of the X-ray spectral shapes in a sample of the highest quality \\xmmn EPIC ULX datasets available to us. Most notably, more than half of the ULXs show at least marginal evidence for curvature in their 2--10 keV spectra, which is somewhat unexpected if they are to be interpreted as the accreting $\\sim 1000$-\\Msun IMBHs suggested by modelling the soft spectral components as accretion discs. Physical modelling shows that this curvature is likely to originate in optically-thick coronae, which in turn leads to interpretations of the ULXs in terms of high accretion-rate stellar-mass (or slightly larger) BHs operating at around the Eddington limit. However, while we conclude that it is likely that the general ULX population does not have a large contribution from IMBHs, we obviously cannot rule out the possibility that some ULXs do possess IMBHs. Perhaps the best candidate on the basis of our spectral fitting is M81 X-9, which is well fitted by cool disc plus power-law/optically-thin corona models, and does not show explicit curvature in its 2--10 keV spectrum, though even this ULX may be fitted using a hot ($\\sim 2.2$ keV) accretion disc plus soft excess model. Clearly, it is difficult to find unique solutions for these sources even with high quality \\xmmn EPIC data. Ultimately, we may perhaps have to wait for radial velocity measurements from the optical counterpart of an ULX, leading to dynamical mass measurements of the compact accretor, before we have conclusive evidence whether any individual ULX does harbour an IMBH." }, "0601/astro-ph0601698_arXiv.txt": { "abstract": "Binary supermassive black holes form naturally in galaxy mergers, but their long-term evolution is uncertain. In spherical galaxies, $N$-body simulations show that binary evolution stalls at separations much too large for significant emission of gravitational waves (the ``final parsec problem''). Here, we follow the long-term evolution of a massive binary in more realistic, triaxial and rotating galaxy models. We find that the binary does not stall. The binary hardening rates that we observe are sufficient to allow complete coalescence of binary SBHs in $10$ Gyr or less, even in the absence of collisional loss-cone refilling or gas-dynamical torques, thus providing a potential solution to the final parsec problem. ", "introduction": "When two galaxies containing supermassive black holes (SBHs) merge, a binary SBH forms at the center of the new galaxy. The two SBHs can eventually coalesce, but only after stellar- or gas-dynamical processes bring them close enough together ($\\lap 10^{-2}$ pc) that gravitational radiation is emitted. There is strong circumstantial evidence that rapid coalescence is the norm. For instance, no binary SBH has ever been unambiguously observed \\citep{LR05}. Furthermore, in a galaxy containing an uncoalesced binary, mergers would eventually bring a third SBH into the nucleus, precipitating a gravitational slingshot interaction that would eject one or more of the SBHs from the nucleus \\citep{MV90,VHM03}. This could produce off-center SBHs, and could also weaken the tight correlations that are observed between SBH mass and galaxy properties \\citep{FM00,G01,MH03}. Unless the binary mass ratio is extreme, dynamical friction rapidly brings the smaller SBH into a distance $\\sim G\\mu/\\sigma^2$ from the larger SBH, where $\\mu\\equiv M_1M_2/(M_1+M_2)$ is the binary reduced mass and $\\sigma$ is the 1D velocity dispersion of the stars. At this separation -- of order $1$ pc -- the two SBHs begin to act like a ``hard'' binary, ejecting passing stars with velocities large enough to remove them from the nucleus. $N$-body simulations \\citep{MF04,SMM05,BMS05} show that continued hardening of the binary takes place at a rate that depends strongly on the number $N$ of ``star'' particles used in the simulation. As $N$ increases, the hardening rate falls, as expected if the binary's loss cone is repopulated by star-star gravitational encounters \\citep{Y02,MM03}. When extrapolated to the much larger $N$ of real galaxies, these results suggest that binary evolution would generally stall (the ``final parsec problem''). To date, $N$-body simulations of the long-term evolution of binary SBHs have only been carried out using spherical or nearly spherical galaxy models. But it has been suggested \\citep{MP04} that binary hardening might be much more efficient in non-axisymmetric galaxies due to the qualitatively different character of the stellar orbits. Here, we test that suggestion by carrying out the first $N$-body simulations of massive binaries in strongly non-axisymmetric galaxy models. We find that the hardening rate is {\\it independent} of $N$ for particle numbers up to at least $0.4\\times 10^6$. To the extent that our galaxy models are similar to real merger remnants, these results imply that binary SBHs can efficiently harden through purely stellar-dynamical interactions in many galaxies, thus providing a plausible solution to the final parsec problem. ", "conclusions": "" }, "0601/astro-ph0601301_arXiv.txt": { "abstract": "We present a \\chandra\\ study of mass profiles in 7 elliptical galaxies, of which 3 have galaxy-scale and 4 group-scale halos, demarcated at $10^{13}$\\msun. These represent the best available data for nearby objects with comparable X-ray luminosities. We measure $\\sim$flat mass-to-light (M/L) profiles within an optical half-light radius (\\reff), rising by an order of magnitude at $\\sim$10\\reff, which confirms the presence of dark matter (DM). The data indicate hydrostatic equilibrium, which is also supported by agreement with studies of stellar kinematics in elliptical galaxies. The data are well-fitted by a model comprising an NFW DM profile and a baryonic component following the optical light. The distribution of DM halo concentration parameters (c) {\\em versus} \\mvir\\ agrees with \\lcdm\\ predictions and our observations of bright groups. Concentrations are slightly higher than expected, which is most likely a selection effect. Omitting the stellar mass drastically increases c, possibly explaining large concentrations found by some past observers. The stellar M/\\lk\\ agree with population synthesis models, assuming a Kroupa IMF. Allowing adiabatic compression (AC) of the DM halo by baryons made M/L more discrepant, casting some doubt on AC. Our best-fitting models imply total baryon fractions $\\sim$0.04--0.09, consistent with models of galaxy formation incorporating strong feedback. The groups exhibit positive temperature gradients, consistent with the ``Universal'' profiles found in other groups and clusters, whereas the galaxies have negative gradients, suggesting a change in the evolutionary history of the systems around \\mvir$\\simeq 10^{13}$\\msun. ", "introduction": "The nature and distribution of dark matter (DM) in the Universe is one of the fundamental problems facing modern physics. Cold DM lies at the heart of our current (\\lcdm) cosmological paradigm, which predicts substantial DM halos for objects at all mass-scales from galaxies to clusters. Although \\lcdm\\ has been remarkably successful at explaining large-scale features \\citep[\\eg][]{spergel03a,perlmutter99a}, observations of galaxies have been more problematical for the theory. Dissipationless dark matter simulations find that dark matter halos are well characterized by a ``Universal'' mass density profile \\citep[][hereafter NFW]{navarro97} over a wide range of Virial masses (\\mvir) \\citep[e.g.][]{bullock01a}. Low mass halos tend to form first in hierarchical cosmologies and are consequently more tightly concentrated than their later forming, high mass counterparts. This tendency produces a predicted correlation between the DM halo concentration parameter (c, which is ratio between Virial radius, \\rvir, and the characteristic scale of the density profile) and \\mvir \\citep{navarro97}. However, since mass and formation epoch are not perfectly correlated, we expect a significant scatter at fixed Virial mass \\citep{jing00a,bullock01a,wechsler02a}. The tight link between halo formation epoch and concentration implies that the precise relation between c and \\mvir\\ is sensitive to the underlying Cosmological parameters, including \\sigmaeight\\ and the dark energy equation of state \\citep{kuhlen05a}, making an observational test of this relation potentially a very powerful tool for cosmology. The mass profiles of galaxies also may provide valuable clues as to the way in which galaxies form in DM halos. In particular, as baryons cool and collapse into stars, the associated increase in the central mass density should in turn modify the shape of the DM halo. This process is typically modelled assuming adiabatic contraction (AC) of the DM particle orbits \\citep[\\eg][]{blumenthal86a,gnedin04a}. If the galaxy halo subsequently evolves by major mergers, simulations are unclear as to whether these features would persist \\citep[\\eg][]{gnedin04a} or whether the merging process may destroy this imprint of star formation, or even mix the DM and baryons sufficiently to produce a {\\em total} gravitating mass profile more akin to NFW \\citep{loeb03a,elzant04a}. Observational tests of the predictions of \\lcdm\\ have proven controversial. In clusters of galaxies there is overwhelming evidence for DM, and an increasing body of work verifying the predictions of \\lcdm. In particular recent, high-quality \\chandra\\ and \\xmm\\ observations have revealed mass profiles in remarkable agreement with the Universal profile from deep in the core to a large fraction of \\rvir\\ \\citep[\\eg][]{lewis03a,zappacosta06a,vikhlinin05b}, and a distribution of c {\\em versus} \\mvir\\ in good agreement with \\lcdm\\ \\citep{pointecouteau05a}. In galaxies, however, the picture is much less clear. Rotation curve analysis of low surface brightness (LSB) disk galaxies has suggested significantly less cuspy density profiles than expected \\citep[\\eg][]{swaters00a}. Although this discrepancy led to a serious discussion of modifications to the standard paradigm \\citep[\\eg][]{hogan00a,spergel00a,zentner02a,kaplinghat05a,cembranos05a}, recent results, taking account of observational bias and the 3-dimensional geometry of the DM halos, have done much to resolve the discrepancy \\citep[\\eg][]{swaters03a,simon05a}. However, some significant discrepancies remain, not least of which is that the DM halos of these galaxies appear less concentrated than expected \\citep[\\eg][]{gonzalez00a,kassin06a}. A possible explanation is that LSB galaxies are preferentially found in low-concentration halos \\citep{bullock01a,bailin05a,wechsler05a}, making additional constraints at the galaxy scale extremely important. % In many respects, kinematical mass measurements are far more challenging for early-type than spiral galaxies. As essentially pressure-supported systems little is known {\\em a priori} about the velocity anisotropy tensor of the stars in elliptical galaxies, which is problematical for the determination of the mass from stellar motions. Nonetheless, stellar kinematical measurements have widely been used as a means to measure the gravitating matter within $\\sim$the optical half-light radius (\\reff) of elliptical galaxies \\citep[\\eg][]{binney90a,vandermarel91a,gerhard01a}. These studies tend to find relatively flat mass-to-light (M/L) ratios within \\reff, implying that most of the matter within this radius is baryonic. Consideration of the tilt in the fundamental plane can also lead to the same conclusion \\citep{borriello03a}. In contrast, \\citet{padmanabhan04a} pointed out that dynamical M/L ratios within \\reff\\ are much larger than predicted by realistic stellar population synthesis models for stars alone, allowing \\gtsim 50\\% of the mass within \\reff\\ to be dark. % Attempts to extend kinematical studies of elliptical galaxies to larger radii, where DM should be dominant, have proven controversial. In particular \\citet{romanowsky03a} argued against the existence of DM in a small sample of elliptical galaxies, based on planetary nebulae dynamics within $\\sim$5\\reff. We note that this sample was heavily biased towards very X-ray faint objects, which might hint at low-mass halos since they have not held onto their hot gas. In any case \\citet{dekel05a} pointed out that their conclusions were very sensitive to the uncertainty in the velocity anisotropy tensor, for plausible values of which the data were consistent with substantial DM halos. In fact globular cluster dynamics in one of these systems, NGC\\thin 3379, does imply a significant amount of DM \\citep{pierce06a,bergond06a}. As more kinematical studies of early-type galaxies at large radii are appearing, it is becoming clear that at least some elliptical galaxies host considerable DM halos \\citep[\\eg][]{statler99a,romanowsky05a}. There persist some questions, however, as to the extent to which all galaxies have DM halos consistent with \\lcdm. In particular \\citet{napolitano05a} argued that a substantial number of early-type galaxy halos appear less concentrated than expected. Gravitational lensing provides further evidence that, at least some, early-type galaxies possess substantial DM halos \\citep[\\eg][]{kochanek95a,fischer00a,rusin02a}. Since weak lensing of galaxies only provides useful mass constraints in a statistical sense, the relatively rare instances of strong lensing are required to study DM in individual systems. Nonetheless it has been possible in a few cases to decompose the mass into stellar and DM components, albeit with strong assumptions or additional observational constraints \\citep[\\eg][]{rusin03a,treu04a}. X-ray observations of the hot gas in early-type galaxies provide a complementary means to infer the mass-profiles {\\em via} techniques similar to those used in studying clusters. Since the X-ray emission from early-type galaxies is typically not very bright, prior to the advent of \\chandra\\ and \\xmm\\ this was limited by the relatively sparse information on the radial temperature and density profiles of the hot gas which could be determined by prior generations of satellites. Notwithstanding this limitation, large M/L ratios (consistent with substantial DM) were inferred for a number of X-ray bright galaxies, albeit with strong assumptions concerning the temperature and density profiles \\citep[\\eg][]{forman85a,loewenstein99b}. Using a novel technique which relied, instead, on the ellipticity of the X-ray halo, \\citet{buote94} were able robustly to detect DM in the isolated elliptical NGC\\thin 720 \\citep[see also][]{buote96a,buote98d,buote02b}. Detailed measurements of the radial mass distribution were, however, largely restricted to a few massive systems, which may be entwined with a group halo \\citep[\\eg][]{irwin96,brighenti97a}. Nevertheless \\citet{brighenti97a} were able to decompose the mass profiles of two systems, NGC\\thin 4472 and NGC\\thin 4649, into stellar and DM components. \\citet{sato00a} investigated the \\mvir-c relation using \\asca\\ for a sample of objects ranging from massive clusters to $\\sim$3 elliptical galaxies. The limited spatial resolution of \\asca\\ necessitated some assumptions about the density profiles and, crucially, the authors neglected any stellar mass component in their fits. This omission may explain the very steep \\mvir-c relation (with c$_{200}$\\gtsim 30 for the galaxies) found by these authors, in conflict with \\lcdm\\ \\citep{mamon05a}. Although mass profiles of early-type galaxies are beginning to appear which exploit the improved sensitivity and resolution of \\chandra\\ and \\xmm, many of the most interesting constraints on DM are still restricted to massive systems, which may be at the centres of groups. For example, \\citet{fukazawa06a} reported \\chandra\\ and \\xmm\\ M/L profiles for $\\sim$50 galaxies and groups, confirming $\\sim$flat profiles within \\reff\\ which rise at larger radii. However, the constraints at large radii were dominated by the massive (group-scale) objects so the implications for the DM content of normal galaxies are unclear. Furthermore, the authors included a substantial number of highly disturbed systems, in which hydrostatic equilibrium may be questioned, and failed to account for the unresolved sources which dominate the emission in the lowest-\\lx\\ objects in their sample\\footnote{Although the authors account for unresolved sources when measuring the gas temperature, they do not account for it when computing the gas density, where its effect is more pronounced}. Recently, however, detailed \\chandra\\ and \\xmm\\ mass profiles have begun to appear for isolated early-type galaxies, also confirming the presence of massive DM halos \\citep[\\eg][]{osullivan04b,khosroshahi04a}. This paper is part of a series \\citep[see also][]{gastaldello06a,zappacosta06a,buote06b,buote06a} using high-quality \\chandra\\ and \\xmm\\ data to investigate the mass profiles of galaxies, groups and clusters. This provides an unprecedented opportunity to place definitive constraints upon the \\mvir-c relation over $\\sim$2 orders of magnitude in \\mvir. In this paper, we focus on the temperature, density and mass profiles of seven galaxies and poor groups chosen from the \\chandra\\ archive. In order to compare to theory we perform spherically-averaged analysis, leaving a discussion of the ellipticities of the X-ray halos to a future paper. In \\S~\\ref{sect_targets} we discuss the target selection. The data-reduction is described in \\S~\\ref{sect_reduction} and the X-ray morphology is addressed in \\S~\\ref{sect_imaging}. We discuss the spectral analysis in \\S~\\ref{sect_spectra}, the mass analysis in \\S~\\ref{sect_mass}, the systematic uncertainties in our analysis in \\S~\\ref{sect_systematics} and reach our conclusions in \\S~\\ref{sect_discussion}. The three systems for which we find \\mvir$<10^{13}$\\msun\\ are optically isolated and so we refer to them as ``galaxies'', and the other systems in our sample as groups. We discuss this in more detail in \\S~\\ref{discussion_groups}. In this paper, all error-bars quoted represent 90\\% confidence limits, unless otherwise stated, and we computed Virial quantities assuming a ``critical overdensity'' factor for the DM halos of $\\rho_{\\rm halo}/\\rho_{\\rm crit} = 103$ (where $\\rho_{\\rm halo}$ is the mean density of a sphere of mass \\mvir\\ and radius \\rvir). ", "conclusions": "\\label{sect_discussion} \\subsection{Hydrostatic equilibrium} Our fit results provide strong evidence that the gas is in hydrostatic equilibrium in these systems. Despite highly nontrivial temperature and density profiles, we were able to recover smooth mass profiles in remarkably good agreement with expectation for these systems, using two complementary techniques. If the gas is significantly out of hydrostatic equilibrium, this would represent a remarkable ``conspiracy'' between the density and temperature profiles. It is unsurprising that the gas is close to hydrostatic equilibrium in these systems, since we took care to choose objects with relaxed X-ray morphology. Based on N-body/ hydrodynamical analysis, X-ray measurements are expected to give reliable constraints on the DM in systems without obvious substructure \\citep{buote95a}. Further support for hydrostatic equilibrium is provided by the general agreement between our measured \\mstars/\\lk\\ ratios and those predicted by SSP models, coupled with the agreement between the measured \\mvir-c relation and that expected. Similarly a comparison between our results and masses determined from stellar dynamics provides even more evidence that our measured mass profiles are reliable. Dynamically-determined \\mgrav/\\lb\\ within the B-band \\reff\\ are typically found to be $\\sim$4--10 \\citep{gerhard01a,trujillo04a}. We found \\mgrav/\\lb\\ within the B-band \\reff\\ (taken from RC3 or \\citealt{faber89}) for our systems ranged from $\\sim$3 to $\\sim$8, in good agreement with this result. For NGC\\thin 4649, outside \\reff\\ there is excellent agreement between our measured \\mgrav/L profile and that obtained from globular cluster kinematics, although at small radii the X-ray data lie $\\sim$30\\% lower (K.\\ Gebhardt et al, in preparation). \\citet{vandermarel91a} constructed stellar kinematical models for 5 galaxies in our sample (NGC\\thin 720, NGC\\thin 1407 and NGC\\thin 4261, NGC\\thin 4472 and NGC\\thin 4649), under the assumption of a constant M/L profile. Strictly speaking a direct comparison cannot be made between their \\mgrav/\\lb\\ measurements and our results since our data indicate this assumption is incorrect. However, if we simply assume that these M/L ratios represent those integrated out to \\reff, the X-ray inferred masses vary from $\\sim$40\\% lower to $\\sim$10\\% higher than those from kinematics. \\citet{kronawitter00a} report \\mgrav/\\lb$\\sim$8$\\pm1.5$ for NGC\\thin 4472 within $\\sim$50\\arcsec, at which radius our X-ray determined value is $\\sim$50\\% lower. The discrepancies between the X-ray and dynamical masses are only modest (the X-ray mass being on average $\\sim$20\\% lower), indicating that the data must be close to hydrostatic equilibrium. Turbulence is expected to contribute only $\\sim$10\\% pressure support in clusters, which are believed to be more turbulent than galaxies, \\citep{rasia06a}. Therefore, on a case-by-case basis, the observed differences are most likely a manifestation of the mass-anisotropy degeneracy \\citep[\\eg][]{dekel05a}. \\subsection{Mass profiles} We obtained detailed mass profiles for 3 galaxies and 4 group-scale systems, out to $\\sim$10\\reff. The data clearly show M/L profiles which are $\\sim$flat within \\reff\\ and rise considerably outside this range. This confirms the presence of substantial DM in at least some early-type galaxies and indicates that a stellar mass component dominates within $\\sim$\\reff. This is consistent with studies of stellar kinematics and similar to the mass decomposition analysis of \\citet{brighenti97a}. The data are well-fitted by a model comprising a stellar mass (H90) component and an NFW DM profile. Omitting the stellar mass component led to systematically poorer fits, smaller \\mvir\\ and vastly larger c ($\\gg$20), which are inconsistent with the predictions of \\lcdm. This effect is easy to understand--- if we add a compact stellar mass component to an (extended) NFW profile, we increase the mass in the core which, by definition, makes the halo more concentrated. However, it is not entirely clear whether this effect, pointed out by \\citet{mamon05a}, can completely account for the significantly steeper \\mvir-c relation found by \\citet{sato00a}. Based on our analysis of group-scale halos \\citep{gastaldello06a} we found that the inclusion of the stellar mass component does not have a strong effect on c in most systems with \\mvir \\gtsim 2$\\times 10^{13}$\\msun, provided the data are fitted to a sufficiently large fraction of \\rvir. The data did not allow us to distinguish statistically between the simple NFW+stars model and scenarios in which the DM halo experiences adiabatic compression due to star formation (however, see \\S~\\ref{discussion_mass_to_light}), or the NFW profile was replaced with the alternative N04 profile. Comparing our inferred \\mvir\\ and c to the predictions of \\lcdm\\ we find general agreement. There is some evidence, however, that the concentrations are systematically higher than one would expect, although the error-bars are typically large. Such a trend is also seen in our analysis of groups \\citep{gastaldello06a}. Whilst the slope of the \\mvir-c relation therefore implied by our data is difficult to explain by varying the cosmological parameters within reasonable limits \\citep{buote06b}, we suspect that the discrepancy can be resolved by taking into account the selection function of our galaxies. Our data were not selected in a statistically complete manner and, by choosing objects with relatively relaxed X-ray morphologies we are probably selecting objects which have not had a recent major merger. This systematically biases us towards early-forming, hence higher concentration halos. In fact, it is striking that all three {\\em de facto} galaxies in our sample are relatively isolated systems (\\S~\\ref{discussion_groups}). Such systems preferentially might be expected to occupy high-c halos \\citep{zentner05a}, which does appear to be the case for 2 out of 3 of the galaxies. We will return to these issues in detail in \\citet{buote06b}. \\subsection{Galaxies, Groups and Fossil Groups} \\label{discussion_groups} All three of the lowest-mass systems in our sample are very isolated optically. NGC\\thin 6482 matches the isolation criteria adopted to identify so-called ``fossil groups'' \\citep{khosroshahi04a}. NGC\\thin 4125 and NGC\\thin 720 are both listed as ``groups'' in G93, but closer inspection actually reveals they are also very isolated. Excepting the central galaxy, only one of the putative members of the NGC\\thin 720 ``group'' listed in the G93 catalogue \\citep[which omits the dwarf galaxy population studied by][]{dressler86a}, actually lies within the projected \\rvir\\ (but outside 0.75$\\times$\\rvir) and it is 2.4 magnitudes fainter in B than the central galaxy. \\citeauthor{dressler86a} remarked upon the optical isolation of this galaxy. Of the two putative companion galaxies to NGC\\thin 4125 given in G93 which lie within the projected \\rvir\\ (but outside 0.67$\\times$\\rvir), both are much fainter (by 2.3 and 3.9 magnitudes, respectively) in B than the central galaxy. In contrast, the four remaining systems in our sample are much less optically isolated. \\citet{schindler99a} show the clear over-density of early-type galaxies around NGC\\thin 4649 and NGC\\thin 4472, and almost 60 group members are associated with these systems by G93. \\citet{gould93a} identified at least 10 members of the NGC\\thin 1407 group, from the dynamics of which he inferred a mass broadly consistent with our measured \\mvir\\ (\\S~\\ref{sect_ngc1407}). \\citet{helsdon03a} report 57 galaxies associated with the NGC\\thin 4261 group within $\\sim$1~Mpc projected radius, which is consistent with our measured \\rvir. Rather than an isolated galaxy \\citet{khosroshahi04a} identify NGC\\thin 6482 as a ``fossil group''. Fossil groups are group-sized X-ray halos centred on essentially a single elliptical galaxy \\citep{ponman94,vikhlinin99a,jones03}. The typical interpretation of these objects is groups in which all of the \\lstar\\ members have merged. Confusingly, using almost the same selection criteria, \\citet{osullivan04b} classify the galaxy NGC\\thin 4555 as an ``isolated elliptical galaxy'' and posit a very different formation scenario. This object appears to be more massive than NGC\\thin 6482; the authors found \\mgrav $\\sim 3\\times 10^{12}$\\msun\\ within 60~kpc which, assuming an NFW profile with c$=$15 would imply \\mvir $\\sim 2\\times 10^{13}$\\msun. Nonetheless, both of these systems have more in common (both optically and in the X-ray) with each other, and the other isolated ellipticals in our sample, than, for example, the massive (\\mvir\\gtsim$10^{14}$\\msun), hotter (kT$\\sim$2~keV) fossil groups considered by \\citet{vikhlinin99a}. We suspect that the distinction made between ``isolated elliptical'' and ``fossil group'' for these two systems is largely semantic, and consider NGC\\thin 6482 more properly an isolated galaxy, too. The clear division in the galaxy content of our sample clearly lends itself to the nomenclature ``galaxies'' for the three lowest-mass systems, and ``groups'' for the others. Strikingly, this separation between galaxies and groups also appears consistent with a difference in temperature profiles (\\S~\\ref{discussion_temp}). That this distinction appears commensurate with \\mvir$\\sim 10^{13}$\\msun\\ is suggestive that this mass-scale may be a useful yard-stick with which to compare to other systems. The error-bars on our mass estimates are sufficiently large that the 90\\% confidence regions of several of the objects (notably NGC\\thin 720, NGC\\thin 6482 and NGC\\thin 1407) actually straddle $10^{13}$\\msun. However, it is clear that {\\em on average}, the systems with \\mvir \\ltsim $10^{13}$\\msun\\ are galaxies. We note that the \\mvir\\ adopted here is that {\\em before} any tidal truncation which is almost certainly occurring as NGC\\thin 4472 and NGC\\thin 4649 merge with Virgo (their untruncated \\rvir\\ would stretch much of the distance to M\\thin 87). \\mvir\\ does not exactly correlate with formation epoch, so that lower-mass halos may still be in the process of forming (hence contain multiple galaxies of similar magnitude), and more massive halos may contain single, dominant ellipticals (fossil groups). Nonetheless, classifying halos primarily on the basis of \\mvir\\ provides a straightforward way to locate them in the formation hierarchy. Traditionally, galaxy-like and group-like systems are distinguished on the basis of local over-densities of galaxies. However, placing optically-identified groups into a cosmological context requires a firm understanding not only of the formation of DM halos but also how galaxies populate them, which is much less well-understood \\citep[\\eg][]{kravtsov04a}. This problem is compounded by the difficulties faced by optical group-finding algorithms in identifying very poor groups \\citep[\\eg][]{gerke05a}. Not only can a significant fraction of putative groups be chance superpositions of galaxies, particularly along filaments, but adjacent groups can be merged, such as happened for NGC\\thin 4649 and NGC\\thin 4472 in G93. If there are only a few identified members, small-number statistics and the treatment of interlopers can affect their interpretation \\citep[\\eg][]{gould93a}. To complicate matters further, some authors refer to {\\em any} over-density of galaxies as a group, even a Milky Way-sized galaxy and its dwarf satellites \\citep[\\eg][]{tully05a}. \\subsection{Stellar Mass-to-Light Ratios} \\label{discussion_mass_to_light} Comparing our measured stellar M/L ratios to the predictions of simple stellar population (SSP) models, we found reasonable agreement provided one assumes a \\citet{kroupa01a} IMF. There is modest disagreement, even when the less-cuspy N04 DM model was adopted. Considering the uncertainties in the SSP modelling (discussed below), however, we believe this discrepancy is not significant. If we allowed the DM profile to be modified by adiabatic compression, we obtained substantially smaller \\mstars/\\lk\\ values from our data, (since it increases the cuspiness of the halo) which are more discrepant with the SSP models. This result casts doubt on AC being as significant an effect as currently modelled. However, the data alone did not allow us statistically to distinguish between the NFW+stars and AC NFW+stars models. Nonetheless, this result is joining a growing body of literature which similarly calls into question whether AC operates as predicted \\citep{zappacosta06a,kassin06a,sand04a}. There are a number of major uncertainties in the computation of the stellar mass-to-light ratios from the SSP models. Specifically, the results are very dependent upon the assumed IMF, which is not confidently known in early-type galaxies. Furthermore there is some evidence that early-type galaxies frequently contain multiple stellar populations of different ages, including a significant young population \\citep[\\eg][]{rembold05a,nolan06a}. Depending on the mass fraction of the young component, this may substantially reduce \\mstars/\\lk\\ in the galaxy, hence possibly reconciling the data and the AC NFW+stars model. A small amount of star formation may also give rise to a population of stars which can dominate the light in the galaxy core, giving rise to significantly lower synthetic \\mstars/\\lk\\ than measured. This may be the case in NGC\\thin 720 (see \\S~\\ref{sect_mass_to_light}). More problematically, there are known to be significant abundance, or possible age, gradients in the stellar populations of early-type galaxies \\citep[\\eg][]{trager00a,kobayashi99a,rembold05a}, which would translate into stellar \\mstars/\\lk\\ gradients. Our simple modelling did not allow us to account for such an effect {\\em per se}. Although we suspect that such gradients will primarily lead to a \\mstars/\\lk\\ value which reflects an average for the galaxy, \\mstars/\\lk\\ does depend to some extent upon the shape of the assumed stellar potential. Properly taking account of this effect is beyond the scope of this present work, but may bring the synthetic M/L ratios and our results into better agreement. Clearly this is only one of a number of other systematic effects which may also reconcile the slight discrepancy (Table~\\ref{table_syserr}). \\subsection{Baryon fractions} An interesting result from our analysis is that these systems, despite having masses \\gtsim 5$\\times 10^{12}$\\msun, do not appear in general to be baryonically closed. To some extent this trend was enforced by applying Eq~\\ref{eqn_baryons} to constrain the data. However, the excellent fits we obtained by this method, in conjunction with the good agreement between the measured \\mvir-c relation and the predictions of \\lcdm\\ and, crucially, our measurements at the group scale (which do not employ this restriction: \\citealt{gastaldello06a}), indicate that the inferred \\fbaryons ($\\sim$0.04--0.09; Table~\\ref{table_results}) are accurate. Furthermore, if we relaxed this constraint and instead restricted \\fbaryons\\ to a finite range, we also found that the data tended to favour modest values of \\fbaryons. In particular, for any given system, the measured \\mvir\\ and \\fbaryons\\ were strongly anti-correlated, so that our upper \\mvir\\ constraint is in part imposed by the {\\em lower} limit we place on \\fbaryons. Given the shapes of the \\mvir-c contours (Fig~\\ref{fig_confidence}), it is clear that good agreement with the \\mvir-c relation predicted from simulations tends, therefore, to require rather modest values of \\fbaryons. This would suggest that strong feedback plays an important role in the evolution of these objects. \\subsection{Temperature profiles} \\label{discussion_temp} By inspection of the temperature profiles (Fig~\\ref{fig_temp}) it is immediately clear that, for all of the galaxy-scale systems in our sample the temperature profiles have negative gradients. In contrast the group-scale objects have positive temperature gradients, similar to observations of other X-ray bright groups and clusters \\citep{gastaldello06a,vikhlinin05b,piffaretti05a}. This radical difference in the temperature profiles seems consistent with our division of galaxies and groups at \\mvir $\\sim 10^{13}$\\msun. The origin of this distinct demarcation between objects around $10^{13}$\\msun\\ is unclear, however. Negative temperature gradients are expected for isolated galaxies containing relatively cool gas, such as that arising from stellar mass-loss. In the deep stellar potential well, compressive heating of the gradually inflowing gas can dominate over radiative cooling to produce a negative temperature slope. In contrast, if hotter ($\\sim$1--2~keV) baryons are allowed to flow in, radiative cooling dominates to produce a positive temperature gradient \\citep{mathews03a}. It is by no means clear, however, why the hot baryons appear to be present only in the systems with \\mvir\\gtsim $10^{13}$\\msun. One possibility is the local environment; all of the galaxy-scale objects are rather isolated, whereas the groups NGC\\thin 4472 and NGC\\thin 4649, in particular, are found in a relatively dense cluster environment, which could provide a reservoir of hot baryons. However, such an explanation cannot easily account for the positive temperature gradient in NGC\\thin 1407, which is comparatively isolated, or the isolated system NGC\\thin 4555, which appears only slightly more massive than our galaxies. It is possible that selection effects may have played some role in the bimodal temperature profile behaviour, since both NGC\\thin 4125 and NGC\\thin 6482 are classified in \\ned\\ as LINERS, and NGC\\thin 720 has a dominant young stellar population (Appendix~\\ref{sect_stars}). However, none of these systems show strong X-ray morphology disturbances in the core, which might indicate a substantial energy input from star-formation or AGN activity. In any case, the cooling time in the core of NGC\\thin 720 is only $\\sim$200~Myr, substantially less than the implied time since the last major burst of star formation, and so the negative temperature gradient cannot simply be related to energy injection during a starburst. Furthermore, at least two of the group-scale systems also harbour AGN and do not show obvious negative temperature gradients in the core. Another example of an object we believe to be a galaxy (rather than a group) which exhibits a negative temperature gradient is the S0 NGC\\thin 1332 \\citep{humphrey04b}. A possible counter-example to this trend might by the ``isolated elliptical galaxy'' NGC\\thin 4555, which exhibits a temperature profile akin to the groups in our sample \\citep{osullivan04b}. However, as we discuss in \\S~\\ref{discussion_groups}, this probably has comparable \\mvir\\ to the groups. Another intriguing feature of two of the group scale objects is a central temperature peak, similar to a feature we found in the cluster A\\thin 644 \\citep{buote05a}. In that system, we found a significant offset between the X-ray centroid and the emission peak in an otherwise fairly relaxed object. We suggested that both of these features may be related to the cD ``sloshing'' in the potential well of the cluster, which is relaxing following disturbance by, for example, a merger. We do not find obvious evidence of a similar offset in either NGC\\thin 1407 or NGC\\thin 4649. However, these groups may be in a comparably more relaxed (evolved) state than A\\thin 644. Alternatively, the central peaks may be related to past AGN activity heating the gas in the core of the galaxies, from which the system has had time to relax dynamically but not cool completely. \\subsection{Is NGC\\thin 1407 a ``dark group''?} \\label{sect_ngc1407} Based on the group member dispersion velocity \\citet{gould93a} suggested that NGC\\thin 1407 may lie in a massive (\\gtsim a few $\\times 10^{13}$\\msun) DM halo. Although such a conclusion was strongly dependent on the association of the galaxy NGC\\thin 1400, which exhibits a large peculiar velocity, with the group, we can now confirm the presence of a substantial DM halo around this system. Both the temperature profile and our best-fit mass are similar to the bright X-ray group NGC\\thin 5044 \\citep{buote06a}, and yet it is almost 2 orders of magnitude fainter in \\lx. NGC\\thin 5044 appears to be close to baryonic closure \\citep{mathews05a}, and so has likely retained most of its large gaseous halo. On the other hand NGC\\thin 1407 is not baryonically closed (we estimate \\fbaryons$\\simeq$0.06) and so the loss of much of its hot gas envelope easily explains its lower \\lx/\\lb. Since the masses of the two systems are not considerably different, this points to substantial variation in the evolutionary history of these two groups. In particular, feedback may have operated more efficiently in evacuating the gas from NGC\\thin 1407. \\citeauthor{gould93a}'s preferred mass estimate ($\\sim10^{14}$\\msun) would imply a remarkably high M/L ratio for the system (\\mvir/\\lb$\\sim$900\\msun/\\lsun), making NGC\\thin 1407 a bona fide ``dark group''. The existence of such an object would provide a valuable insight into the process of star formation in DM halos, as it would imply star formation was somehow inhibited in that system. This mass estimate is, however, considerably larger than our preferred value $\\sim 1.5 \\times 10^{13}$\\msun, which implies a more modest M/L ratio (\\mvir/\\lb$\\sim$140\\msun/\\lsun). To some extent, though, our constraint on \\fbaryons, which was necessary to obtain interesting \\mvir\\ constraints, has probably enforced this behaviour. Such a restriction may not be valid in a system with an unusual star-formation history and so we experimented with freeing \\fbaryons. To enable \\mvir\\ to be constrained, we restricted c to lie on the best-fit \\mvir-c relation found by \\citet{bullock01a}. The best-fitting mass, \\mvir$=(9.7^{+17.8}_{-6.2})\\times 10^{13}$\\msun, was in good agreement with \\citet{gould93a}'s values, but implies a baryon fraction only of $\\sim$0.003. Since this fit was statistically indistinguishable from the preferred model, we cannot determine which mass estimate is more likely." }, "0601/astro-ph0601137_arXiv.txt": { "abstract": " ", "introduction": "The photometric data sets obtained in the course of large microlensing surveys provide an abundance of opportunities to carry out scientific investigations not related to lensing. In one such project, Sumi~(2004; S04 hereafter) presented extinction maps of the Galactic bulge fields observed in the Optical Gravitational Lensing Experiment (OGLE) during its second phase. These fields range in Galactic longitude between approximately $-11$~deg$$5 \\\\ CS 22956-028* & 6700 & 3.5 & -2.38 & 3 \\\\ CS 30322-023 & 4000 & 0.0 & -3.78 & 5 \\\\ HD 26* & 5200 & 2.6 & -0.71 & ? \\\\ HD 5424* & 5000 & 3.0 & -0.41 & 4 \\\\ HD 24035* & 5000 & 3.7 & 0.16 & 4 \\\\ HD 168214 & 5200 & 3.5 & -0.03 & ? \\\\ HD 187861* & 4800 & 1.8 & -2.21 & 10 \\\\ HD 196944* & 5250 & 1.7 & -2.21 & ? \\\\ HD 206983 & 4550 & 1.6 & -0.93 & 4 \\\\ HD 207585* & 5800 & 4.0 & -0.35 & 10 \\\\ HD 211173 & 4800 & 2.5 & -0.17 & 13 \\\\ HD 218875 & 4600 & 1.5 & -0.63 & 100 \\\\ HD 219116 & 4800 & 1.8 & -0.45 & 7 \\\\ HD 224959* & 5000 & 2.0 & -2.08 & 7 \\\\ HE 1419-1324 & 4900 & 1.8 & -3.28 & 4 \\\\ HE 1410+0213* & 4550 & 1.0 & -2.38 & 4 \\\\ HE 1001-0243* & 5000 & 2.3 & -3.12 & 30 \\\\ \\hline \\multicolumn{5}{l}{* probably binary, from radial velocity monitoring}\\\\ \\end{tabular} \\end{center} \\end{table} ", "conclusions": "" }, "0601/astro-ph0601409_arXiv.txt": { "abstract": "We have obtained optical and infrared photometry of the quiescent soft X-ray transient XTE J1118+480. In addition to optical and $J$-band variations, we present the first observed $H$- and $K_s$-band ellipsoidal variations for this system. We model the variations in all bands simultaneously with the WD98 light curve modeling code. The infrared colors of the secondary star in this system are consistent with a K7V, while there is evidence for light from the accretion disk in the optical. Combining the models with the observed spectral energy distribution of the system, the most likely value for the orbital inclination angle is $68^{\\circ}\\pm2^{\\circ}$. This inclination angle corresponds to a primary black hole mass of $8.53\\pm0.60 \\,M_{\\odot}$. Based on the derived physical parameters and infrared colors of the system, we determine a distance of 1.72$\\pm$0.10 kpc to XTE J1118+480. ", "introduction": "Transient low mass X-ray binaries (LMXBs) exhibit large and abrupt X-ray and optical outbursts that can be separated by decades of quiescence \\citep{csl97}. For these systems, the compact object is a black hole or a neutron star, and the companion is normally a low-mass K- or M-type dwarf-like star \\citep{cc03}. During their periods of quiescence, these systems are faint at X-ray, optical, and infrared (IR) wavelengths. While the quiescent X-ray emission can be caused by accretion onto a compact object or thermal emission from the surface of a neutron star \\citep{garcia01,wijnands04}, the companion can dominate the luminosity at optical and IR energies. During the binary orbit, the changing aspect of the tidally-distorted companion causes a periodic modulation of the optical and IR emission \\citep{gelino01}. Measurements of these ``ellipsoidal variations'' provide information about the physical parameters of the binary. XTE J1118+480 ($\\alpha_{2000}$ = 11$^{\\rm h}$18$^{\\rm m}$10.85$^{\\rm s}$, $\\delta_{2000}$ = 48$^{\\circ}$02$\\arcmin$12.9$\\arcsec$) was discovered with the all-sky monitor (ASM) on the {\\it Rossi X-Ray Timing Explorer} by \\citet{rem00} on 2000 March 29, while \\citet{gar00} spectroscopically identified its 12.9 magnitude optical counterpart. Owing to its position in the Galactic halo, this high latitude ({\\it b} = +62$^{\\circ}$) system has been observed by numerous groups over many wavelength regimes. Recent orbital parameters determined from optical spectra suggest an orbital period of 4.078 hr and a secondary star radial velocity semi-amplitude of 709$\\pm$7 km s$^{-1}$ \\citep{tor04}. These values imply a mass function of {\\it f(M)}=6.3$\\pm$0.2 M$_{\\odot}$, identifying the compact object as a black hole. Determining a precise black hole mass requires an accurate measurement of the orbital inclination angle of the system. As discussed in \\citet{gho01}, the best way to find the inclination angle in a non-eclipsing system is to model its infrared ellipsoidal light curves. In the IR regime, there is a smaller chance of contamination from other sources of light in the system. While modeling several light curves from one wavelength regime helps to constrain model parameters, simultaneously modeling light curves that span more than one wavelength regime provides tighter constraints than modeling IR light curves alone. Previous inclination estimates for XTE J1118+480 have come from modeling optical ellipsoidal variations as the system approached quiescence \\citep{mcc01, wag01, zur02}. These inclination angles range from 55$^{\\circ}$\\ \\citep{mcc01} to 83$^{\\circ}$\\ \\citep{wag01}, and correspond to primary masses of 10 M$_{\\odot}$ and 6.0 M$_{\\odot}$, respectively. Since this system has been known to exhibit optical superhumps from the precession of an eccentric accretion disk on its way to quiescence \\citep{zur02}, it is important to determine the orbital inclination angle while XTE J1118+480 is in a truly quiescent state. In order to determine an accurate orbital inclination angle for XTE J1118+480, we have obtained $B$-, $V$-, $R$-, $J$-, $H$-, and $K_s$-band light curves of the system while in quiescence, and simultaneously model them here with the WD98 light curve modeling code \\citep{wil98}. To date, this is the most comprehensive ellipsoidal variation data set published for this system. The modeled inclination angle is combined with recently published orbital parameters to determine a highly constrained mass of the black hole in this X-ray binary. ", "conclusions": "Figure \\ref{fig1} presents the resulting $B$-, $V$-, and $R$-band light curves of XTE J1118+480, while the final $J$, $H$, and $K_s$ differential light curves are presented in Figure \\ref{fig2}. Despite the expectation of detecting superhumps from the 52 day accretion disk precession period, there was no evidence of a superhump period when the data were run through a periodogram. These results are consistent with the findings of \\citet{sha05} whose optical data taken on 2003 June suggested that if a superhump modulation existed, it was at the $<$0.50$\\%$ level. \\citet{sha05} also observed stochastic variability and fast flares in their light curves. However, the observed stochastic variability had the same magnitude as their photometric error bars, and despite the significant power in the power density spectrum, the flares seen in XTE~J1118+480 had roughly the same magnitude as the spread in the comparison star measurements. The 2003 and 2004 $V$- and $R$-band light curves presented here were consistent in shape and amplitude. As Table \\ref{obstab} shows, two thirds of the nights that XTE~J1118+480 was observed, data was gathered for at least 75\\% of an orbit, with more than one full orbit covered on four of the nine nights. No evidence for unequal maxima, flares, or light curve distortions were found in the unphased data. The periodogram results are consistent with the results from \\citet{tor04} and therefore all data presented here have been phased to their ephemeris. These are the first $H$- and $K_s$-band detections of ellipsoidal variations from this X-ray binary system. \\subsection{Ellipsoidal Models} The spectral type of the secondary star can be estimated by comparing its red optical spectrum with the spectra of stars with various spectral types from the same luminosity class. Alternatively, one can use a spectral energy distribution (SED), and published limits on the spectral type to not only derive an effective temperature of the secondary star, but to also estimate both the visual extinction and contamination level. Given that photometric data usually has a higher S/N than spectroscopic data, an effective temperature derived using photometry can be just as useful as a spectral type derived from a spectroscopic data set. To this end, we compared the observed optical/IR ($BVRJHK_s$) SED for XTE J1118+480 with the observed SEDs for K0V -- M4V stars \\citep{bb88,bes91,mik82}. We present the best fitting SED in Figure \\ref{fig3}, and find that it predicts a visual extinction of A$_V$=0.065$\\pm$0.020 mags, and a secondary star spectral type of K7V. It also includes 60 -- 70\\% light from the accretion disk at $B$ and 30 -- 35\\% at $V$. A K5V gives a slightly worse fit, and predicts more $R$-band light than is observed, as well as a smaller amount of disk light in the $B$- and $V$-bands, inconsistent with previously published values. The visual extinction found here is consistent with the column density adopted by \\citet[$N_H=1.2 \\times 10^{20}$ cm$^{-2}$]{mcc04} based on three independent measurements, and the spectral type found here is consistent with those found through spectral fitting \\citep[K5/7 V]{mcc03,tor04}. Light from the primary object or accretion disk in an X-ray binary will act to dilute the amplitude of the ellipsoidal variations of the secondary star. \\citet{tor04} estimate that the secondary star in the XTE J1118+480 system contributes roughly 55\\% of the total flux between 5800\\AA ~and 6400\\AA; however, at $R$ = 18.6, the system was not in true quiescence when their observations were taken. \\citet{zuriauc02} determined that XTE J1118+480 has a true quiescent $R$-band magnitude of 18.9. % The optical data presented here ($R$=19.00) are consistent with this value, and thus we contend that the system was in a quiescent state during our observations. In addition, Doppler imaging of the system did not detect any H$\\alpha$ emission from a hot spot or accretion stream \\citep{tor04}. Since our observations took place while XTE J1118+480 was in true quiescence, the optical data presented here are consistent with both of the \\citet{tor04} results. The most difficult contamination source to extract in the case of XTE J1118+480 is the one that has the shallowest spectral slope. If the disk contamination in the infrared is based on the assumption that the optically thin disk radiates through free-free emission processes, and we therefore ascribe the entire $V$-band excess to free-free emission, then in the $K_s$-band, the contamination would be $\\sim$ 8\\%. An 8\\% contamination in the infrared bands would cause the observed orbital inclination angle of the system to be underestimated by 2$^{\\circ}$. What if we instead assume that any IR contamination is ascribed to a jet or some form of an advection dominated accretion flow (ADAF)? Models by \\citet{yua05} for XTE~J1118+480 in quiescence show that both the jet and ADAF flux in the IR are predicted to constitute significantly less than 8\\% of the companion flux. Irradiation of the secondary star by the accretion disk will affect the symmetry of the ellipsoidal light curves. \\citet{tor04} investigated the possibility of X-ray irradiation powering the H$\\alpha$ emission seen in their Doppler tomograms of the XTE J1118+480 system. They concluded that it was improbable that this could be the case, and found that the strength of the H$\\alpha$ emission from the secondary star was comparable with other K dwarfs. Similarly, we see no significant evidence for irradiation effects in the unphased data or the light curves presented here. As in \\citet{gho01}, we simultaneously modeled the optical and infrared light curves of XTE J1118+480 with WD98 \\citep{wil98}. See \\citet{gel01} for references and a basic description of the code, and \\citet{gelino01} for a more comprehensive description. The data were run through WD98's DC program which uses a Simplex algorithm for initial parameter searches and a damped least squares algorithm for error minimization between the data and model light curves. We ran the code for a semi-detached binary with the primary component's gravitational potential set so that all of its mass is concentrated at a point. The most important wavelength-independent input values to WD98 are listed in Table~\\ref{tab1}. The models were run for a range of inclination angles with parameters for a K3V through an M1V secondary star. The secondary star atmosphere was determined from solar-metallicity Kurucz models. We used normal, non-irradiated, square-root limb darkening coefficients \\citep{van93}. We also adopted gravity darkening exponents found by \\citet{cla00}, and mass ratios ranging from $q$ = 0.030 - 0.056 \\citep{har99,oro01}. We assumed that the secondary star was not exhibiting any star spots during our observations. The models were run with varying amounts of additional light in all bands. Solving for six consistent light curve solutions simultaneously allowed the rejection of many disk light scenarios. With 118 degrees of freedom, the best fit model had a reduced $\\chi^2$ of 1.65. While the error on the best fit inclination angle was dependent on combining the uncertainties from varying all of the model parameters simultaneously, we found that changing the spectral type of the secondary from a K5V to an M1V resulted in a change in the orbital inclination angle of 1$^{\\circ}$. Similarly, varying $q$ from 0.030 to 0.056 affected $i$ by $\\le 1^{\\circ}$. We find that the best fitting $B$-, $V$-, $R$-, $J$-, $H$-, and $K_s$-band model has $i$ = 68$^{\\circ}$, and the parameters found in Table \\ref{tab2}. Figure 1 presents this model for the optical bands, while Figure \\ref{fig2} presents this model for the $J$-, $H$-, and $K_s$-band light curves. \\subsection{The Model and its Uncertainties} Based on our optical/IR SED, as well as published values, we adopt a secondary star spectral type of K7 \\citep{tor04, mcc01} with a temperature of $T_{\\rm eff}$ = 4250 K \\citep{gra92}. The corresponding gravity darkening exponent is $\\beta_1$ = 0.34. Modeling six light curves simultaneously is a robust method for constraining the amount of extra light in the system. If we model the light curves individually and assume that all of the light in the system comes from the secondary star, the best fitting orbital inclination angle for XTE J1118+480 varies with wavelength. Since the contaminating light does not have the same spectrum as the secondary star the light curves at each wavelength are diluted by different amounts, causing the best fit inclination for the affected bands to be different from the others. This is the case for the $B$- and $V$-band light curves. If a jet or other contaminant were to have a flat spectrum that only affected the IR, the results from the SED would most likely not match those obtained from spectral measurements, and we would not be able to determine a reasonable parameter set and inclination angle that is consistent throughout the data. When fit simultaneously, the best fit inclination angle is $i$ = 68$^{\\circ}$ with 62\\% disk light in $B$, and 31\\% disk light in $V$. These disk light contributions are consistent with those found through the SED fitting. Using an estimate of 8\\% for the infrared accretion disk contamination in the system gives an inclination of 68$^{+2.8}_{-2}$$^{\\circ}$, however, if we artificially add in 8\\% or more infrared light, we are unable to obtain a reasonable solution for the orbital inclination angle that is consistent throughout the entire data set. Therefore, based on the IR colors of the system ($J$-$K$=1.1$\\pm$0.3), the SED fit, and the results of the simultaneous light curve modeling, it appears unlikely that the infrared light curves are significantly affected by any such contamination. In order to determine the final error on the orbital inclination angle, we plotted the $\\chi^2$ values as a function of $i$. Therefore, based on the error in each of the model parameters including $q$, the spectral type (i.e. temperature) of the secondary star, the amount of observed disk light, as well as the photometric error bars, the orbital inclination angle is 68$^{\\circ}$$\\pm$2$^{\\circ}$. We combined the determined inclination angle with the orbital period (P = 0.1699167 $\\pm$ 1.72$\\times$10$^{-5}$ d), radial velocity of the secondary star (K$_2$ = 709 $\\pm$ 7 km; \\citet{tor04}), and and the mass ratio ($q$ = 0.0435 $\\pm$ 0.0100) to find the mass of the primary object. A Monte Carlo routine was used to propagate the errors on the above quantities and gives a primary mass of 8.53$\\pm$0.60 M$_{\\odot}$, confirming it as a black hole. The constraints on the mass of the black hole in this system presented here represent a considerable improvement over those previously published. \\citet{wag01} determined a mass of 6.0 -- 7.7 M$_{\\odot}$ from data obtained before XTE 1118+480 had entered a quiescent state ($R\\sim$18.3). \\citet{mcc01} gave an upper limit of $M_1 \\le$ 10 M$_{\\odot}$, and more recently, \\citet{mcc04} adopted a mass of $\\sim$ 8 M$_{\\odot}$ for their calculations of the thermal emission from the black hole in the system. The 8.53 M$_{\\odot}$ black hole mass determined here is consistent with the adopted mass of \\citet{mcc04}, and falls nicely into the current observed black hole mass distribution. Theoretical models by \\citet{fry01} predict that there should exist a greater number of 3 -- 5 M$_{\\odot}$ black holes than 5 -- 12 M$_{\\odot}$ black holes; however, most of the determined black hole masses thus far have fallen into a 6 -- 14 M$_{\\odot}$ range. In fact, thus far, GRO J0422+32 is the only system with a compact object that falls into the 3 -- 5 M$_{\\odot}$ range \\citep[3.97$\\pm$0.95 M$_{\\odot}$]{gel03}. Using the mass of the compact object and the orbital period, we computed the orbital separation of the two components in the system. We then combined the separation with the mass ratio to find the size of the Roche lobe for the secondary star. The temperature of the secondary and its Roche lobe radius were then used to find the secondary's bolometric luminosity and bolometric absolute magnitude. After accounting for the bolometric correction \\citep{bes91}, the distance modulus for the $J$, $H$, and $K_s$ bands were used to find an average distance of 1.72$\\pm$0.10 kpc. Consistent with results from the modeling and SED fitting, this calculation assumes that all of the IR light in the system originates from the secondary star. Table~\\ref{tab2} lists all of the derived parameters for the XTE~J1118+480 system. Both the mass and radius of the secondary star are smaller than that of a K7V ZAMS star. In addition, the infrared colors of the system and its position in the Galactic halo, support the notion that the secondary star in the XTE J1118+480 system may be evolved." }, "0601/astro-ph0601315_arXiv.txt": { "abstract": "We have carried out large scale CO observations with a mm/sub-mm telescope NANTEN toward a far infrared loop-like structure whose angular extent is about 20$\\times$20 degrees around ($l$, $b$) $\\sim$ (109$\\degr$, $-$45$\\degr$) in Pegasus. Its diameter corresponds to $\\sim$ 25 pc at a distance of 100 pc, adopted from that of a star HD886 (B2IV) near the center of the loop. We covered the loop-like structure in the $^{12}$CO ($J$ = 1--0) emission at 4$\\arcmin$--8$\\arcmin$ grid spacing and in the $^{13}$CO ($J$ = 1--0) emission at 2$\\arcmin$ grid spacing for the $^{12}$CO emitting regions. The $^{12}$CO distribution is found to consist of 78 small clumpy clouds whose masses range from 0.04 $M_{\\sun}$ to 11 $M_{\\sun}$, and $\\sim$ 83\\% of the $^{12}$CO clouds have very small masses less than 1.0 $M_{\\sun}$. $^{13}$CO observations revealed that 18 of the 78 $^{12}$CO clouds show significant $^{13}$CO emission. $^{13}$CO emission was detected in the region where the column density of H$_{2}$ derived from $^{12}$CO is greater than 5$\\times$10$^{20}$ cm$^{-2}$, corresponding to $A$v of $\\sim$ 1 mag, which takes into account that of H{\\small \\,I}. We find no indication of star formation in these clouds in IRAS Point Source Catalog and 2MASS Point Source Catalog. The very low mass clouds, $M$ $\\leq$ 1$M_{\\sun}$, identified are unusual in the sense that they have very weak $^{12}$CO peak temperature of 0.5 K--2.7 K and that they aggregate in a region of a few pc with no main massive clouds; contrarily to this, similar low mass clouds less than 1 $M_{\\sun}$ in other regions previously observed including those at high Galactic latitude are all associated with more massive main clouds of $\\sim$ 100 $M_{\\sun}$. A comparison with a theoretical work on molecular cloud formation (Koyama \\& Inutsuka 2002) suggests that the very low-mass clouds may have been formed in the shocked layer through the thermal instability. The star HD886 (B2IV) may be the source of the mechanical luminosity via stellar winds to create shocks, forming the loop-like structure where the very low-mass clouds are embedded. ", "introduction": "High Galactic latitude molecular clouds (hereafter HLCs) are typically located at $\\mid$b$\\mid$ $\\gtrsim$ 20$\\degr$--30$\\degr$. Since the Gaussiun scale height of CO is estimated to be $\\sim$ 100 pc in the inner Galactic disk (e.g., Magnani et al. 2000), HLCs are likely located very close to the Sun, within a few hundred pc or less. Their proximity to the Sun and the low possibility of overlapping with other objects along the line of sight enable us to study them with a high spatial resolution and to compare CO data unambiguously with the data at other wavelengths. HLCs have lower molecular densities compared with dark clouds where the optical obscuration is significant. Therefore, HLCs are often called as translucent clouds (e.g., van Dishoeck \\& Black 1988) and most of the known HLCs are not the sites of active star formation, although a few of them are known to be associated with T Tauri stars (e.g., Magnani et al. 1995; Pound 1996; Hearty et al. 1999). Given the very small distances of HLCs, it is a challenging task for observers to make a complete survey for HLCs over a significant portion of the whole sky. $^{12}$CO ($J$ = 1--0) emission has been used to search for HLCs because the line emission in the mm band is strongest among the thermally or sub-thermally excited spectral lines of interstellar molecular species. It is however difficult to cover an area as large as tens of square degrees subtended by some of the HLCs because of the general weakness of the $^{12}$CO emission, typically $\\sim$ a few K (e.g., Magnani et al. 1996), with existing mm-wave telescopes in a reasonable time scale. HLCs have been therefore searched for by employing various large-scale datasets at other wavelengths including the optical obscuration (Magnani et al. 1985; Keto \\& Myers 1986), the infrared radiation (Reach et al. 1994), and the far-infrared excess over H{\\small \\,I} (=FIR excess)(Blitz et al. 1990; Onishi et al. 2001). On the other hand, unbiased surveys in CO at high Galactic latitudes have been performed at very coarse grid separations of 1$\\degr$ resulting in a small sampling factor of a few \\% (Hartmann et al. 1998; Magnani et al. 2000). Most recently, Onishi et al. (2001) discovered 32 HLCs or HLC complexes. This search was made based on the FIR excess, demonstrating the correlation among FIR excess clouds with CO clouds is a useful indicator of CO HLCs. Previous CO observations of individual HLCs at higher angular resolutions show that HLCs exhibit often loop-like or shell-like distributions having filamentary features with widths of several arc min or less (Hartmann et al. 1998; Magnani et al. 2000; Bhatt 2000), and in addition that HLCs often compose a group, whose angular extent is $\\sim$ 10 degrees or larger. In order to better understand the structure of HLCs and to pursue the evolution of HLC complexes, CO observations covering tens of square degrees at a high angular resolution are therefore crucial. The past observations of such complexes of HLCs are limited to a few regions including Polaris flare (Heithausen \\& Thaddeus 1990), Ursa Major (Pound \\& Goodman 1997) and the HLC complex toward MBM 53, 54, and 55 (Yamamoto et al. 2003). Pound \\& Goodman (1997) showed an arc-like structure of the molecular cloud system and suggested that the origin of such structures could be some explosive events. Most recently, Yamamoto et al. (2003) carried out extensive observations of the molecular cloud complex including MBM 53, 54, and 55 and suggest that the HLCs may be significantly affected by past explosive events based on the arc-like morphologies of molecular hydrogen (see also Gir et al. 1994). The region of MBM 53, 54, and 55 is of particular interest among the three, because it is associated with a large H{\\small \\,I} cloud of $\\sim$ 590 $M_{\\sun}$ at a latitude of $-$35 degrees and because there is a newly discovered HLC of 330$M_{\\sun}$, HLCG92$-$35, which is significantly H{\\small \\,I} rich with a mass ratio $M$(H$_{2}$)/$M$(H{\\small \\,I}) of $\\sim$ 1, among the known HLCs (Yamamoto et al. 2003). This cloud was in fact missed in the previous surveys based on optical extinction (Magnani et al. 1985). Subsequent to these observations we became aware of that the region is also very rich in interstellar matter as shown by the 100$\\mu$m dust features (Kiss et al. 2004). There is a loop-like structure shown at 100 $\\mu$m around ($l$, $b$) $\\sim$ (109$\\degr$, $-$45$\\degr$). Toward the center of the loop, an early type star HD886(B2IV) is located and may play a role in creating the loop. Its proper motion is large at a velocity of a few km s$^{-1}$, suggesting that the stellar winds of the star might have continued to interact with the surrounding neutral matter over a few tens of pc in $\\sim$ a few Myr. Magnani et al. (1985) and Onishi et al. (2001) yet observed only a small part of this region. In order to reveal the large scale CO distribution of the region, we have carried out observations toward ($l$, $b$) $\\sim$ (109$\\degr$, $-$45$\\degr$) by $^{12}$CO ($J$ = 1--0) and $^{13}$CO ($J$ = 1--0) with NANTEN 4-meter millimeter/sub-mm telescope of Nagoya University at Las Campanas, Chile. We shall adopt the distance of 100 pc from the sun to the loop-like structure which is equal to the distance of the B2 star in the center of the loop, and is also a typical value for the HLCs. ", "conclusions": "We have made a large-scale survey of high Galactic latitude molecular clouds in the $J$ = 1--0 lines of $^{12}$CO and $^{13}$CO toward a large scale structure located around ($l$, $b$) $\\sim$ (109$\\degr$, $-$45$\\degr$) with NANTEN. This survey spatially resolved the distribution of molecular gas associated with the large scale structure. The main conclusions of the present study are summarized as follows: \\begin{enumerate} \\item The $^{12}$CO observation covered the entire large loop-like structure. The loop-like structure consits of very small clumpy clouds. The $^{12}$CO clouds are concentrated on the north to north-west of the loop-like structure and toward the south of that. We identified 78 $^{12}$CO clouds in the observed region. The total mass is estimated to be $\\sim$ 64 $M_{\\sun}$ if we assume the conversion factor from CO intensity to $N$(H$_2$) as 1.0$\\times$10$^{20}$ cm$^{-2}$/(K km s$^{-1}$). \\item We performed $^{13}$CO observations in and around the whole area where the peak temperature of $^{12}$CO is more than 2.0 K. We identified 33 $^{13}$CO clouds and derived physical properties under the assumption of LTE. \\item The mass spectra are well fitted by a power law, $dN/dM$ $\\propto$ $M^{-1.53\\pm0.13}$ for the $^{12}$CO clouds and $dN/dM$ $\\propto$ $M^{-1.36\\pm0.10}$ for the $^{13}$CO clouds. These spectral indices are similar to those derived in the other regions. \\item The size and the line width relation of $^{13}$CO clouds is fitted by a least-squares method, log($\\Delta V$) = (0.22$\\pm$0.43) $\\times$ log($R$) + (0.37$\\pm$0.52) (c.c.=0.23), but the correlation is not good. \\item Present $^{13}$CO clouds are far from the virial equilibrium, indicating that $^{13}$CO clouds are not gravitationally bound. $M_{\\rm vir}$ and $M_{\\rm LTE}$ relation can be fitted by a least-squares method as log($M_{\\rm vir}$) = (0.91$\\pm$0.30) $\\times$ log($M_{\\rm LTE}$) + (2.23$\\pm$0.29) (c.c.=0.66). This index is slightly different from the indices in the other regions although the tendency that molecular clouds are more vilialized as the mass increases is consistent with the other regions. \\item There is no sign of star formation from the comparison of IRAS point sources and Point Source Catalog of Two-Micron All-Sky survey in the present region. This suggests that molecular clouds in this region are not the site of present star formation or the remnants of past star formation. \\item There may be two expanding shells in the present region as inferred from H{\\small \\,I} although we cannot identify them from CO. The total mechanical luminosity of HD886 during the last few $\\times$ 10$^{6}$ yr is comparable to the expanding energy of the northern expanding H{\\small \\,I} shell. This indicates that some additional source of energy other than HD886 is needed to explain the expanding energy. \\item $^{13}$CO emission is significantly detected in the $^{12}$CO clouds having molecular column density greater than 5$\\times$10$^{20}$ cm$^{-2}$. This may be explained as that the $^{13}$CO emitting regions become significant when $A$v becomes larger than $\\sim$ 1 mag, marginally enough to shield the ultraviolet radiation to protect $^{13}$CO molecules. \\item There is a possibility that very small clouds have been formed in the shoked layer through the thermal instability. The stellar wind of HD886 may be the source to creat shocks, forming the loop-like structure where the very small clouds are embedded. \\end{enumerate}" }, "0601/astro-ph0601645_arXiv.txt": { "abstract": "\\par We present new rates for the \\iso{22}Ne$(\\alpha, n$)\\iso{25}Mg and \\iso{22}Ne$(\\alpha,\\gamma)$\\iso{26}Mg reactions, with uncertainties that have been considerably reduced compared to previous estimates, and we study how these new rates affect the production of the heavy magnesium isotopes in models of intermediate mass Asymptotic Giant Branch (AGB) stars of different initial compositions. All the models have deep third dredge-up, hot bottom burning and mass loss. Calculations have been performed using the two most commonly used estimates of the \\iso{22}Ne $+ \\, \\alpha$ rates as well as the new recommended rates, and with combinations of their upper and lower limits. The main result of the present study is that with the new rates, uncertainties on the production of isotopes from Mg to P coming from the \\iso{22}Ne $+ \\alpha$-capture rates have been considerably reduced. We have therefore removed one of the important sources of uncertainty to effect models of AGB stars. We have studied the effects of varying the mass-loss rate on nucleosynthesis and discuss other uncertainties related to the physics employed in the computation of stellar structure, such as the modeling of convection, the inclusion of a partial mixing zone and the definition of convective borders. These uncertainties are found to be much larger than those coming from \\iso{22}Ne $+ \\, \\alpha$-capture rates, when using our new estimates. Much effort is needed to improve the situation for AGB models. ", "introduction": "\\par The origin of the stable magnesium isotopes, \\iso{24}Mg, \\iso{25}Mg and \\iso{26}Mg, is of particular interest to astrophysics because Mg is one of the few elements for which we can obtain isotopic information from stellar spectroscopy. The ratio \\iso{24}Mg:\\iso{25}Mg:\\iso{26}Mg has been derived from high-resolution spectra of cool dwarfs and giants in the thin and thick disk of the Galaxy \\citep*{gl00,yong03b}, and for giants stars in the globular cluster (GC) NGC 6752 \\citep{yong03a}. These observations show that many of the stars, including relatively metal-poor stars ([Fe/H] $\\lesssim -1.0$), have non-solar Mg isotopic ratios\\footnote{The solar Mg isotopic ratios are \\iso{24}Mg:\\iso{25}Mg:\\iso{26}Mg = 79:10:11 \\citep{lodder03}.} with enhancements in the neutron-rich isotopes, \\iso{25}Mg and \\iso{26}Mg, compared to what is expected from galactic chemical evolution (GCE) models. The main stellar nucleosynthesis site for all three stable isotopes is hydrostatic burning in the carbon and neon shells of massive stars that explode as Type II supernovae \\citep{ww95,cl04}. The abundances of the neutron-rich Mg isotopes are further enhanced by secondary $\\alpha$-capture processes operating in the helium shell. The amount of \\iso{24}Mg produced does not strongly depend on the initial metallicity of the model and is an example of primary nucleosynthesis\\footnote{Primary production means that the species is produced from the hydrogen and helium initially present in the star and the amount produced is relatively independent of the metallicity, $Z$, whereas secondary production requires some heavier seed nuclei to be present, and the amount produced scales with $Z$.}, whereas the amounts of \\iso{25}Mg and \\iso{26}Mg produced scale with the initial metallicity of the star. This means that the Mg content of the ejecta of low metallicity supernovae will be mostly \\iso{24}Mg with very little \\iso{25}Mg and \\iso{26}Mg produced, with typical ratios \\iso{24}Mg:\\iso{25}Mg:\\iso{26}Mg $\\approx$ 99.0:0.50:0.50 from a 25$\\Msun$ supernova model with metallicity $Z=0.01 Z_{\\odot}$ \\citep{ww95}. \\par Previous studies have shown that GCE models using ejecta from massive stars match the observational data well for [Fe/H] $> -1.0$ but severely underestimate \\iso{25,26}Mg/\\iso{24}Mg at lower metallicities \\citep*{timmes95}, indicating that another production site for the neutron-rich Mg isotopes at low metallicities is required to account for the observations. Recently, \\citet{fenner03} included the predicted stellar yields of \\iso{25}Mg and \\iso{26}Mg from Asymptotic Giant Branch (AGB) stars \\citep{kara03} along with yields from Type II supernovae \\citep{ww95,lc02} into a GCE model of the solar neighborhood. The GCE model with the AGB contribution could successfully match the Mg isotopic ratios of the metal-poor Galactic disk stars while the model without an AGB contribution could not. This result indicates that low-metallicity intermediate mass AGB stars may play an important role in the production of these species in galaxies and stellar systems. The production of the neutron-rich isotopes in AGB stars is also of interest in relation to the non-solar Mg isotopic ratios observed in giant stars in globular clusters \\citep{yong03a,yong05}. The non-solar Mg isotopic ratios observed in NGC 6752 have been attributed to AGB stars, but \\citet{fenner04} used a GCE model with tailor made AGB yields from \\citet{campbell05} and failed to match the abundance patterns observed in stars in that cluster. As \\citet{ventura05a,ventura05b} have pointed out, there are still many uncertainties that effect the stellar yields, so an AGB solution to the globular cluster anomalies cannot be ruled out at present. \\par Further motivation for the study of the production of the Mg isotopes in AGB stars is given by their relevance in the important current debate on the apparent variation of the fine-structure constant \\citep*{murphy01,ashen04,fenner05}, and the origin of pre-solar spinel grains, some of which show enhancements in both \\iso{25}Mg and \\iso{26}Mg compared to solar \\citep{zinner05}. \\par Briefly, intermediate-mass stars (initial mass $\\sim$4 to 8$\\Msun$) will enter the thermally-pulsing phase with a hydrogen (H)-exhausted core mass (hereafter core mass) $M_{\\rm c} \\gtrsim$ 0.8$\\Msun$, after experiencing the second dredge-up (SDU) \\citep{latt96,bgw99,herwig05}. The SDU brings the products of H-burning to the stellar surface (mostly \\iso{4}He and \\iso{14}N), and will slightly alter the composition of the Mg isotopic ratios, with an enrichment in \\iso{26}Mg at the expense of \\iso{25}Mg. For the massive, $Z = 0.0001$ models, the SDU is the first time the surface abundances are altered, because there is essentially no first giant branch phase \\citep{herwig04a}. Following the SDU, the He-burning shell becomes thermally unstable and flashes every few thousand years or so. The energy from the thermal pulse (TP) drives a convective pocket in the He-rich intershell, which thoroughly mixes the products of He-nucleosynthesis within this region. Following the TP, the convective envelope moves inward in mass and may reach the region previously mixed by the flash-driven convective pocket. This mixing event is known as the third dredge-up (TDU) and if it occurs, is responsible for enriching the envelope in material from the H-exhausted core. Following the TDU the star contracts and the H-shell is re-ignited and provides nearly all of the surface luminosity for the next interpulse period. The thermal pulse -- TDU -- interpulse cycle may occur many times during the TP-AGB phase; how many times depends on a number of factors including the convective model which determines the surface luminosity and mass-loss rate and hence the total AGB lifetime \\citep{ventura05a}. \\par Hot bottom burning can also occur when the base of the convective envelope becomes hot enough to sustain proton-capture nucleosynthesis \\citep{latt96}. If temperature at the base of the envelope is sufficiently hot (over $\\sim 60 \\times 10^{6}$\\,K), the NeNa and MgAl chains may operate alongside the CNO cycle; \\iso{7}Li production is also possible via the Cameron-Fowler mechanism \\citep{sb92,latt96}, which operates at lower temperatures (typically $\\sim 30-40 \\times 10^{6}\\,$K). \\citet{frost98} noted that intermediate-mass AGB stars may become luminous, optically obscured carbon stars near the end of the TP-AGB, when mass loss has removed much of the envelope, extinguishing HBB but allowing dredge-up to continue. \\par \\citet{kara03} described in detail the various nucleosynthesis processes that alter the Mg isotopic ratios in AGB stars. To summarize, \\iso{25}Mg and \\iso{26}Mg are synthesized in the He-shell during thermal pulses by the reactions \\iso{22}Ne($\\alpha,n$)\\iso{25}Mg and \\iso{22}Ne($\\alpha,\\gamma$)\\iso{26}Mg, when the temperature exceeds about 300$\\times 10^{6}\\,$K. The amount of Mg produced depends on the thermodynamic conditions inside the pulse as well as on the composition of the intershell, which will have been altered by previous H and He-burning. Neutron captures, in particular the \\iso{25}Mg($n, \\gamma$)\\iso{26}Mg reaction, can also alter the Mg isotopic ratio in the intershell, where the neutrons come from the \\iso{22}Ne($\\alpha, n$)\\iso{25}Mg reaction \\citep{herwig04b}. HBB can also significantly alter the surface Mg isotopic ratio via the activation of the MgAl chain, which can result in the destruction of \\iso{24}Mg if the temperature exceeds $\\sim 90 \\times 10^{6}\\,$K. \\par The stellar yields of \\iso{25,26}Mg presented in \\citet{kara03} and shown in figure~\\ref{fig:mg-previous} were calculated from models covering a range in mass (1$\\Msun$ to 6$\\Msun$) and metallicity ([Fe/H] $= 0, -0.3, -0.7$). From this figure we see that the most massive AGB models produce the most \\iso{25,26}Mg, as a consequence of higher temperatures in the He-shells compared to lower mass stars; we also notice an increase in production at a given mass with a decrease in metallicity. The computationally demanding nature of AGB models precluded a detailed study in that paper of the effect of the major uncertainties (mass loss, nuclear reaction rates, convection). A recent comprehensive study by \\citet{ventura05a,ventura05b} demonstrated that the predictive power of AGB models is still seriously undermined by these uncertainties. The theory of convection has a significant effect on the structure and nucleosynthesis \\citep{ventura05a}, whilst varying the mass-loss rate results in larger changes to the stellar yields than varying the nuclear reaction rates \\citep{ventura05b}. However, the magnitude of the errors associated with the relevant nuclear reaction rates is still one of the key questions concerning predictions of magnesium production in intermediate-mass AGB models. Whilst the rates of the \\iso{14}N($\\alpha,\\gamma$)\\iso{18}F and \\iso{18}O($\\alpha,\\gamma$)\\iso{22}Ne reactions are well determined \\citep{gorres00,dab03}, the two key $\\alpha$-capture reaction rates: \\iso{22}Ne($\\alpha,n$)\\iso{25}Mg and \\iso{22}Ne($\\alpha,\\gamma$)\\iso{26}Mg, suffer from large uncertainties at the stellar energies appropriate for AGB stars \\citep{koehler}. For example, at typical He-shell burning temperatures, $T \\approx 300 \\times 10^{6}\\,$K, the NACRE compilation \\citep{nacre} give an upper limit to the \\iso{22}Ne($\\alpha,n$)\\iso{25}Mg reaction rate that is about 47 times larger than the recommended rate. At lower temperatures, the uncertainties are even larger. \\par The aims of this paper are two fold. First, we present new reaction rates for the two key $\\alpha$-capture reactions, with considerably reduced uncertainties compared to those given by the NACRE compilation. Second, we use these new rates within models of different metallicities ([Fe/H] $\\approx 0, -0.3, -0.7, -2.3$), for a typical mass (5$\\Msun$) that produces the Mg isotopes during the thermally-pulsing AGB (TP-AGB) phase. For each model we calculate the stellar yields and compare to previous nucleosynthesis calculations using older estimates of these reaction rates, including those by \\citet{herwig04b} and \\citet{ventura05a,ventura05b}. We also examine the effect of other model uncertainties, in particular the inclusion of a partial mixing zone and mass loss, on the stellar yields. \\par The paper is organized as follows. In \\S\\ref{section:method} we discuss the numerical method used for the stellar model calculations, including a discussion of the input physics used. In \\S\\ref{section:rates} we present new rates for the \\iso{22}Ne $+ \\alpha$ capture reactions. The results of the calculations are presented in \\S\\ref{section:results} and discussed in \\S\\ref{section:discussion}. ", "conclusions": "\\label{section:discussion} \\par The main important result of the present study is that the reduction of the uncertainties on the \\iso{22}Ne $+ \\, \\alpha$ reaction rates has allowed us to considerably reduce the uncertainties coming from these rates on the production of isotopes from Mg to P in AGB stars of intermediate mass. The uncertainties on the Mg yields are now at a level of $\\sim 30$\\%, much lower than those obtained when using the NACRE upper or lower limits. The yields of \\iso{25}Mg and \\iso{26}Mg are 20\\% to 45\\% and 9\\% to 16\\%, respectively, smaller with the new rates, as compared to NACRE. The uncertainties have also been reduced for species heavier than Mg, where for example, the uncertainty on the production of P in solar metallicity models is now at a level of 35\\%, much lower than the 400\\% obtained when using the NACRE upper limit. These results are clearly illustrated by figures~\\ref{yield-compare1} and~\\ref{yield-compare2} and in table~\\ref{tab:mgyields}. \\par From our analysis it would appear that among the uncertainties related to the stellar models, those coming from the treatment of convection and of the mass loss are the largest. We described earlier that these are enormous when compared to the current uncertainties coming from the \\iso{22}Ne $+ \\, \\alpha$ reaction rates. Much effort is needed to improve the situation for AGB models, in particular with respect to convection, by trying to evaluate and reduce the uncertainties, perhaps by exploiting all the available observational constraints. \\par Our new evaluation of the \\iso{22}Ne $+ \\alpha$ reaction rates will encourage much future work, as these rates are important for many nucleosynthetic processes and sites. The \\iso{22}Ne($\\alpha, n$)\\iso{25}Mg reaction is an important source of neutrons both during the final evolutionary stages of massive stars, and during the AGB phase of low to intermediate mass stars. It is responsible for the production of heavy s-process elements in these environments \\citep{gallino98,rauscher02}. The new rate and its uncertainties have to be tested in relation to this process. Moreover, the smaller uncertainties of the rates presented here with respect to those in NACRE, appear to rule out the possibility that the production of the relatively abundant p-only isotopes of Mo and Ru could be related to a high value of the \\iso{22}Ne($\\alpha, n$)\\iso{25}Mg rates \\citep{costa00}. \\par It is important to study the relative production of \\iso{25}Mg and \\iso{26}Mg, as we have done in table~\\ref{tab:mgyields}, because both spectroscopic observations and the analysis of pre-solar grains are able to separate these two isotopes. Also the contribution of radioactive \\iso{26}Al to the abundance of \\iso{26}Mg has to be carefully evaluated. In particular, one pre-solar spinel grain \\citep[][OC2]{zinner05} appears to bear the signature of nucleosynthesis in intermediate mass AGB stars, with excesses in both \\iso{25}Mg and \\iso{26}Mg. A future application of our present work will be to compare our detailed results to the composition of this grain, extending the study to the oxygen isotopic ratios in AGB stars and their uncertainties, as these are also measured in the grain." }, "0601/astro-ph0601535_arXiv.txt": { "abstract": "{ {We present a multiwavelength study of the massive star forming region associated with IRAS 06055+2039 which reveals an interesting scenario of this complex where regions are at different stages of evolution of star formation. Narrow band near-infrared (NIR) observations were carried out with UKIRT-UFTI in molecular hydrogen and Br$\\gamma$ lines to trace the shocked and ionized gases respectively. We have used 2MASS $J H K_{s}$ data to study the nature of the embedded cluster associated with IRAS 06055+2039. We obtain a power-law slope of 0.43$\\pm$0.09 for the $K_{s}$-band Luminosity Function (KLF) which is in good agreement with other young embedded clusters. We estimate an age of 2 -- 3 Myr for this cluster. The radio emission from the ionized gas has been mapped at 610 and 1280 MHz using the Giant Metrewave Radio Telescope (GMRT), India. Apart from the diffuse emission, the high resolution 1280 MHz map also shows the presence of several discrete sources which possibly represent high density clumps. The morphology of shocked molecular hydrogen forms an arc towards the N-E of the central IRAS point source and envelopes the radio emission. Submillimetre emission using JCMT-SCUBA show the presence of a dense cloud core which is probably at an earlier evolutionary stage compared to the ionized region with shocked molecular gas lying in between the two. Emission from warm dust and the Unidentified Infrared Bands (UIBs) have been estimated using the mid-infrared (8 -- 21\\,$\\mu$m) data from the MSX survey. From the submillimetre emission at 450 and 850\\,$\\mu$m the total mass of the cloud is estimated to be $\\sim$ 7000 -- 9000 $\\rm M_{\\odot}$. } ", "introduction": "Massive stars are preferentially formed in dense cores of molecular clouds. They remain deeply embedded in the prenatal molecular gas and obscuring dust and their pre-main sequence time scales are much shorter compared to the low mass stars. The luminous high mass stars also affect the parent cloud. In addition, massive stars do not form in isolation but often in clusters and associations. All these factors contribute in making the study of the formation mechanisms of these systems very difficult. Multiwavelength studies, therefore, hold the potential to probe these complexes at different depths and unravel the least understood aspects of massive star formation. IRAS 06055+2039 (G189.78+0.34, RAFGL 5179) is a massive star forming region chosen from the catalog of massive young stellar objects by Chan et al. (1996). G189.78+0.34 is listed as an ultracompact (UC) HII region (Shepherd \\& Churchwell 1996; Bronfman et al. 1996). It belongs to the Gem OB1 molecular cloud complex and is a part of the extended HII region Sh 252. It is associated with S252 A which is one of the six compact radio sources in Sh 252 revealed from the 5 GHz aperture synthesis observations by Felli et al. (1977). IRAS fluxes yield a far infrared luminosity of $\\sim 10^{4}\\rm L_{\\odot}$ (Carpenter et al. 1995) for this star forming region. There are several kinematic distance estimates used in literature for this source which range from 1.25 kpc (Mirabel et al. 1987) to 2.9 kpc (Wouterloot \\& Brand 1989). In this paper we use the value of 2.6 kpc (Wu et al. 2001) which is the most widely used distance estimate for this source. This high mass star forming region has been observed as part of many surveys. H$_{2}$O maser (K\\\"{o}mpe et al. 1989; Lada \\& Wooden 1979) and the 6.7 GHz methanol maser (Szymczak et al. 2000a) have been detected towards IRAS 06055+2039. Positive detection has also been made in SiO (Harju et al. 1998), CO (2 - 1) (Wu et al. 2001) and CS (2 - 1) (Bronfman et al. 1996). Zinchenko et al. (1998) in their study of dense molecular cores have also mapped this source in CS (2 - 1). CO maps of Shepherd \\& Churchwell (1996) do not show evidence of any bipolar outflows. Search for the 6 cm (Szymczak et al. 2000b) and 5 cm (Baudry et al. 1997) OH masers show negative results. The IRAS low resolution spectra show a relatively red continuum from 13 to 23\\,$\\mu$m with the presence of an emission feature at 11.3\\,$\\mu$m which is attributed to the presence of Polycyclic Aromatic Hydrocarbon (PAH) molecules (Kwok et al. 1997). In this paper, we present a multiwavelength study of this star forming region. In Sect.\\,\\,\\ref{obvs.sec}, we present the narrow-band near-infrared (NIR) and radio continuum observations and a brief description of the related data reduction procedures. In Sect.\\,\\,\\ref{arch.sec}, we discuss other available datasets used in the present study. Section\\,\\,\\ref{results.sec} gives a comprehensive discussion on the results obtained and in Sect.\\,\\,\\ref{concl.sec}, we summarize the results. ", "conclusions": "\\label{results.sec} \\subsection{Embedded Cluster in the Near-Infrared} \\subsubsection{Radial Profile and Stellar Surface Number Density} \\label{ssnd.sect} The 2MASS $K_{s}$ - band image of the region around IRAS 06055+2039 is shown in Figure\\,\\,\\ref{2mass_k.fig}. We see the presence of a diffuse emission region harbouring an infrared cluster. \\begin{figure} \\centering \\resizebox{\\hsize}{!}{\\includegraphics{fig4.ps}} \\caption{ The 2MASS $K_{s}$ - band image of the region around IRAS 06055+2039. The presence of a diffuse emission region and a NIR cluster is seen. The plus sign marks the position of the IRAS point source.} \\label{2mass_k.fig} \\end{figure} This cluster is also listed in the catalog of embedded infrared clusters compiled by Bica et al. (2003). We use the 2MASS $JHK_{s}$ data to study the nature of this cluster. To determine the $K_{s}$ - band radial profile and the stellar surface number density (SSND), we have selected sources which are detected in the $K_{s}$ band. To estimate the cluster radius we select a large region of radius $\\rm 300\\arcsec$ centered on IRAS 06055+2039 ($\\rm \\alpha_{2000.0} = 06^{h} 08^{m} 32^{s}.1\\,;\\,\\delta_{2000.0} = +20^{\\circ} 39\\arcmin 18\\arcsec$). To account for the contribution from the field stars we select a control field ($\\rm \\alpha_{2000.0} = 06^{h} 09^{m} 52^{s}.0\\,;\\,\\delta_{2000.0} = +20^{\\circ} 39\\arcmin 18\\arcsec$) which is $\\sim$ $\\rm 20\\arcmin$ to the east of IRAS 06055+2039. Figure\\,\\,\\ref{cl_rad.fig} shows the radial profile of the stellar density in log-log scale. This profile was created by counting the number of stars in $\\rm 10\\arcsec$ annuli and normalizing by the annulus area. We fit two models to the surface density radial profile -- the King's profile and the inverse radius ($r^{-1}$) model. Neglecting the tidal radius, the King's profile can be written as \\begin{equation} f(r) = a + \\frac{f_{0}}{1+(r/r_{c})^{2}} \\end{equation} where, $f_{0}$ is the core concentration at radius zero, $r_{c}$ is the core radius and $a$ is a constant for the background offset. As seen in Fig.\\,\\,\\ref{cl_rad.fig}, both the models describe the density distribution fairly well. However, the $r^{-1}$ model has a better overall fit (reduced $\\chi^{2}$ = 1.2) as compared to the King's model (reduced $\\chi^{2}$ = 3.3). Several studies have shown that young embedded clusters have been fitted by $r^{-1}$ profiles (e.g McCaughrean \\& Stauffer 1994; Lada \\& Lada 1995) or, by both the inverse radius model and the King's model (e.g Horner et al. 1997; Teixeira et al. 2004; Baba et al. 2004). As discussed in Baba et al. (2004), the $r^{-1}$ dependence is likely to be a reminiscent of the parental cloud core whereas the King's profile represents systems in dynamical equilibrium. Hence, the better fitting of the $r^{-1}$ model could possibly suggest that the cluster associated with IRAS 06055+2039 is not yet in complete dynamical equilibrium. Within errors, the cluster profile merges with the field star level at $\\rm \\sim 85\\arcsec$ which translates to $\\sim$ 1.1 pc at a distance of 2.6 kpc. We take this as the cluster radius. The background level as estimated from the control field is $\\sim \\rm 9\\, stars\\, pc^{-2}$ which is in reasonable agreement with the value of $\\rm 11.8\\pm1.6\\, stars\\, pc^{-2}$ yielded by the King's profile fitting. The King's profile fitting also gives a core radius $\\rm r_{c} \\,\\sim 0.1\\, pc$. The core radius is a scale parameter and depends mainly on cluster parameters like density, luminosity, total mass etc. Several studies have shown that the core radius of the cluster is also correlated with its age. In their study of young clusters in the LMC, Elson et al. (1989) show that the core radii increases between $\\sim 10^{6}$ and $10^{9}$ yr and then begin to decrease again. Such trends in core radius evolution has also been discussed by Wilkinson et al. (2003) for the LMC clusters and Mackey \\& Gilmore (2003) for clusters in the SMC. These authors have shown that apart from the general increasing trend, the spread in the core radii also increases with the age of the cluster. Teixeira et al. (2004) and Baba et al. (2004) derive core radius values of 0.05 and 0.08 pc for the clusters NGC 2316 and RCW 36 respectively. The age estimate for both these clusters is 2 -- 3 Myr. For a comparatively older (5 -- 10 Myr) cluster NGC 2282, Horner et al. (1997) obtain a core radius of 0.19 pc. The above values of core radii seem to suggest that the cluster associated with IRAS 06055+2039 is likely to be of the same age ($\\sim$ 2 -- 3 Myr) as NGC 2316 and RCW 36. The number of stars detected in the $K_{s}$ band within $\\rm 85\\arcsec$ radius of the cluster is 114. The total background population is 34 [$\\rm = \\,9 (background) \\times \\pi \\times (1.1)^{2} (area \\,of\\, cluster)$]. Hence, the total number of cluster members is estimated to be 80. This yields a space density of $\\sim$ 14 stars\\,pc$^{-3}$. However, it should be noted here that this density is a lower limit as we are not completely sampling the stellar population at fainter magnitudes due to the 2MASS sensitivity limits. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig5.ps}} \\caption{The radial profile of the surface number density for the cluster associated with IRAS 06055+2039 in log-log scale. Also plotted are the two fitted models -- King's model (solid) and inverse radius (dashed). As compared to the King's model, the inverse radius model fits the radial profile much better. The horizontal dotted line corresponds to the background field star level which is $\\sim$ 9 stars pc$^{-2}$. Statistical errors are shown. } \\label{cl_rad.fig} \\end{figure} \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig6.ps}} \\caption{The contour map of the stellar surface number density obtained by counting stars in a $\\rm 10\\arcsec \\times 10\\arcsec$ ( $\\sim$ 0.1 $\\times$ 0.1 pc) grid for the cluster associated with IRAS 06055+2039. The contours are from 30 to 140 stars pc$^{-2}$ in steps of 20 stars pc$^{-2}$. The lowest contour is at the 3$\\sigma$ level. The positions of the IRAS point source (plus), the radio peak (open triangle) and the sub-mm peak (star) are shown in the figure.} \\label{cl_sdm.fig} \\end{figure} Figure\\,\\,\\ref{cl_sdm.fig} shows the surface density map of the region around IRAS 06055+2039. We see the presence of two peaks. The prominent peak which lies to the west coincides spatially with the central bright source. The secondary peak, which lies $\\rm \\sim 50\\arcsec$ to the east of the main peak, coincides well with the departure seen in the surface density profile from the model profiles around a radial distance of $\\sim$ 0.6 pc. This secondary peak is situated close to the edge of the sub-mm peaks presented in Fig.\\,\\,\\ref{jcmt.fig}. The stellar surface density distribution exhibits a centrally condensed-type structure rather than a hierarchical-type structure (Lada \\& Lada 2003). This is consistent with the fact that the derived radial profile fits reasonably well both to the King's model as well as the inverse radius model. Hierarchical-type complexes do not follow any well defined profile. Though at a less significant level, centrally condensed clusters have also been seen to show the presence of structures. For example, Lada \\& Lada (1995) show the presence of satellite subclusters in the outer regions of IC 348. This is consistent with the double peaked structure that we see for the cluster associated with IRAS 06055+2039. \\subsubsection{Colour-Colour (CC) and Colour-Magnitude (CM) Diagrams} \\label{cccmd.sect} The ($H-K$) versus ($J-H$) CC diagram for the cluster associated with IRAS 06055+2039 is shown in Fig.\\,\\,\\ref{cc.fig}. In this figure we have plotted 53 sources which have good quality photometric magnitudes in all the three $JHK_{s}$ bands (2MASS `read-flag' value of 1 -- 3). Henceforth, for the analysis of 2MASS data, we will be using only good quality photometric data which have the above `read-flag' values. For clarity we have classified the CC diagram into three regions (e.g. Sugitani et al. 2002; Ojha et al. 2004a \\& b). The ``F\" sources are located within the reddening bands of the main sequence and the giant stars. These sources are generally considered to be either field stars, Class III objects or Class II objects with small NIR excess. ``T\" sources populate the region redward of the ``F\" region but blueward of the reddening line corresponding to the red end of the T Tauri locus. These sources are classical T Tauri stars (Class II objects) with large NIR excess or Herbig AeBe stars with small NIR excess. Redward of the ``T\" region is the ``P\" region which are mostly protostar-like Class I objects and Herbig AeBe stars. Majority of sources in our sample are almost equally distributed in the ``F\" and the ``T\" regions, whereas, only four lie in the ``P\" region. A total of 18 out of 80 (22\\%) sources show infrared excess (i.e. sources populating the ``T\" and the ``P\" regions). However, it is important to note here that this NIR excess fraction is just the lower limit as several cluster members detected in the $K_{s}$ band were not detected in the other two shorter wavelength bands and hence are not part of this sample. The NIR excess in pre-main sequence (PMS) stars is due to the optically thick circumstellar disks/envelopes. These disks/envelopes become optically thin with age hence the fraction of NIR excess stars decreases with age. For very young ($\\le$ 1 Myr) embedded clusters the fraction is $\\sim$ 50 \\% (Lada et al. 2000; Haisch et al. 2000) and it decreases to $\\sim$ 20 \\% for more evolved (2 -- 3 Myr) clusters (Haisch et al. 2001; Texeira et al. 2004). The fraction of stars with NIR excess seen in this cluster suggests an upper limit 2 -- 3 Myr on the age which is reasonably consistent with that suggested from the core radius value. The age estimates for the cluster NGC 2175 associated with the extended HII region Sh 252 are 2 Myr (Grasdalen \\& Carrasco 1975) and 1 -- 2 Myr (K\\\"{o}mpe et al. 1989) which agrees rather well with our estimates for the cluster associated with IRAS 06055+2039 which also belongs to the Sh 252 complex. We have calculated the extinction by de-reddening the stars in the CC diagram. The stars are shifted to a line drawn tangential to the turn-off point of the main sequence locus (see Fig.\\,\\ref{cc.fig}). The amount of shift gives an estimate of the extinction of individual stars. The extinction values range from $A_{V}$ $\\sim$ 0 to 13 mag with an average foreground extinction of $A_{V}$ $\\sim$ 7 mag. The range of extinction values obtained shows up as the spread of stars along the reddening band in the CC diagram. This indicates that the cluster is partially embedded (Teixeira et al. 2004). Also according to Lada \\& Lada (2003), low extinction values ($A_{V}$ $\\sim$ 1 -- 5 mag) are typical of partially embedded clusters. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig7.ps}} \\caption{Colour-colour diagram of the infrared cluster in the IRAS 06055+2039 region. The two solid curves represent the loci of the main sequence (thin line) and the giant stars (thicker line) derived from Bessell \\& Brett (1988). The long-dashed line is the classical T Tauri locus from Meyer et al. (1997). The parallel dotted lines are reddening vectors with the crosses placed along these lines at intervals corresponding to five magnitudes of visual extinction. We have assumed the interstellar reddening law of Rieke \\& Lebofsky (1985) ($A_{J}/A_{V}$ = 0.282; $A_{H}/A_{V}$ = 0.175 and $A_{K}/A_{V}$ = 0.112). The short-dashed line represents the locus of the Herbig AeBe stars (Lada \\& Adams 1992). The plot is classified into three regions namely ``F\", ``T\" and ``P\" (see text for details). The colours and the curves shown in the figure are all transformed to the Bessell \\& Brett (1988) system. The solid line shown is drawn tangential to the turn-off point of the main sequence locus. The arrow points to the position corresponding to the central brightest star in the cluster. The mean photometric errors are shown in the upper right corner.} \\label{cc.fig} \\end{figure} Figure\\,\\,\\ref{cmd.fig} shows the ($H-K$) versus $K$ colour-magnitude (CM) diagram for 79 sources with good quality $HK$ magnitudes. Using the zero age main sequence (ZAMS) loci and the reddening vectors, we estimate the spectral type of the brightest star in the cluster to be $\\sim$ B0.5. This is the central IRAS point source and from the NIR estimates seems to be the most massive star in the cluster. A similar estimate for the spectral type is also obtained from the analysis of the ($J-H$) versus $J$ CM diagram. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig8.ps}} \\caption{Colour-magnitude diagram of the infrared cluster in the IRAS 06055+2039 region. The nearly vertical solid lines represent the zero age main sequence (ZAMS) loci with 0, 20 and 40 magnitudes of visual extinction corrected for the distance. The slanting lines show the reddening vectors for each spectral type. The magnitudes and the ZAMS loci are all plotted in the Bessell \\& Brett (1988) system. The arrow points to the position corresponding to the central brightest star in the cluster. The mean photometric errors are shown in the upper right corner.} \\label{cmd.fig} \\end{figure} \\subsubsection{$K_{s}$ - Band Luminosity Function} We use the 2MASS $K_{s}$ - band star counts to derive the $K_{s}$ - band luminosity function (KLF) for the embedded cluster. In order to obtain the KLF of the cluster it is essential to account for the background and foreground source contamination. For this purpose we use both the Besan\\c{c}on Galactic model of stellar population synthesis (Robin et al. 2003) and the observed control field star counts. We use the same control field as described in Sect\\,\\,\\ref{ssnd.sect}. Star counts were predicted using the Besan\\c{c}on model in the direction of the control field. We have checked the validity of the simulated model by comparing the model KLF with that of the control field and found both the KLFs to match rather well. As mentioned in the previous section, the average foreground extinction is determined to be A$_{V}$$\\sim$ 7 mag. Hence, assuming spherically symmetric geometry, the background population is then seen through a cloud with extinction upto A$_{V}$$\\sim$ 14 mag (7 $\\times$ 2). Model simulations with A$_{V}$ = 0 mag and d $<$ 2.6 kpc gives the foreground contamination. The background population is generated with A$_{V}$ = 14 mag and d $>$ 2.6 kpc. We determine the fraction of the contaminating stars (foreground + background) over the total model counts. This fraction is used to scale the observed control field and subsequently the star counts of the modified control field are subtracted from the KLF of the cluster to obtain the final corrected KLF which is shown in the left panel of Fig.\\,\\,\\ref{klf_klf_slope.fig}. \\begin{figure*} \\centering \\resizebox{\\hsize}{!}{\\includegraphics{fig9l.ps} \\includegraphics{fig9r.ps}} \\caption{Left: The corrected $K_{s}$ - band luminosity function (KLF) for the cluster around IRAS 06055+2039 is shown (solid line). The dotted line is the luminosity function without foreground/background correction. Right: The KLF shown as logN versus the $\\rm K_{s}$ magnitude. The straight line is the least squares fit to the data points in the magnitude range 11.5 -- 14.5.} \\label{klf_klf_slope.fig} \\end{figure*} The right panel of Fig.\\,\\,\\ref{klf_klf_slope.fig} shows the KLF plotted as $\\rm log N$ versus the $K$ magnitude. The KLF can be written as a power-law \\begin{equation} \\frac{dN(K_{s})}{dK_{s}} \\propto 10^{\\alpha K_{s}} \\end{equation} where, the left side of the equation denotes the number of stars per unit magnitude bin and $\\alpha$ is the slope of the power-law. A linear least squares fitting algorithm is used to fit the above power-law to the KLF in the magnitude range $K_{s} = 11.5 \\,\\rm to \\,14.5$. We obtain a value of $\\alpha$ = 0.43$\\pm$0.09 for the cluster. The least square fitting is done taking into account the statistical errors on individual data points. Within the quoted errors the estimate of the power-law slope is consistent with the average value of slopes ($\\alpha \\sim 0.4$) obtained for other young clusters (Lada et al. 1991; Lada \\& Lada 1995; Lada \\& Lada 2003). The power-law slope values obtained for other embedded clusters like W3 main and NGC 7538 are significantly lower ($\\alpha$ $\\sim$ 0.17 to 0.33 -- Ojha et al. 2004a \\& b; Balog et al. 2004). However, it should be noted here that these clusters are much younger ($\\la$ 1 Myr) and the surveys are deeper ($K_{s}\\le 17.5$) which probe the low mass stellar population down to $\\sim \\rm 0.1M_{\\odot}$. We estimate the masses of the sources in the cluster by comparing them with the evolutionary models of Palla \\& Stahler (1999). Figure\\,\\,\\ref{mass_spec.fig} shows the ($J-H$) versus $J$ CM diagram for the cluster field. We use the $J$-band magnitudes rather than $K_{s}$ because it is less affected by emission from circumstellar material. The solid curve represents the ZAMS isochrone for a 2 Myr cluster from Palla \\& Stahler (1999). Majority of sources have a typical mass of $\\sim \\rm 2 M_{\\odot}$ which we assume to be representative of the stellar population in the cluster. For a log-normal IMF, the power law slope ($\\gamma$), which is 1.35 for a Salpeter IMF, is variable and is given by $\\rm \\gamma = 0.94 + 0.94\\,log(m_{\\star})$, where $m_{\\star}$ is the stellar mass (Miller \\& Scalo 1979; Lada et al. 1993). Using this relation, we estimate $\\gamma \\sim 1.2$ for this mean mass. The power-law slope for the mass to luminosity relation is $\\beta$ $\\approx 1$ for clusters of age $\\sim 10^{6}$ Myr (Simon et al. 1992; Lada et al. 1993). The corresponding slope of the KLF, $\\alpha$ ($=\\gamma/2.5\\beta$), is 0.48 which is consistent with the value obtained from the least squares fit to the KLF. As is seen from Fig.\\,\\ref{mass_spec.fig}, the mass of the majority of sources in the cluster are below 2.5 $\\rm M_{\\odot}$ and the lowest mass limit is $\\sim$ 0.4 $\\rm M_{\\odot}$ from our sample. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig10.ps}} \\caption{The mass spectrum for the cluster associated with IRAS 06055+2039. The solid curve represents the model isochrone from Palla \\& Stahler (1999) for a 2 Myr PMS stellar population for the mass range 0.1 -- 3 $\\rm M_{\\odot}$. The slanted dotted lines are the reddening vectors for 2.5 and 0.2 $\\rm M_{\\odot}$ PMS stars respectively. Sources which lie above the reddening vector for 2.5 $\\rm M_{\\odot}$ are luminous massive stars. The arrow points to the position corresponding to the central massive star in the cluster.} \\label{mass_spec.fig} \\end{figure} \\subsection{Spatial Distribution of UIBs from MSX Data} \\label{uib.sect} We have used the scheme developed by Ghosh \\& Ojha (2002) to extract the contribution of UIBs (due to the Polycyclic Aromatic Hydrocarbons (PAHs)) from the mid-infrared images in the four MSX bands. The emission from each pixel is assumed to be a combination of two components. The first is the thermal continuum from dust grains (gray body) and the second is the emission from the UIB features falling within the MSX bands. The scheme assumes the dust emissivity to follow the power law of the form $\\epsilon_{\\lambda} \\propto \\lambda^{-1}$ and the total radiance due to the UIBs in band C to be proportional to that in band A. A self consistent non-linear chi-square minimization technique is used to estimate the total emission from the UIBs, the temperature and the optical depth. The spatial distribution of emission in the UIBs with an angular resolution of $\\rm \\sim 18\\arcsec$ (for the MSX survey) extracted for the region around IRAS 06055+2039 is shown in Fig.\\,\\,\\ref{pah.fig}. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig11.ps}} \\caption{Spatial distribution of the total radiance in the UIBs for the region around IRAS 06055+2039 as extracted from the MSX four band images. The contour levels are at 5, 10, 20, 30, 40, 50, 60, 70, 80, 90 and 95 \\% of the peak emission of $\\rm 4.1\\times10^{-5}$ W\\,m$^{-2}$\\,Sr$^{-1}$. The peak position of the UIB distribution (filled triangle), the 1280 MHz radio emission (open star) and the IRAS point source position (plus) are also indicated.} \\label{pah.fig} \\end{figure} Comparing the morphology with the radio continuum maps, we see that the emission from the UIBs is much more extended though the gross morphologies are similar with a relatively steep intensity gradient towards the N-E and a smoothly decreasing intensity distribution to the S-W. The peak position of the UIB distribution matches rather well with the radio peak. The integrated emission from the region around IRAS 06055+2039 in the UIB features within the band A of MSX (viz., 6.2, 7.7 \\& 8.6\\,$\\mu$m) is found to be 2.85 $\\times 10^{-12}$ Wm$^{-2}$ (see Fig.\\,\\,\\ref{pah.fig}). For comparison, we have estimated the emission in individual UIB features from the IRAS LRS spectrum (covering 8 -- 22\\,$\\mu$m), for this source. The total emission in the 7.7 and 8.6 $\\mu$m features is $\\sim$ 7.33 $\\times 10^{-12}$ Wm$^{-2}$, in the 11.3\\,$\\mu$m feature is $\\sim$ 1.32 $\\times 10^{-12}$ Wm$^{-2}$ and an upper limit for the 12.7\\,$\\mu$m feature is 5.9 $\\times 10^{-13}$ Wm$^{-2}$. The last value is an upper limit due to possible contamination from the Ne [II] line at 12.8\\,$\\mu$m. Hence, the UIB emission extracted from the MSX band A is $\\la$ 39\\% of the estimate from LRS. This is reasonable considering the larger effective field of view for the latter. \\subsection{Emission from Ionized Gas} The radio continuum emission from the ionized gas associated with the region around IRAS 06055+2039 at 1280 and 610 MHz is shown in Fig.\\,\\,\\ref{1280_610.fig}. The integrated flux densities from these maps are 183 and 282 mJy at 1280 and 610 MHz respectively. It should be noted here that the flux densities are obtained by integrating upto 3\\,$\\sigma$ level, where $\\sigma$ is the rms noise of the maps. The integrated flux densities obtained from the GMRT maps are consistent with the results of Felli et al. (1977) and White \\& Gee (1986). Felli et al. (1977) get values of 230 and 205 mJy at 1415 and 4995 MHz respectively. White \\& Gee (1986) estimate the total integrated flux density at 5 GHz to be 198 mJy. The contour maps display a cometary morphology with a bright arc-shaped edge on the N-E side and a smoothly decreasing intensity distribution on the opposite side. The cometary morphology is more clearly seen in the 610 MHz map. Such a morphology implies that the HII region is ionization bounded towards the N-E and density bounded towards the S-W. The position of the brightest radio peak matches with the central bright star seen in the infrared images. Similar morphology is seen from the 4995 MHz map of Felli et al. (1977). They report the presence of several dense clumps. According to them the radio continuum emission from this HII region (S252 A) is spatially separate from the larger extended HII region associated with Sh 252. CO line observations of Lada \\& Wooden (1979) show that this compact HII region (S252 A), the associated H$_{2}$O maser and CO bright spot are nearly coincident and located near the interface of the molecular cloud with the extended S252 HII region. The detection of the H$_{2}$O maser also implies the relative youth of the region. Lada \\& Wooden (1979) also suggest S252 A to be in the early stages of stellar evolution in which a massive star (or stars) has reached the main sequence and created a compact HII region within its parental molecular cloud. VLA maps at 15 and 5 GHz of White \\& Gee (1986) also trace similar cometary morphologies. Using the low frequency flux densities at 1280 and 610 MHz from our GMRT observations and 5 GHz data from White \\& Gee (1986), we derive the physical properties of the compact core of the HII region associated with IRAS 06055+2039. Mezger \\& Henderson (1967) have shown that for a homogeneous and spherically symmetric core, the flux density can be written as \\begin{equation} S = 3.07 \\times 10^{-2} T_{e} \\nu^{2} \\Omega (1 - e^{-\\tau(\\nu)}) \\end{equation} where, \\begin{equation} \\tau(\\nu) = 1.643a \\times 10^{5} \\nu^{-2.1} (EM) T_{e}^{-1.35} \\end{equation} where, $S$ is the integrated flux density in Jy, $T_{e}$ is the electron temperature in K, $\\nu$ is the frequency of observation in MHz, $\\tau$ is the optical depth, $\\Omega$ is the solid angle subtended by the source in steradians and $EM$ is the emission measure in $\\rm cm^{-6} pc$. $a$ is a correction factor and we use a value of 0.99 (using Table 6 of Mezger \\& Henderson 1967) for the frequency range 0.6 -- 5 GHz and $T_{e} = \\rm 8000 K$. The two GMRT maps are convolved to a common angular resolution of $\\rm 12\\arcsec \\times 12\\arcsec$ which is the resolution of the 5 GHz map of White \\& Gee (1986). In our case since the core is unresolved, $\\Omega$ is taken as this synthesized beam size (i.e $\\Omega = 1.133 \\times \\theta_{a} \\times \\theta_{b}$, where $\\theta_{a}$ and $\\theta_{b}$ are the half power beam sizes). The peak flux densities of the core in the 0.6 -- 5 GHz frequency range appear to lie in the optically thin region. Using these peak flux densities we derive the emission measure for an estimated electron temperature. The electron temperature ($ T_{e}$) of HII regions is known to increase linearly with Galacto-centric distance ($ D_{G}$) (Deharveng et al 2000 and references therein). This is due to the decrease in heavy element abundance with $ D_{G}$ which results in higher $ T_{e}$. The values of $ T_{e}$ derived by Omar et al. (2002) for a sample of three Galactic HII regions are also consistent with the relationship given in Deharveng et al. (2000). Assuming $ D_{G}$ as 10 kpc (Shirley et al. 2003) for IRAS 06055+2039 (S252A), we obtain a value of $\\sim$ 8000 K for the electron temperature. Using this value of the electron temperature, the peak flux densities were used to fit the above equations (Eqns. 3 \\& 4). The best fit value for the emission measure is $\\rm 8.8 \\pm 0.4 \\times 10^{4} cm^{-6} pc$. We get an estimate of $\\rm 1.05 \\times 10^{3} cm^{-3}$ for the electron density ($n_{e}$ $= (EM/r)^{0.5}$, $r$ being the core size which in this case corresponds to the synthesized beam size). These values agree reasonably well with the estimates of Felli et al. (1977) for the brightest peak (A3) of the component S252 A (see Fig. 5 of Felli et al. 1977). They derive a value of $\\rm 9.9 \\times 10^{4} cm^{-6} pc$ and $\\rm 5.95 \\times 10^{2} cm^{-3}$ for $EM$ and $n_{e}$ respectively. They assumed an electron temperature of $\\rm 10^{4} K$ and a distance of 2 kpc. It should also be noted here that the 5 GHz map of Felli et al. (1977) has a larger beam size ($\\rm \\sim 8\\arcsec\\times21\\arcsec$). Taking the total integrated flux density of 183 mJy at 1280 MHz and using the formulation of Schraml \\& Mezger (1969) and the table from Panagia (1973; Table II), we estimate the exciting star of this HII region to be of the spectral type B0 -- B0.5. This is consistent with the spectral class obtained by Felli et al. (1977) and White \\& Gee (1986). The FIR flux densities from the IRAS PSC yield a luminosity of $\\sim$ 10$^{4}\\rm L_{\\odot}$ which implies an exciting star of spectral type B0.5, in good agreement with radio measurements. In addition to the diffuse emission seen in our 1280 MHz map, we also detect a few discrete sources probably representing high density clumps which are listed in Table\\,\\,\\ref{radio_sources.tab}. We designate them as R1, R2,....R7 and their positions are marked in Fig.\\,\\ref{1280_610.fig}. Three such dense clumps were also detected in the 5 GHz map of Felli et al. (1977). The position of R1 is spatially coincident with the position of the central bright and massive ($\\sim$ B0.5) star seen in the infrared cluster and is possibly the exciting source of the HII region. The other dense clumps could also be possible discrete radio sources but the resolution of our map makes it difficult to comment on their nature. \\begin{table} \\caption{Discrete sources extracted from the 1280 MHz map of the region associated with IRAS 06055+2039.} \\label{radio_sources.tab} \\begin{tabular}{c|c|c|c} \\hline\\hline Source & RA (2000.0) & DEC (2000.0) & Peak Flux density \\\\ & (h m s) & (d m s) & (mJy/beam) \\\\ \\hline R1 & 06 08 32.16 & +20 39 18.7 & 5.7 \\\\ R2 & 06 08 31.95 & +20 39 23.1 & 5.4 \t \\\\ R3 & 06 08 32.79 & +20 39 16.1 & 4.3 \\\\ R4 & 06 08 31.04 & +20 39 24.9 & 3.3 \\\\ R5 & 06 08 30.92 & +20 39 30.3 & 2.8 \\\\ R6 & 06 08 31.39 & +20 39 02.1 & 3.2 \\\\ R7 & 06 08 32.20 & +20 39 01.7 & 2.8 \\\\ \\hline \\end{tabular} \\end{table} These seven clumps contribute $\\sim$ 5\\% of the total integrated emission from the ionized region around IRAS 06055+2039, the remaining being of diffuse nature. \\subsection{Emission from Shocked Neutral Gas} The rotational-vibrational line of molecular hydrogen (H$_{2}$ (1-0)S1, 2.12\\,$\\mu$m) traces the shocked neutral gas at the interface between the ionized and the molecular gas. In the photo-dissociation regions (PDRs), the molecular emission traces the first neutral layer beyond the ionization front. In Fig.\\,\\,\\ref{h2_brg_rad.fig}, we compare the morphologies of the 610 MHz continuum map with the continuum subtracted narrow-band H$_{2}$ (left) and Br$\\gamma$ (right) images. It is interesting to note that the shocked molecular hydrogen envelopes the radio emitting region. The Br$\\gamma$ image shows the presence of a faint diffuse emission which correlates well with the cometary morphology of the radio continuum emission. The H$_{2}$ arc which traces the ionization front lies beyond the Br$\\gamma$ emission. Comparison with the 1280 MHz continuum map also shows similar morphology. From the position of the H$_{2}$ arc, we estimate the radius of the HII region to be $\\sim$ 0.4 pc. \\begin{figure*} \\resizebox{\\hsize}{!}{\\includegraphics{fig12l.ps} \\includegraphics{fig12r.ps}} \\caption{610 MHz radio contours overlaid over the continuum subtracted H$_{2}$ (left) and Br$\\gamma$ (right) images for the region associated with IRAS 06055+2039. The contour levels are same as in Fig.\\,\\,\\ref{1280_610.fig}. The plus sign in both the images marks the position of the IRAS point source.} \\label{h2_brg_rad.fig} \\end{figure*} \\subsection{Emission from Dust: Temperature, Optical Depth and Dust Mass} As discussed in Sect.\\,\\,\\ref{uib.sect}, the MSX images were used to obtain the spatial distribution of temperature and optical depth ($\\tau_{10}$) of warm dust with the assumption that the dust is optically thin and the dust emissivity follows the power law of the form $\\epsilon_{\\lambda} \\propto \\lambda^{-1}$ (Mathis et al. 1983; Scoville \\& Kwan 1976). We obtain peak values of $1.4\\times10^{-4}$ and 155 K for $\\tau_{10}$ and the dust temperature respectively. Figure\\,\\,\\ref{msx_tautemp.fig} shows the optical depth and the mid-infrared dust temperature maps. \\begin{figure*} \\centering \\resizebox{\\hsize}{!}{\\includegraphics{fig13l.ps}\\includegraphics{fig13r.ps}} \\caption{Left: The spatial distribution of dust optical depth $\\tau_{10}$ in the region around IRAS 06055+2029. The contour levels are at 5, 10, 20, 30, 40, 50, 60, 70, 80 and 90 \\% of the peak value of $\\rm 1.4 \\times 10^{-4}$. Right: The spatial distribution of the mid-infrared dust temperature in this region. The contour levels are for the 100,115,125,135,140,150 and 154 K. The peak temperature is 155 K. The plus sign in both the images marks the position of the IRAS point source.} \\label{msx_tautemp.fig} \\end{figure*} Morphologically, the spatial distribution of the UIBs and the optical depth contours are similar with the intensity peaks matching rather well. This indicates the presence of higher dust densities near the embedded cluster. The temperature distribution shows lower values near the centre with a plateau like feature running diagonally along the S-E and N-W direction. Higher temperatures are towards the periphery with the peak seen towards the north. The optical depth and temperature are inherently anti-correlated. Hence, we see that the region has higher optical depth at the centre with the values decreasing outwards. Several peaks seen in the $\\tau_{10}$ map could possibly indicate the clumpy nature of the region. As will be evident later, we see similar trends for the $T\\rm(12/25)$ and $\\tau_{25}$ maps derived from the IRAS-HIRES images. The MSX Point Source Catalog (MSX PSC) lists two mid-infrared sources which fall within the radio nebulosity of IRAS 06055+2039. We designate them as M1 and M2. The MSX PSC flux densities for these two sources are listed in Table\\,\\,\\ref{msx_psc.tab}. \\begin{table} \\caption{Flux densities for the MSX point sources possibly associated with IRAS 06055+2039} \\label{msx_psc.tab} \\begin{tabular}{c|c c} \\hline\\hline & \\multicolumn{2}{|c}{Flux density (Jy)}\\\\ \\hline Wavelength ($\\mu$m) & M1$^{1}$ & M2$^{2}$\t\\\\ \\hline 8.3 & 3.52 & 0.97 \\\\ 12.1 & 6.24 & 1.40 \\\\ 14.7 & 2.42 & 4.19 \\\\ 21.3 & 6.66 & 13.98 \\\\ \\hline \\end{tabular} {\\tiny $^{1}$G189.7672+00.3407 ($\\alpha_{2000}= 06^{h}08^{m}33.02^{s}$; $\\delta_{2000} = 20^{\\circ}39\\arcmin32.04\\arcsec$)\\\\ $^{2}$G189.7677+00.3376 ($\\alpha_{2000}= 06^{h}08^{m}32.40^{s}$; $\\delta_{2000} = 20^{\\circ}39\\arcmin25.20\\arcsec$)\\\\} \\end{table} Comparing the MSX mid-infrared colours $\\rm F_{21/8}$, $\\rm F_{14/12}$, $\\rm F_{14/8}$ and $\\rm F_{21/14}$ of these two sources with study of the Galactic plane population by Lumsden et al. (2002), we infer that M1 is possibly a Herbig AeBe or a foreground star, whereas, M2 falls in the zone occupied mostly by compact HII regions. The NIR counterpart from 2MASS catalog for M2 is the central bright IRAS point source. The mid- and near-infrared colours $\\rm F_{21/8}$, $\\rm F_{8/K}$, $\\rm F_{21/12}$ and $\\rm F_{K/J}$ of this source are also consistent with compact HII regions. The IRAS-HIRES maps (at 12, 25, 60 and 100\\,$\\mu$m) were also used to obtain the spatial distribution of warm and cold dust colour temperatures ($T\\rm (12/25)$, $T\\rm (60/100)$) and optical depths ($\\tau_{25}$, $\\tau_{100}$). We assume the dust emissivity to follow the power law of the form $\\epsilon_{\\lambda} \\propto \\lambda^{-1}$. Figure\\,\\,\\ref{hires_1234.fig} shows the dust temperature and optical depth maps. The maps for the optical depth $\\tau_{25}$ and colour temperature $T\\rm (12/25)$ represent the warmer dust component. The distribution is centrally dense with the optical depth peak and hence lower derived temperature at the centre. This is similar to the distribution seen in Fig.\\,\\,\\ref{msx_tautemp.fig}. On the other hand $\\tau_{100}$ and $T\\rm (60/100)$ distributions are from a relatively colder component which probably forms an envelope around the warmer dust. Unlike the warmer dust temperature distributions (obtained from mid-infrared emission), the $T\\rm (60/100)$ distribution has its peak at the centre with the temperature decreasing towards the periphery. \\begin{figure*} \\centering \\resizebox{\\hsize}{!}{\\includegraphics{fig14ul.ps} \\hskip -2 cm\\includegraphics{fig14ur.ps}} \\resizebox{\\hsize}{!}{\\includegraphics{fig14ll.ps} \\hskip -2 cm\\includegraphics{fig14lr.ps}} \\caption{Upper panel -- The dust optical depth ($\\tau_{25}$) (left) and colour temperature ($T\\rm(12/25)$) (right) maps of the region around IRAS 06055+2039. The optical depth contours are 5, 10, 20, 30, 40, 50, 60, 70, 80, 90 \\% of the peak value of $2.2\\times10^{-5}$. The dust temperature contour levels are at 140, 145, 150, 155, 160, 170 and 180 K from the centre to the periphery. The peak value is 189 K. Lower panel -- The dust optical depth ($\\tau_{100}$) (left) and colour temperature ($T\\rm(60/100)$) (right) maps of the region around IRAS 06055+2039. The optical depth contours are 10, 20, 30, 40, 50, 60, 70, 80, 90 \\% of the peak value of $1.2\\times10^{-2}$ from the centre to the periphery. The dust temperature contour levels are at 20, 30, 33, 39, 46, 53, 59 and 62 K. The peak value is 66 K. The plus sign in all the images marks the position of the IRAS point source.} \\label{hires_1234.fig} \\end{figure*} For $\\tau_{25}$ and $\\tau_{100}$, we obtain peak values of $2.2\\times10^{-5}$ and $1.2\\times10^{-2}$ respectively. The dust temperature distributions for $T\\rm(12/25)$ and $T\\rm(60/100)$ peak at 189 and 66 K respectively. Schreyer et al. (1996) derive a value of 31.2 K for $T\\rm(60/100)$ and $1.21\\times10^{-3}$ for $\\tau_{100}$ from the IRAS PSC flux densities. This difference is due to the different resolution of the raw IRAS and the HIRES processed maps. Scaling the $\\tau_{25}$ peak value, we obtain a value of $\\tau_{10}$ $\\sim$ $5.6\\times10^{-5}$. The difference between this value and that obtained from the MSX data could be a result of different beam sizes and/or an inhomogeneous medium. Using the peak value of $\\tau_{100}$ distribution, derived from the IRAS-HIRES maps, we estimate the warm dust mass to be $\\sim$ 6 $\\rm M_{\\odot}$. We use the emission at submillimetre wavebands to study the cold dust environment in the region around IRAS 06055+2039. The spatial distribution of the sub-mm emission is shown in Fig.\\,\\,\\ref{jcmt.fig}. The angular resolutions are $\\rm 7\\arcsec.8$ and $\\rm 15\\arcsec.2$ for the 450 and 850\\,$\\mu$m wave bands respectively. The main source (central dense core which covers the region upto $\\sim$ 25\\% of the peak intensity) seems elongated in both the maps. Apart from this dense core, several dust clumps are seen which could have probably formed due to the fragmentation of the original cloud. The dust mass can be estimated from the following relation: \\begin{equation} M_{dust} = 1.88 \\times 10^{-4} \\left(\\frac{1200}{\\nu}\\right)^{3+\\beta} S_{\\nu}(e^{0.048\\nu/T_{d}} - 1) d^{2} \\end{equation} This is taken from Sandell (2000) and is a simplified version of Eqn. 6 of Hildebrand (1983). The above equation assumes the standard Hildebrand opacities (i.e. $\\rm \\kappa_{1200GHz} = 0.1 cm^{2}g^{-1}$). Here, $S_{\\nu}$ is the flux density at frequency $\\nu$, $T_{d}$ is the dust temperature which we assume to be 20 K (Klein et al. 2005; Mueller et al. 2002), $\\beta$ is the dust emissivity index and is taken to be 2 (Hildebrand 1983) and $d$ is the distance to the source in kpc. The flux densities are obtained from the JCMT-SCUBA maps shown in Fig.\\,\\,\\ref{jcmt.fig}. To obtain the flux density of the entire cloud, we have integrated upto the last contour (which is at 5\\% of the peak value). Using the above relation, we estimate dust masses of $\\sim$ 70 and 90 $\\rm M_{\\odot}$ from the 450 and 850\\,$\\mu$m maps respectively. Assuming a gas-to-dust ratio of 100, the above values translate to total masses of 7000 and 9000 $\\rm M_{\\odot}$ for the cloud from the 450 and 850\\,$\\mu$m maps respectively. We also estimate the total mass of only the central dense core to be $\\sim$ 875 and 1250 $\\rm M_{\\odot}$ from 450 and 850\\,$\\mu$m maps respectively. This source (S252A) has also been studied by Mueller et al. (2002) at 350\\,$\\mu$m and more recently by Klein et al. (2005) at 850 and 1300\\,$\\mu$m. Scaling to the distance and gas-to-dust ratio assumed by us, the corresponding mass from Mueller et al. (2002) is $\\sim$ 600 $\\rm M_{\\odot}$ and from Klein et al. (2005) is $\\sim$ 100 $\\rm M_{\\odot}$. Comparing the mass derived from the 850\\,$\\mu$m map, our estimate is very close to the mass obtained by Mueller et al. (2002) considering the fact that they have assumed a higher dust temperature (29 K). The mass estimate from the 450\\,$\\mu$m maps is $\\sim$ 35 \\% lower which could have been affected by the large atmospheric extinction correction applied to the data. The mass derived by Klein et al. (2005) are relatively lower. This could be possibly due to the fact that the flux density values presented by Klein et al. (2005) are lower compared to our values. From the CS line maps, Zinchenko et al. (1998) derive a cloud mass of 3132 $\\rm M_{\\odot}$. It should be noted here that the CS maps are from a larger region. Furthermore, a point worth discussing here is the effect of varying $T_{d}$ and $\\beta$. Exploring the range of $T_{d}$ (20 -- 40 K) and $\\beta$ (1 -- 2), we infer that the mass estimates can vary by upto a factor of $\\sim$ 8. \\subsection{Comparison of the Different Components Associated with IRAS 06055+2039} In Fig.\\,\\,\\ref{all.fig}, we present the various components of the region associated with IRAS 06055+2039. The plot displays the contour maps of the ionized gas and the emission from dust overlaid on the 2MASS $K_{s}$-band image. We show the contour plots of 610 MHz radio emission, 850\\,$\\mu$m cold dust emission and mark the peak positions for the 100 and 12\\,$\\mu$m emission from warm dust. The 850\\,$\\mu$m emission core lies to the S-E of the ionized region with the warmer dust in between. The 12 and 100\\,$\\mu$m peaks lie relatively closer to the radio peak. The ionized region is seen to be close to the edge of the molecular cloud. Comparing with Fig.\\,\\,\\ref{h2_brg_rad.fig}, we note that the shocked molecular gas lies in between the ionized region and the dense molecular core. The central region of the infrared cluster is located within the HII region which is at the edge of the molecular cloud. \\begin{figure} \\resizebox{\\hsize}{!}{\\includegraphics{fig15.ps}} \\caption{Contour maps of the emission from the ionized gas at 610 MHz (thick line) and dust emission at 850\\,$\\mu$m (thin line) are overlaid on the 2MASS $K_{s}$ band image for the region around IRAS 06055+2039. The dashed arc represents the position of shocked H$_{2}$. The peak positions of the IRAS-HIRES 12\\,$\\mu$m (open triangle) and 100\\,$\\mu$m (open star) maps are indicated. The cross marks the position of the methanol and water masers.} \\label{all.fig} \\end{figure} The ionized region and the dense molecular core as seen in the radio and the sub-mm maps respectively are possibly at different stages of evolution. The dense molecular core seen in our sub-mm maps does not show any radio emission (down to level of the rms noise which is 0.4 mJy/beam). The sub-mm peak is spatially offset from the FIR peaks and there are no MSX or NIR counterparts seen. This indicates a very early evolutionary stage for this dense and massive molecular core. It is most probably a very early protocluster candidate and we are sampling the initial collapse phase of the star forming core before the formation of the UCHII region (Williams et al. 2004 and references therein). This fact is further supported by the positions of the water and methanol masers which are seen to be coincident with the peak of the molecular core. The central peak of the CS map (Zinchenko et al. 1998) and the peaks of the 450 and 850\\,$\\mu$m JCMT-SCUBA maps match with the position of the masers. More specifically, it is known that methanol maser sites are generally radio quiet (as is the case here) and trace high mass star forming protoclusters in very early evolutionary phases (Minier et al. 2005). On the other hand, the ionized region which is associated with IRAS 06055+2039, has FIR emission, free-free radio emission and has NIR and MSX counterparts. This could probably indicate that the source is at a later evolutionary stage in between an evolved protocluster and a young cluster. Here, massive stars have started forming and a detectable HII region has been created. The cluster is partially embedded in the parental cloud. The sub-mm emission is weak here. A small subcluster is seen close to the edge of the sub-mm core spatially coincident with the secondary peak mentioned in Sect.\\,\\,\\ref{ssnd.sect}. Thus, from this multiwavelength study of the region associated with IRAS 06055+2039, we see the signatures of different evolutionary stages at different locations." }, "0601/astro-ph0601425_arXiv.txt": { "abstract": "{The blue Main Sequence (bMS) of $\\omega$ Cen implies a ratio of helium to metal enrichment $\\Delta Y/\\Delta Z \\approx 70$, which is a major enigma. We show that rotating models of low metallicity stars, which account for the anomalous abundance ratios of extremely metal poor stars, are also useful for understanding the very high $\\Delta Y/\\Delta Z$ ratio in $\\omega$ Cen. Models of massive stars with moderate initial rotation velocities produce stellar winds with large He-- and N--excesses, but without the large C-- (and O--) excesses made by very fast rotation, in agreement with the observed chemical abundance ratios in $\\omega$ Cen. It is still uncertain whether the abundance peculiarities of $\\omega$ Cen result from the fact that the high velocity contributions of supernovae escaped the globular cluster, usually considered as a tidally stripped core of a dwarf galaxy. Another possibility is a general dominance of wind ejecta at very low $Z$, due to the formation of black holes. Some abundance and isotopic ratios like $Mg/Al$, $Na/Mg$, $Ne/N$, $^{12}C/^{13}C$, $^{16}O/^{18}O$ and $^{17}O/^{18}O$ may allow us to further discriminate between these scenarios and between the AGB and massive star contributions. \\keywords stars:Omega Centauri -- Helium -- stars: evolution } \\titlerunning{The high Helium Sequence in $\\omega$ Cen} ", "introduction": "The globular cluster $\\omega$ Centauri has remarkable properties: it is the most massive globular cluster in the Galaxy and is often interpreted as the remaining core of an ancient dwarf galaxy (Bekki \\& Freeman \\cite{BekkiF03}), a possibility supported by the density profile of the cluster (Ideta \\& Makino \\cite{IdetaM04}). Dynamical studies support a formation of $\\omega$ Cen from one of the small progenitor galaxies of the Milky Way (e.g. Gnedin et al. \\cite{Gnedin02}). The stars in $\\omega$ Cen show a wide spread in metallicity (Norris \\& Da Costa \\cite{NorrisDC95}) from $[Fe/H] = -2$ to $-0.5$. Among its many abundance peculiarities, $\\omega$ Cen shows a large N--excess, an overabundance of s--elements relatively to Fe (Norris \\& Da Costa \\cite{NorrisDC95}) and an unusually low $[Cu/Fe]$ ratio relatively to other metal poor stars (Cunha et al. \\cite{Cunha02}; McWilliam \\& Smecker--Hane \\cite{McWilliam05}), which is interpreted as a relative lack of contributions from supernovae SNIa (Cunha et al. \\cite{Cunha02}). The finding of a double sequence in the globular cluster $\\omega$ Cen by Anderson (\\cite{Anderson97}; see also Bedin et al. \\cite{Bedin04}, Gratton \\cite{Gratton05}) and the further interpretation of the bluer sequence by a strong excess of helium constitutes a major enigma for stellar and galactic evolution (Norris \\cite{Norris04}). The interpretation in terms of an He excess is convincing and supported by stellar models as well by the morphology of the horizontal branch stars (Piotto et al. \\cite{Piotto05}). The great problem is that the bluer sequence with a metallicity $[Fe/H]= -1.2$ or $Z = 2 \\cdot 10^{-3}$ implies an He--content $Y=0.38$ (0.35-0.45), i.e. an He--enrichment $\\Delta Y = 0.14$ (see Norris \\cite{Norris04}). In turn, this demands a relative helium to metal enrichment $\\Delta Y/\\Delta Z$ of the order of 70 (Piotto et al. \\cite{Piotto05}; Gratton \\cite{Gratton05}). At the opposite, the system of globular clusters has a constant $Y=0.250$ (Salaris et al. \\cite{Salaris04}), while $[Fe/H]$ varies a lot, which implies a ratio $\\Delta Y/\\Delta X(\\mathrm{Fe})=0$, where $X(\\mathrm{Fe})$ is the iron mass fraction. The value $\\Delta Y/\\Delta Z=70$ is enormous and more than one order of magnitude larger than the current value of $\\Delta Y/\\Delta Z= 4-5$ (Pagel et al. \\cite{Pagel92}) obtained from extragalactic HII regions. A value of 4 -- 5 is consistent with the chemical yields from supernovae (Maeder \\cite{Maeder92}) forming black holes above about 20 -- 25 M$_{\\odot}$. The subject of the present work is to examine whether this extreme value of $\\Delta Y/\\Delta Z$ can be accounted for by models of rotating stars at very low metallicity. These models, which have the same physical ingredients as the models successfully used at solar $Z$, well account (Meynet et al. \\cite{MEMZfirst}) for the abundance anomalies observed in extremely metal poor halo stars. Sect. 2 collects the relevant observational determinations of the chemical abundances in $\\omega$ Cen and mentions the possible interpretations of the observed abundance peculiarities. In Sect. 3, we show results of models of rotating stars and in Sect. 4 we compare the results to observations of the bMS in $\\omega$ Cen. Sect. 5 gives the conclusions. ", "conclusions": "The wind contributions of low $Z$ massive rotating stars are able to produce the high $\\Delta Y/\\Delta Z$ observed in the bMS sequence in $\\omega$ Cen. The observations tend to favour an origin of the high helium observed by contributions from massive stars of intermediate rotation velocities. At this stage, it is not clear whether the dominance of wind contribution is a general feature of low $Z$ stars or whether the tidal evaporation experienced by $\\omega$ Cen has enabled it to lose most of the supernovae ejecta, keeping the enrichment from stellar winds. Some critical abundance and isotopic ratios may offer further signatures of the contributions of AGB and massive stars." }, "0601/astro-ph0601613_arXiv.txt": { "abstract": "The subpulse drifting phenomenon in pulsar radio emission is considered within the partially screened inner gap model, in which the sub-Goldreich-Julian thermionic flow of iron ions or electrons coexists with the spark-associated electron-positron plasma flow. We derive a simple formula that relates the thermal X-ray luminosity $L_{\\rm x}$ from the spark-heated polar cap and the \\EB\\ subpulse periodicity $\\hat{P}_3$ (polar cap carousel time). For PSRs B0943+10 and B1133+16, the only two pulsars for which both $\\hat{P}_3$ and $L_{\\rm x}$ are known observationally, this formula holds well. For a few other pulsars, for which only one quantity is measured observationally, we predict the value of the other quantity and propose relevant observations that can confirm or discard the model. Then we further study the detailed physical conditions that allow such partially screened inner gap to form. By means of the condition $T_{\\rm c}/T_{\\rm s}>1$ (where $T_{\\rm c}$ is the critical temperature above which the surface delivers a thermal flow to adequately supply the corotation charge density, and $T_{\\rm s}$ is the actual surface temperature), it is found that a partially-screened gap (PSG) can be formed given that the near surface magnetic fields are very strong and curved. We consider both curvature radiation (CR) and resonant inverse Compton scattering (ICS) to produce seed photons for pair production, and find that the former is the main agency to produce gamma-rays to discharge PSG. ", "introduction": "Pulsar radio emission typically occurs in a form of periodic series of narrow bursts of radiation. These burst often have a complex structure, in which higher order periodicities can be found. The individual pulses consist of one, few to several subpulses. In some pulsars the subpulses demonstrate a very systematic drift across the pulse window. If the pulses are folded with the basic pulsar period then the drifting subpulses form amazingly spectacular patterns called drift-bands. The phenomenon of drifting subpulses is a long standing puzzle in the pulsar research and its solution would likely result in deeper understanding to the nature of pulsar radiation. It is generally believed that this phenomenon is inherently associated with the so-called inner acceleration region above the polar cap, in which the magnetospheres plasma does not corotate with the neutron star surface. The first model based on this idea was proposed by \\citet[][ RS75 henceforth]{rs75}. The predictions of RS75 model were successfully compared with a handful of pulsars known to show this phenomenon at that time. A decade later, \\citet[][ R86 hereafter]{r86} compiled a list of about 40 drifting pulsars, but drifting subpulses were still regarded as some kind of exceptional phenomenon. However, recently \\citet[][ WES06 hereafter]{wes06} presented the results of a systematic, unbiased search for subpulse modulation in 187 pulsars and found that the fraction of pulsars showing drifting subpulse phenomenon is likely to be larger than 55\\%. They identified 102 pulsars with drifting subpulses in their sample, with a large fraction of newly discovered drifters. The authors concluded that the conditions required for the drifting mechanism to work cannot be very different from the emission mechanism of radio pulsars. WES06 then suggest that the subpulse drifting phenomenon is an intrinsic property of the pulsar emission mechanism, although drifting could in some cases be very difficult or even impossible to detect due to insufficient signal-to-noise ratio. It is therefore essential to attempt to unravel the physical conditions that can lead to formation of an inner acceleration region above the polar cap that could lead to development of the subpulse drift phenomenon. The classical vacuum gap model of RS75, in which spark-associated sub-beams of subpulse emission circulate around the magnetic axis due to $\\mathbf{E}\\times\\mathbf{B}$ drift of spark plasma filaments, provides the most natural and plausible explanation of drifting subpulse phenomenon. However, despite its popularity, it suffers from the so-called binding energy problem. Namely, under canonical conditions the surface charges (ions or electrons) are likely to be directly pulled out of the surface so that a pure vacuum gap is difficult to form. The alternative steady flow polar cap models, the so-called space-charge-limited flow models \\citep{as79,hm98}, cannot give rise to the intermittent ``sparking'' behavior, which seems necessary to explain the subpulse drift phenomenon in radio pulsars. \\citet[][ GM01 hereafter]{gm01} revisited the binding energy problem of RS75 model and argued that the formation of the vacuum gap (VG) is, in principle possible, although it requires a very strong non-dipolar surface magnetic fields, much stronger than a canonical dipolar component inferred from the observed spindown rate. Once the binding is strong enough to prevent the thermionic emission at the full space charge limited flow, the inevitable \\EB\\ drift of plasma filaments will result in the observable subpulse drift phenomenon. It has been known for a long time that in order to allow all radio pulsars to produce electron-positron pairs (the necessary condition for coherent radio emission), the near-surface magnetic fields must include multipole components dominating over global dipole field \\citep{rs75,as79,zhm00}. Growing observational evidence of non-dipolar structure of surface magnetic field \\footnote{It is worth mentioning that although RS75 implicitly assumed non-dipolar surface magnetic fields to treat the $\\gamma$-$B$ pair production processes, in their calculations of many other physical quantities (such as the surface charge density) they still used dipolar form, presumably for the sake of simplicity.} accumulates, and the suggestion that such a sunspot-like fields form during the early proto-neutron star stage has been proposed \\footnote{Another possible mechanism of creating small scale anomalies of surface magnetic fields was proposed by \\citet{grg03}. They argued that due to a Hall drift instability, the poloidal magnetic field structures can be generated from strong subsurface toroidal fields.} \\citep[e.g.][]{ug04}. \\citet[][ GM02 hereafter]{gm02} calculated the non-dipolar VG model for 42 drifting subpulse pulsars tabulated by R86 and argued that VG can be formed in all considered pulsars, provided that the actual surface magnetic field was close to $10^{13}$ G independently of the value of the canonical dipolar magnetic field. Although the binding energy problem could be, at least in principle, resolved by assuming an appropriately strong surface magnetic field, yet another difficulty of the RS75 model was that it predicted a much too fast \\EB\\ drift rate. Motivated by this issue, \\citet[][ GMG03 hereafter]{gmg03} developed further the idea of the inner acceleration region above the polar cap by including the partial screening by a sub-Goldreich-Julian thermal flow from the surface due to the spark-associated polar cap heating. This idea was first introduced by \\citep[ CR80 henceforth]{cr80}, who argued that even with thermionic ions included in the flow, the condition above the polar cap is close to that of a pure vacuum gap. A similar model was also invoked by \\citet{um95,um96}. GMG03 reanalyzed this model and argued that a thermostatic self-regulation should keep the surface temperature just few percent below the critical ion temperature at which the gap potential drop is completely screened. This results in more than 90 \\% of screening due to thermionic emission. Given a similar gap height, the actual potential drop, and hence the \\EB\\ drift rate, is about 10 \\% of that of the pure vacuum case. This is still above the threshold for the magnetic pair production. Similar to the avalanche pair production cascade introduced by RS75, the discharge of the growing potential drop above the polar cap would also occur in the form of a number of sparks. In this paper we call such an inner accelerator a partially screened gap (PSG henceforth). The latest XMM-Newton observation of the drifting pulsar PSR B0943+10 \\citep[][ ZSP05 hereafter]{zsp05} reveals a possible hot spot with the surface area much smaller than the conventional polar cap. This is consistent with the polar cap heating from such a PSG, which at the same time gives the right \\EB\\ drift rate. This lends strong support to the PSG model. In this paper we study PSG model in greater detail and explore the physical conditions for the model to work. We also apply this model to a new set of 102 pulsars from WES05 and show that it can work in every case, provided that the surface non-dipolar magnetic field is strong enough, even stronger than $10^{13}$ G suggested by GM02 and GMG03. Our new treatment is a combination of those used in GM02 and GMG03. Our working hypothesis is that drifting subpulses manifest the existence of a thin inner acceleration region, with an acceleration length scale much shorter than the polar cap size. The ultra-high accelerating potential drop discharges via a number of localized $\\mathbf{E}\\times\\mathbf{B}$ drifting sparks, as has been envisioned by RS75 . These sparks produce isolated columns of electron-positron plasma that stream into the magnetosphere to generate radio-beams of the observed subpulse emission due to some kind of plasma instability (see the \\S8 for more discussion). Due to charge depletion with respect to the co-rotational Goldreich-Julian (1969) value, sparks experience an unavoidable \\EB\\ drift with respect to the polar cap surface. As a consequence, the spark-associated sub-beams of radio emission perform a slow circumferential motion that is responsible for the observed subpulse drift. This model, which is often called ``a pulsar carousel model'' \\citep{dr99}, is examined in this paper \\footnote{Other suggestions of subpulse drifting as phenomena occurring outside the inner gap have been made \\citep{kmm91,sa02,w03,gmm05,fkk06}, but the connection between the radio drifting rate and the X-ray properties in those models is not yet clear, and we do not discuss them in the current paper.}. We are particularly interested in the thermal effect associated with the surface bombardment by back-flowing particles produced by sparks. The intrinsic drift rate (manifested by the tertriary drift periodicity) and the polar cap heating rate (manifested by the thermal X-ray luminosity) should be correlated with each other, since they are determined by the same value of the accelerating potential drop. The properties of drifting subpulses are discussed in \\S2 and the properties of charge depleted acceleration region above the polar cap are discussed \\S3. We find a specific relationship between the appropriate observables and conclud that it holds for a number of pulsars for which good quality data is available (especially PSR B0943+10, ZSP05). It turns out that this relationship depends only on the observational quantities and thus it is a powerful tool for testing this theoretical model. Although the number of pulsars that have all necessary data for such testing is small at the moment, the clean prediction holds the promise to ultimately confirm (or discard) the PSG model in the future. In \\S\\S4-7 we analyze this model in a more detailed manner to investigate the microscopic conditions (e.g. near-surface configuration, radiation mechanism, etc) that are needed to form such PSGs. The paper is summarized in \\S8. ", "conclusions": "The phenomenon of drifting subpulses has been widely regarded as a powerful tool for understanding the mechanism of coherent pulsar radio emission. RS75 first proposed that drifting subpulses are related to \\EB\\ drifting sparks discharging the high potential drop within the inner acceleration region above the polar cap. The subpulse-associated streams of secondary electron-positron plasma created by sparks were penetrated by much faster primary beam. This system was supposed to undergo a two-stream instability, which should lead to generation of the coherent radiation at radio wave lengths. However, careful calculations of the binding energy (critical for spark ignition) and the growth rate of the two-stream instability have shown that neither the sparking discharge nor the two-stream instability were able to work in a way proposed by RS75. Nonetheless, qualitatively their idea was still considered attractive, at least to the authors of this paper who have been continuing efforts to search for mechanisms that would actually make the RS75 model to work. In this paper we applied the partially screened gap model proposed by GMG03 to PSR B0943+10, a famous drifter for which ZSP06 successfully attempted to measure the thermal X-ray from hot polar cap. We derived a simple relationship between the X-ray luminosity $L_{\\rm x}$ from the polar cap heated by sparks and the tertiary periodicity $\\hat{P}_3$ of the spark-associated subpulse drift observed in radio band. This relationship reflects the fact that both the drifting (radio) and the heating (X-rays) rates are determined by the same value of the electric field in the partially screened gap. As a consequence of this coupling equations~(\\ref{Lx}) and (\\ref{LxE}) are independent of details of the acceleration region. In PSRs B0943$+$10 and B1133+16, the only two pulsars for which both $L_{\\rm x}$ and $\\hat{P}_3$ are measured, the predicted relationship between observational quantities holds very well. We note that $\\hat{P}_3$ is very difficult to measure and its value is known only for four pulsars: B0943$+$10, B1133+16, B0826$-$34 and B0834$+$06. The successful confrontation of the predicted X-ray luminosity with the observations in PSRs B0943$+$10 and B1133+16 encourages further tests of the model with future X-ray observations of other drifting pulsars. This is particularly relevant to PSR B0834$+$06, whose predicted X-ray luminosity is much higher than in PSR B0943+10, while the distance to both pulsars is almost identical. It is worth mentioning that due to relatively poor photon statistics it is still not absolutely clear whether the X-ray radiation associated with polar cap of PSR B0943+10 is thermal or magnetospheric (or both) in origin. However, the case of Geminga pulsar B0633+17 strongly supports thermal origin. Observations of PSR B0834$+$06 with {\\it XMM-Newton} should help to resolve this question ultimately. In both the steady SCLF model and the pure vacuum gap model, the potential increases with height quadratically at lower altitudes. However the growth rate is significantly different - the latter is faster by a factor of $R/r_{\\rm p}$. It is well known that in the pure vacuum case (RS75), the potential grows so fast that a primary particle could quickly generate pairs with a high multiplicity, and that some backward returning electrons generate more pairs and soon a ``pair avalanche'' occurs and the potential is short out by a spark. In the PSG model we are advocating, the potential drops by a factor of $\\eta$. For essentially all the cases we are discussing, this $\\eta$ value is much larger than the $r_{\\rm p}/R$ value required for the steady SCLF to operate, although it is much less than unity. The steady state condition is not satisfied, and the the gap is more analogous to a vacuum gap, i.e. the pair discharging happening intermittently. In fact, the partially screened potential drop is still above the threshold for the magnetic pair production, which in strong and curved surface magnetic field is a condition necessary and sufficient for the sparking breakdown. The original RS75 pure VG model predicts much too high a subpulse drift rate and an X-ray luminosity to explain the case of PSR B0943+10 and other similar cases. Other available acceleration models predict too low a luminosity and the explanation of drifting subpulse phenomenon is generally not clear at all (see ZSP05 for more detailed discussion). In summary, the bolometric X-ray luminosity for the space charge limited flow \\citep{as79} pure vacuum gap (RS75) and partially screened gap (GMG03) is $(10^{-4} - 10^{-5})\\dot{E}$ \\citep{hm02}, $(10^{-1} - 10^{-2})\\dot{E}$ (ZSP05), and $\\sim 10^{-3}\\dot{E}$ (this paper), respectively. The latter model also predicts right \\EB\\ plasma drift rate. Thus, combined radio and X-ray data are consistent only with the partially screened VG model, which requires very strong (generally non-dipolar) surface magnetic fields. Observations of the hot-spot thermal radiation almost always indicate bolometric polar cap radius much smaller than the canonical Goldreich-Julian value. Most probably such a significant reduction of the polar cap size is caused by the flux conservation of the non-dipolar surface magnetic fields connecting with the open dipolar magnetic field lines at distances much larger than the neutron star radius. Our analysis suggests the following pulsar picture: In the strong magnetic fields approaching $10^{14}$~G at the neutron star surface, the binding energy is high enough to prevent a full thermionic flow from the hot polar cap at the corotation limited level. A partially screened vacuum gap develops with the acceleration potential drop exceeding the threshold for the magnetic pair formation. The growth of this potential drop should be naturally limited by a number of isolated electron-positron spark discharges. As a consequence, the polar cap surface is heated by back-flowing particles to temperatures $T_{\\rm s}\\sim 10^6$~K, just below the critical temperature $T_{\\rm c}$ at which the thermionic flow screens the gap completely. The typical radii of curvature of the field lines ${\\cal R}$ is of the order of polar cap radii $r_{\\rm p}\\sim 10^3-10^4$~cm. The only parameter that is thermostatically adjusted in a given pulsar is the shielding parameter $\\eta=10^{-3}(B_{\\rm s}/B_{\\rm q}){\\cal R}^{-0.5}_6 P \\sim 0.001 T_6^{1.43}{\\cal R}^{-0.5}_6 P$, which determines the actual level of charge depletion with respect to the pure vacuum case ($\\eta$=1), and in consequence the polar cap heating rate as well as the spark drifting rate. It is worth to emphasize that $\\eta\\sim 0.1$ for longer period pulsars $P\\sim 1$ s and $\\eta\\sim 0.01$ for shorter period pulsars. Our calculations are consistent with PSR B0943$+$10 and few other drifting pulsars, for which the signatures of X-ray emission from the hot polar cap were detected. The sparks operating at the polar cap generate streams of secondary electron-positron plasma flowing through the magnetosphere. These streams are likely to generate beams of coherent radio emission that can be observed in the form of subpulses. \\citet{u87} first pointed out that the non-stationarity associated with sparking discharges naturally leads to a two-stream instability as the result of mutual penetration between the slower and the faster plasma components. \\citet{am98} developed this idea further, calculated the growth rates, and demonstrated that the instability is very efficient in generating Langmuir plasma waves at distances of many stellar radii $r_{\\rm ins} \\sim 10^{4-5} h^{CR}=10^{7-8}$ cm, where $ h^{CR}$ is the height of the acceleration region described by equation (21) \\citep{mgp00}. This is exactly where pulsar's radio emission is supposed to originate \\citep[e.g.][]{kg98}. Conversion of these waves into coherent electromagnetic radiation escaping from the pulsar magnetosphere was considered and discussed by \\citet{mgp00} and \\citet{glm04}. These authors demonstrated that the nonlinear evolution of Langmuir oscillations developing in pulsar's magnetosphere leads to formation of charged, relativistic solitons, able to emit coherent curvature radiation at radio wavelengths. The component of this radiation that is polarized perpendicularly to the planes of dipolar magnetic field can escape from the magnetosphere \\citep[see][for observational evidence of such polarization of pulsar radiation]{lea01}. The observed pulsar radiation in this picture is an indirect consequence of sparking discharges within the inner acceleration region just above the polar cap. In light of this paper we can therefore argue that the coherent pulsar radio emission is conditional on the presence of strong non-dipolar surface magnetic fields at the polar cap, with a strength about $10^{13-14}$~G and radius of curvature of the order $10^{4}$~cm. In the very strong surface magnetic field assumed within the accelerator, processes such as photon splitting \\citep{bh01} and bound pair creation \\citep{um95} may become important. It has been suggested that such processes could potentially delay pair creation and thus increase the height and voltage of the accelerator. For the photon splitting case, the delay is significant only if photons with both polarization modes split - a hypothesis in strong magnetic fields \\citep{bh01}. It could be possible that only one mode split \\citep{u02}. In such a case the gap height and potential of a PSG would not be significantly affected due to the high pair multiplicity in strong, curved magnetic fields. For bound pairs \\citep{um95,um96}, in the very hot environment near the neutron star surface (with temperatures exceeding 1 MK), it is possible that bound pairs could not survive long from photoionization. Following \\citet{bcp92} and \\citet{um96} one can roughly estimate the mean free path for bound pair photo-ionization. It turns out that for temperatures around (2-3) MK this mean free path is of the order of few meters, which is considerably less than the height of CR-driven PSG. In such a case, even if the bound pairs are initially produced, they would not significantly delay the pair formation. One could address this potential potential problem by referring to the detailed case study of PSR B0943+10 analyzed in this paper. This is the only pulsar in which we have full information concerning \\EB\\ drift rate and the polar cap heating rate. In the analysis we have used two methods. The first method is independent on details of accelerating region (such as height, potential drop, etc.) but is based only on the subpulse drift radio-data (eqs.~[\\ref{Lx},\\ref{LxE}]). The second one includes the detailed treatments of model parameters without considering the delaying effect by photon splitting and bound pairs (eqs.~[\\ref{Ts},\\ref{LxE2}]). Both methods give consistent results, as illustrated in Figures (1-4). We therefore conclude that the delaying processes, if any, are not significant in the PSGs given the physical conditions we invoke. Finally, we note that \\citet{um95} presented a model of the inner accelerator that shares some similar features with our model, although it is basically different. In their model the potential drop is self-regulated by partial screening close to the threshold for field ionization of bound pairs, so the surface temperature is maintained at the level necessary for this screening. In our model, the partial screening keeps the surface temperature slightly below the critical temperature at which the thermionic flow is space charge limited \\citet{as79}. In the light of results presented in this paper we claim that bound pairs do not affect the formation of inner accelerators in pulsars." }, "0601/astro-ph0601339_arXiv.txt": { "abstract": "% The aim of this paper is to determine the initial rotational velocities required on the ZAMS for single stars to reach the critical velocity, sometimes called the break-up velocity, during the Main-Sequence (MS) phase. Some useful relations between $\\Omega/\\Omega_{\\rm crit}$, $\\upsilon/\\upsilon_{\\rm crit}$ ($\\upsilon$ is the velocity at the equator), the moments of inertia, the angular momenta, the kinetic energy in the rotation and various other basic physical quantities are obtained. ", "introduction": "The new results from VLTI on the shape of Achernar \\citep{Do05} show that detailed parameters of rotating stars become accessible to interferometric observations. In order to make valuable comparisons with models of rotating stars, we provide here a few basic data on rotating stellar models. We also search the initial conditions required for a star to reach the critical limit during its Main-Sequence phase. When the surface velocity of the star reaches the critical velocity ({\\it i.e.} the velocity such that the centrifugal acceleration exactly balances gravity), one expects that equatorial ejections of matters ensue \\citep{Town04}. The physics involved in the ejection process, determining the quantity of mass ejected, the timescales between outburst episodes, the conditions required for a star to present such outbursts are important questions not only for understanding Be, B[e] or Luminous Blue Variable stars, but probably also for having a better knowledge on how the first stellar generations in the Universe behaved \\citep{EkTar}. Indeed, at lower metallicity, the stellar winds are weaker and much less angular momentum is removed at the surface. This favors an evolution to the critical limit and may have important consequences for the evolution of the first stellar generations. Also there are some indications that the distribution of the initial velocities contains more rapid rotators at lower metallicity \\citep{MG99}. Finally, realistic simulations of the formation of the first stars in the Universe show that the problem of the dissipation of the angular momentum is more severe at very low $Z$ than at the solar $Z$. Thus these stars might begin their evolution with a higher amount of angular momentum \\citep{Ab02}. Rotation affects all the outputs of the stellar models. The reader will find reviews on the effects of rotation in \\citet{MMAA}, \\citet{Hel00}, \\citet{Ta04} and \\citet{MT}. In this work, we compute 112 different stellar models with initial masses equal to 3, 9, 20 and 60 M$_\\odot$, for metallicities $Z$ equal to 0, 0.00001, 0.002 and 0.020, and for values of the ratio of the angular velocity $\\Omega$ to the critical angular velocity $\\Omega_{\\rm crit}$ equal to 0.1, 0.3, 0.5, 0.7, 0.8, 0.9, 0.99. For the 3, 9 and 20 M$_\\odot$ models, the computation was performed until the end of the Main-Sequence phase or until the star reaches the critical velocity. For the 60 M$_\\odot$ stars, we have obtained all the models on the ZAMS, but only a subset of them were computed further on the Main-Sequence. We consider that a star arrives on the ZAMS, when a fraction of 0.003 in mass of hydrogen has been burned at the center. On the ZAMS, the star is supposed to have a solid body rotation. During the Main-Sequence phase the variation of $\\Omega$ inside the star is computed self-consistently taking into account the various transport mechanisms, {\\it i.e.} convection, shear diffusion, meridional circulation and horizontal turbulence. The removal of angular momentum at the surface by the stellar winds and the changes of $\\Omega$ resulting from movements of contraction or expansion are also accounted for. A detailed description of the physical ingredients used in this grid of models will be given elsewhere (Ekstr\\\"om et al., in preparation), let us just mention here that the prescriptions for the opacity tables, the initial chemical compositions and the treatment of convection are as in \\citet{MMXI}. \\begin{figure}% \\plottwo{meynet_fig1a.eps}{meynet_fig1b.eps} \\caption{{\\it Left:} Variations of the momentum of inertia on the ZAMS as a function of the initial masses for various metallicities $Z$ (squares: $Z=0.020$; triangles: $Z=0.002$; pentagons: $Z=0.00001$ and circles: $Z=0$). {\\it Right:} Variations of the gravitational energy on the ZAMS as a function of $M_{\\rm ini}$ for various $Z$. } \\label{moin} \\end{figure} \\begin{figure}% \\plottwo{meynet_fig2a.eps}{meynet_fig2b.eps} \\caption{{\\it Left:} Variations of the polar radius on the ZAMS as a function of $M_{\\rm ini}$ for various metallicities. {\\it Right:} Fraction of the total mass (m$_r$ is the mass inside the radius $r$) as a function of the fraction of the total radius for solar metallicity models with $\\Omega/\\Omega_{\\rm crit}$=0.1. } \\label{radii} \\end{figure} \\begin{figure}% \\plottwo{meynet_fig3a.eps}{meynet_fig3b.eps} \\caption{{\\it Left:} Variation of the critical equatorial velocity on the ZAMS as a function of the initial mass for various metallicities. {\\it Right:} Variation of the critical rotation period on the ZAMS as a function of the initial mass for various metallicities. } \\label{vcrit} \\end{figure} Based on the set of numerical results described above, we first present useful relations between basic quantities, and then explore the initial conditions required for a star to reach the critical velocity during it MS evolution. ", "conclusions": "" }, "0601/astro-ph0601080_arXiv.txt": { "abstract": "{In order to determine the incidence of black hole accretion-driven nuclear activity in nearby galaxies, as manifested by their radio emission, we have carried out a high-resolution Multi-Element Radio-Linked Interferometer Network (MERLIN) survey of LINERs and composite LINER/H{\\sc ii} galaxies from a complete magnitude-limited sample of bright nearby galaxies (Palomar sample) with unknown arcsecond-scale radio properties. There are fifteen radio detections, of which three are new subarcsecond-scale radio core detections, all being candidate AGN. The detected galaxies supplement the already known low-luminosity AGN -- low-luminosity Seyferts, LINERs and composite LINER/H{\\sc ii} galaxies -- in the Palomar sample. Combining all radio-detected Seyferts, LINERs and composite LINER/H{\\sc ii} galaxies (LTS sources), we obtain an overall radio detection rate of 54\\% (22\\% of all bright nearby galaxies) and we estimate that at least $\\sim$50\\% ($\\sim$20\\% of all bright nearby galaxies) are true AGN. The radio powers of the LTS galaxies allow the construction of a local radio luminosity function. By comparing the luminosity function with those of selected moderate-redshift AGN, selected from the 2dF/NVSS survey, we find that LTS sources naturally extend the RLF of powerful AGN down to powers of about 10 times that of Sgr A*. ", "introduction": "The search for low-luminosity active galactic nuclei (LLAGN) in nearby galaxies has been the subject of many optical surveys. Results show that nuclear activity may be a common phenomenon. The Palomar survey (Ho, Filippenko \\& Sargent 1995, 1997a, b) has been very useful in this regard by providing a sensitive magnitude-limited (B$_{\\rm T}<$12.5~mag) sample of almost 500 bright nearby galaxies. About half of the sources are emission-line nuclei, classified as Seyferts, LINERs or composite LINER/H{\\sc ii} galaxies, the last category displaying both LINER and H{\\sc ii} properties. However, characterizing the powering mechanisms of the sources is not straightforward, particularly in low-luminosity sources. Many of these galaxies possess circum-nuclear star-forming regions which blend with and may even drown out the presence of a weak active galactic nucleus (AGN). Optimally it is necessary to pick spectral regions where the contrast between any hypothetical LLAGN component and circum-nuclear stellar component is maximized. X-rays are very useful in this regard as shown by the hard X-ray studies of LLAGN (Terashima \\etal 2000; Terashima, Ho \\& Ptak 2000; Ho \\etal 2001; Terashima \\& Wilson 2003). In the absence of spectral AGN signatures such as a Seyfert or quasar-type continuum or broad emission lines, radio observations can offer an alternative method for determining the LLAGN incidence in nearby galaxies. Measurements of radio flux, compactness, radio spectral index\\footnote{F$_{\\nu} \\propto \\nu^{-\\alpha}$ throughout.} and brightness temperatures provide the necessary diagnostic tools for determining the nature of the radio emission. \\begin{table*} \\setcounter{table}{0} \\scriptsize \\begin{center} \\begin{minipage}[!ht]{119mm} \\caption{Target MERLIN sources. Col. 1 Source name. Col. 2 and 3: Optical position from NASA/IPAC Extragalactic Database (NED). Col. 4: Adopted distance from Ho, Filippenko \\& Sargent (1997a) with H$_{\\rm 0}$=75 km s$^{-1}$ Mpc$^{-1}$. Col. 5: Spectral Type from Ho, Filippenko \\& Sargent (1997a). L = LINERs, S = Seyferts, and T = composite LINER/H{\\sc ii} galaxies. Colons refer to uncertain (:) or highly uncertain (::) spectral classification. The number 1.9 refers to the presence of broad H$\\alpha$ detection and 2 refers to the absence of broad H$\\alpha$ emission. Col. 6: Hubble type from Ho, Filippenko \\& Sargent (1997a).} \\begin{tabular}[h!]{l c c c c c r} \\hline \\hline & RA(J2000) & Dec(J2000) & $D$ & & \\\\ Galaxy & (\\hrs \\ramin \\rasec) & (\\degree \\arcmin \\arcsec~) & (Mpc) & Spectral Type & Hubble Type \\\\ (1) & (2) & (3) & (4) & (5) & (6) \\\\ \\hline IC\\,356 & 04 07 46.9 & $+$69 48 45 & 18.1 & T2 & SA(s)ab pec \\\\ IC\\,520 & 08 53 42.2 & $+$73 29 27 & 47.0 & T2: & SAB(rs)ab? \\\\ NGC\\,428 & 01 12 55.6 & $+$00 58 54 & 14.9 & L2/T2: & SAB(s)m \\\\ NGC\\,488 & 01 21 46.8 & $+$05 15 25 & 29.3 & T2:: & SA(r)b \\\\ NGC\\,521 & 01 24 33.8 & $+$01 43 52 & 67.0 & T2/H: & SB(r)bc \\\\ NGC\\,718 & 01 53 13.3 & $+$04 11 44 & 21.4 & L2 & SAB(s)a \\\\ NGC\\,777 & 02 00 14.9 & $+$31 25 46 & 66.5 & S2/L2:: & E1 \\\\ NGC\\,841 & 02 11 17.3 & $+$37 29 50 & 59.5 & L1.9: & (R')SAB(s)ab \\\\ NGC\\,1169 & 03 03 34.7 & $+$46 23 09 & 33.7 & L2 & SAB(r)b \\\\ NGC\\,1961 & 05 42 04.8 & $+$69 22 43 & 53.1 & L2 & SAB(rs)c\\\\ NGC\\,2336 & 07 27 03.7 & $+$80 10 42 & 2.9 & L2/S2 & SAB(r)bc\\\\ NGC\\,2681 & 08 53 32.8 & $+$51 18 50 & 13.3 & L1.9 & (R')SAB(rs)0/a \\\\ NGC\\,2768 & 09 11 37.5 & $+$60 02 15 & 23.7 & L2 & E6: \\\\ NGC\\,2832 & 09 19 46.9 & $+$33 44 59 & 91.6 & L2:: & E$+$2: \\\\ NGC\\,2841 & 09 22 02.6 & $+$50 58 35 & 12.0 & L2 & SA(r)b: \\\\ NGC\\,2985 & 09 50 21.6 & $+$72 16 44 & 22.4 & T1.9 & (R')SA(rs)ab \\\\ NGC\\,3166 & 10 13 45.6 & $+$03 25 32 & 22.0 & L2 & SAB(rs)0/a \\\\ NGC\\,3169 & 10 14 15.0 & $+$03 27 57 & 19.7 & L2 & SA(s)a pec \\\\ NGC\\,3190 & 10 18 05.8 & $+$21 49 56 & 22.4 & L2 & SA(s)a pec spin \\\\ NGC\\,3226 & 10 23 27.0 & $+$19 53 54 & 23.4 & L1.9 & E2: pec \\\\ NGC\\,3301 & 10 36 55.8 & $+$21 52 55 & 23.3 & L2 & (R')SB(rs)0/a \\\\ NGC\\,3368 (M\\,96) & 10 46 45.7 & $+$11 49 12 & 8.1 & L2 & SAB(rs)ab \\\\ NGC\\,3414 & 10 51 16.2 & $+$27 58 30 & 24.9 & L2 & S0 pec \\\\ NGC\\,3433 & 10 52 03.9 & $+$10 08 54 & 39.5 & L2/T2:: & SA(s)c \\\\ NGC\\,3507 & 11 03 25.4 & $+$18 08 12 & 19.8 & L2 & SB(s)b \\\\ NGC\\,3607 & 11 16 54.3 & $+$18 03 10 & 19.9 & L2 & SA(s)0: \\\\ NGC\\,3623 (M\\,65) & 11 18 55.9 & $+$13 05 32 & 7.3 & L2: & SAB(rs)a \\\\ NGC\\,3626 & 11 20 03.8 & $+$18 21 25 & 26.3 & L2: & (R)SA(rs)0$+$ \\\\ NGC\\,3627 (M\\, 66) & 11 20 15.0 & $+$12 59 30 & 6.6 & T2/S2 & SAB(s)b \\\\ NGC\\,3628 & 11 20 17.0 & $+$13 35 22 & 7.7 & T2 & SAb pec spin \\\\ NGC\\,3646 & 11 22 14.7 & $+$20 12 31 & 56.8 & H/T2: & SB(s)a \\\\ NGC\\,3675 & 11 26 07.9 & $+$43 35 10 & 12.8 & T2 & SA(S)b \\\\ NGC\\,3718 & 11 32 35.3 & $+$53 04 01 & 17.0 & L1.9 & SB(s)a pec \\\\ NGC\\,3780 & 11 39 22.3 & $+$56 16 14 & 37.2 & L2:: & SA(s)c: \\\\ NGC\\,3884$^a$ & 11 46 12.2 & $+$20 23 30 & 91.6 & L1.9 & SA(r)0/a; \\\\ NGC\\,3898 & 11 49 15.2 & $+$56 05 04 & 21.9 & T2 & SA(s)ab; \\\\ NGC\\,3945 & 11 53 13.6 & $+$60 40 32 & 22.5 & L2 & SB(rs)0$+$ \\\\ NGC\\,4013 & 11 58 31.3 & $+$43 56 49 & 17.0 & T2 & SAb \\\\ NGC\\,4036 & 12 01 26.9 & $+$61 53 44 & 24.6 & L1.9 & S0$-$ \\\\ NGC\\,4143 & 12 09 36.1 & $+$42 32 03 & 17.0 & L1.9 & SAB(s)0 \\\\ NGC\\,4203 & 12 15 05.0 & $+$33 11 50 & 9.7 & L1.9 & SAB0$-$: \\\\ NGC\\,4293 & 12 21 12.8 & $+$18 22 58 & 17.0 & L2 & (R)SB(s)0/a \\\\ NGC\\,4321 (M\\,100) & 12 22 54.9 & $+$15 49 21 & 16.8 & T2 & SAB(s)bc \\\\ NGC\\,4414 & 12 26 27.1 & $+$31 13 24 & 9.7 & T2: & SA(rs)c? \\\\ NGC\\,4435 & 12 27 40.5 & $+$13 04 44 & 16.8 & T2/H: & SB(s)0 \\\\ NGC\\,4438 & 12 27 45.6 & $+$13 00 32 & 16.8 & L1.9 & SA(s)0/a \\\\ NGC\\,4450 & 12 28 29.5 & $+$17 05 06 & 16.8 & L1.9 & SA(s)ab \\\\ NGC\\,4457 & 12 28 59.1 & $+$03 34 14 & 17.4 & L2 & (R)SAB(s)0/a \\\\ NGC\\,4548 (M\\,91) & 12 35 26.4 & $+$14 29 47 & 16.8 & L2 & SBb(rs) \\\\ NGC\\,4589 & 12 37 25.0 & $+$74 11 31 & 30.0 & L2 & E2 \\\\ NGC\\,4636 & 12 42 50.0 & $+$02 41 17 & 17.0 & L1.9 & E0$+$ \\\\ NGC\\,4736 (M\\,94) & 12 50 53.0 & $+$41 07 14 & 4.3 & L2 & (R)SA(r)ab \\\\ NGC\\,4750 & 12 50 07.1 & $+$72 52 30 & 26.1 & L1.9 & (R)SA(rs)ab \\\\ NGC\\,4772 & 12 53 29.0 & $+$02 10 02 & 16.3 & L1.9 & SA(s)a \\\\ NGC\\,4826 (M\\,64) & 12 56 43.7 & $+$21 40 52 & 4.1 & T2 & (R)SA(rs)ab \\\\ NGC\\,5077$^b$ & 13 19 31.6 & $-$12 39 26 & 40.6 & L1.9 & E3$+$\\\\ NGC\\,5297 & 13 46 23.7 & $+$43 52 20 & 37.8 & L2 & SAB(s)c: spin \\\\ NGC\\,5322 & 13 49 15.2 & $+$60 11 26 & 31.6 & L2:: & E3$+$ \\\\ NGC\\,5353 & 13 53 26.7 & $+$40 16 59 & 37.8 & L2/T2: & SA0 spin \\\\ NGC\\,5363 & 13 56 07.1 & $+$05 15 20 & 22.4 & L2 & IA0? \\\\ NGC\\,5371 & 13 55 39.9 & $+$40 27 43 & 37.8 & L2 & SAB(rs)bc \\\\ NGC\\,5377 & 13 56 16.6 & $+$47 14 08 & 31.0 & L2 & (R)SB(s)a \\\\ NGC\\,5448 & 14 02 49.7 & $+$49 10 21 & 32.6 & L2 & (R)SAB(r)a \\\\ NGC\\,5678 & 14 32 05.8 & $+$57 55 17 & 35.6 & T2 & SAB(rs)b \\\\ NGC\\,5813 & 15 01 11.2 & $+$01 42 08 & 28.5 & L2: & E1$+$ \\\\ NGC\\,5838 & 15 05 26.2 & $+$02 05 58 & 28.5 & T2:: & SA0$-$ \\\\ NGC\\,6340 & 17 10 24.9 & $+$72 18 16 & 22.0 & L2 & SA(s)0/a \\\\ NGC\\,6501 & 17 56 03.7 & $+$18 22 23 & 39.6 & L2:: & SA0$+$: \\\\ NGC\\,6702 & 18 46 57.6 & $+$45 42 20 & 62.8 & L2:: & E: \\\\ NGC\\,6703 & 18 47 18.8 & $+$45 33 02 & 35.9 & L2:: & SA0$-$ \\\\ \\hline \\end{tabular} \\smallskip $^a$ Does not fulfill the magnitude criterion for the Palomar sample. $^b$ Does not fulfill declination criterion for the Palomar sample. \\end{minipage} \\end{center} \\end{table*} \\normalsize Several important radio surveys have been conducted on the magnitude-limited Palomar bright nearby galaxy sample (Ho, Filippenko \\& Sargent 1995, 1997a, b), revealing a large fraction of radio cores, not only in ellipticals but also in bulge-dominated spirals. In a recent Very Large Array (VLA) 5 and 1.4\\,GHz, 1\\arcsec~resolution survey of the low-luminosity Seyferts of the Palomar sample (Ho \\& Ulvestad 2001; HU01 hereafter), it was found that over 80\\% of the sources harbour a radio core. In a distance-limited sample of low- luminosity Seyferts, LINERs and composite LINER/H{\\sc ii} galaxies observed with the VLA at 15\\,GHz, 0\\arcsecpoint25 resolution (Nagar \\etal 2000, 2002; VLA/N00 and VLA/N02 hereafter; Nagar, Falcke \\& Wilson 2005; VLA/N05 hereafter), it was found that $\\sim$40\\% of the objects harbour subarcsecond-scale compact radio cores. A similar study with the VLA, at 8.4\\,GHz, 2\\arcsecpoint5~resolution of all the composite LINER/H{\\sc ii} galaxies in the Palomar sample has been presented in Filho, Barthel \\& Ho (2000, 2002a), revealing radio cores in $\\sim$25\\% of the sample sources. However, although the radio core emission in these sources is consistent with the presence of a LLAGN, we cannot exclude a stellar origin from the brightness temperature figures (T$_{\\rm B}$\\lax10$^{5}$~K; Condon 1992) obtained at these resolutions. As conclusive judgement requires Very Long Baseline (VLBI)-resolution, multi-wavelength Very Long Baseline Array (VLBA) and European Very Long Baseline Interferometer Network (EVN) observations have been obtained for selected subsamples of low-luminosity Seyferts, LINERs and composite LINER/H{\\sc ii} galaxies that showed arcsecond- or subarcsecond-scale radio cores (Falcke \\etal 2000; F00 hereafter; Nagar \\etal 2002; VLBA/N02 hereafter; Nagar, Falcke \\& Wilson 2005; VLBA/N05 hereafter; Ulvestad \\& Ho 2001b; Filho, Barthel \\& Ho 2002b; Filho \\etal 2004; Anderson, Ulvestad \\& Ho 2004; AU04 hereafter). In sources with subarcsecond- or arcsecond-scale radio peak emission above 2.5~mJy, results reveal a 100\\% detection rate of high-brightness temperature (T$_{\\rm B}$\\gax10$^8$~K), compact, flat spectrum ($\\alpha$$<$0.5) radio cores, enforcing the LLAGN scenario for the radio emission (VLBA/N05; see also Ulvestad \\& Ho 2001b; Filho, Barthel \\& Ho 2002b; Filho \\etal 2004; AU04). Their low radio luminosities suggest we are probing the very faint end of the AGN population. Unambiguously determining the physical nature of the nearby galaxy radio cores is more than of mere phenomenological interest. If they truly contain an accretion-powered nucleus, then they obviously need to be included in the AGN population. Their non-trivial numbers impact on several astrophysical problems ranging from the cosmological evolution of the AGN luminosity function to their contribution to the X-ray background. The present paper deals with high-resolution radio-imaging of LLAGN, carried out with the Jodrell Bank Multi-Element Radio-Linked Interferometer Network (MERLIN), completing the radio census of nuclear activity in the Palomar galaxy survey of Ho, Filippenko \\& Sargent (1997a). ", "conclusions": "We have undertaken a MERLIN survey of nearby galaxies that did not have available 2\\arcsecpoint5~resolution or better radio observations. Results reveal a 21\\% radio-detection rate among the sources, with fifteen radio detections, three of which are new AGN candidates. A compilation of radio observations of all low-luminosity Seyferts, LINERs and composite LINER/H{\\sc ii} galaxies in the magnitude-limited Palomar survey reveal a radio-detection rate of 54\\% (or 22\\% of all bright nearby galaxies), with a more than 50\\% detection rate (or 20\\% for all bright nearby galaxies) of low-luminosity active nuclei. The radio detection of the Seyferts, LINERs and composite LINER/H{\\sc ii} sources in the Palomar survey allow the construction of a local radio luminosity function. Our results show that the Seyferts, LINERs and composite LINER/H{\\sc ii} sources form a smooth luminosity transition from higher redshift, more luminous AGN as sampled by the 2dF/NVSS survey." }, "0601/astro-ph0601049_arXiv.txt": { "abstract": "{This paper presents a method to determine effective temperatures, angular semi-diameters and bolometric corrections for population I and II FGK type stars based on $V$ and \\MASS IR photometry. Accurate calibration is accomplished by using a sample of solar analogues, whose average temperature is assumed to be equal to the solar effective temperature of 5777 K. By taking into account all possible sources of error we estimate associated uncertainties better than 1\\% in effective temperature and in the range 1.0--2.5\\% in angular semi-diameter for unreddened stars. Comparison of our new temperatures with other determinations extracted from the literature indicates, in general, remarkably good agreement. These results suggest that the effective temperaure scale of FGK stars is currently established with an accuracy better than 0.5\\%--1\\%. The application of the method to a sample of 10999 dwarfs in the Hipparcos catalogue allows us to define temperature and bolometric correction ($K$ band) calibrations as a function of $(V-K)$, $[m/H]$ and \\logg. Bolometric corrections in the $V$ and $K$ bands as a function of \\teft, $[m/H]$ and \\logg are also given. We provide effective temperatures, angular semi-diameters, radii and bolometric corrections in the $V$ and $K$ bands for the 10999 FGK stars in our sample with the corresponding uncertainties. ", "introduction": "Effective temperature and luminosity are two fundamental stellar parameters that are crucial to carry out tests of theoretical models of stellar structure and evolution by comparing them with observations. The accuracy in the determination of other stellar properties, such as metallicity, age or radius, hinges on our ability to estimate the effective temperatures and luminosities. There are several approaches in the literature to compute effective temperature and/or luminosity. Except when applied to the Sun, very few of them are {\\it direct} methods that permit an empirical measurement of these parameters. Usually, {\\it semi-empirical} or {\\it indirect} methods are based to a certain extent on stellar atmosphere models. Among the {\\it direct} approaches we find the remarkable work by \\citet{code76}, which is based on interferometric measurements of stellar angular semi-diameters ($\\theta$) and total fluxes ($F_{\\rm bol}$) at Earth, and the more recent works of \\citet{mozurkewich03} and \\citet{kervella04}, also based on interferometry. On the other hand, {\\it indirect} methods are mainly based on the use of photometry, spectroscopy, or a combination of both. In the case of the temperatures, although many of the published calibrations claim to have uncertainties of the order of several tens of degrees, values obtained by different authors can easily have discrepancies of 100 K or even larger. The reason for such nagging differences must be found somewhere in the {\\it ingredients} of the methods: atmosphere models, absolute flux calibrations, oscillator strengths, calibration stars, etc. In this paper we present a {\\it semi-empirical} method to determine effective temperatures (\\teft) and bolometric corrections ($BC$) from 2MASS\\footnote{\\tt http://www.ipac.caltech.edu/2mass} $JHK$\\footnote {Throughout the paper, $K$ refers to $K_s$ band.} photometry \\citep{cutri03b} that is applicable to FGK type stars. As all others, our method is susceptible to problems derived from the uncertainties in the ingredients mentioned above. However, our approach benefits from two major features: First, it provides a way to evaluate realistic individual uncertainties in \\teft, $\\theta$ and luminosity by considering all the involved errors; and second, as it is calibrated to use the \\MASS photometry, it allows the calculation of consistent and homogeneous \\tefts and $BC$ for several million stars in the \\MASS catalogue. This paper also provides \\teft, angular semi-diameters, radii and $BC$s for 10999 dwarfs and subdwarfs in the Hipparcos catalogue \\citet{ESA97}. Such large sample has allowed us to construct simple parametric calibrations as a function of $(V-K)_0$, \\feh and \\logG. Note that a preliminary version of the method presented here was already successfully applied to the characterization of the properties of planet-hosting stars \\citep{iribas03a}. The present paper is organized as follows. Section \\ref{sedf} presents the method and explains in detail the procedure to obtain \\tefts and angular semi-diameters, including the fitting algorithm, zero point corrections and error estimates. The comparison of our temperatures with several previous works, both based on photometric and spectroscopic techniques, is described in Sect. \\ref{comparacions}. In Sect. \\ref{calib} we present simple parametric calibrations of \\tefts and $BC$ as a function of $(V-K)_0$, $[m/H]$ and \\logg valid for dwarf and subdwarf stars. The sample of 10999 stars used to build the calibrations is also described in this section together with a detailed explanation of the different contributors to the final uncertainties. Finally, the results are discussed in Sect. \\ref{disc} and the conclusions of the present work are presented in Sect. \\ref{conclu}. ", "conclusions": "\\label{conclu} We have presented a method (called SEDF) to compute effective temperatures, angular semi-diameters and bolometric corrections from \\MASS photometry. We have adopted an approach based on the fit of the observed $VJHK$ magnitudes using synthetic photometry, and it yields accuracies around 1\\% in \\teft, 2\\% in $\\theta$, and 0.05 mag in $BC$, in the temperature range 4000--8000 K. A zero point offset was added to the synthetic photometry computed from the Kurucz atmosphere models to tie in our temperature scale with the Sun's temperature through a sample of solar analogues. From the application to a large sample of FGK Hipparcos dwarfs and subdwarfs we provide parametric calibrations for both effective temperature and bolometric correction as a function of $(V-K)_0$, $[m/H]$ and \\logG. Note that the method presented here has been selected as one of the main sources of effective temperatures to characterize the primary and secondary targets of the COROT space mission \\citep {baglin00}. Also, it is being currently implemented as one of the tools offered by the Spanish Virtual Observatory \\citep{solano05}. The resulting temperatures have been compared with several photometric and spectroscopic determinations. Although we obtained remarkably good agreement, slight systematic differences with other semi-empirical methods, such as the IRFM, are present. This is probably due to the uncertainties in the absolute flux calibration used by different techniques. It is possible that, in spite of the great effort carried out by \\citet{cohen03a} and others to construct a consistent absolute flux calibration in both the optical and the IR regions, some problems still remain, which introduce small systematic effects in the temperatures. However, these effects seem to be as small as 20--30 K and could be explained through uncertainties in the IR fluxes of about 2\\%. In conclusion, the results presented here strongly suggest that, given the small differences found between methods, the effective temperature scale of FGK stars (4000--8000 K) is currently established with a net accuracy better than 0.5--1.0\\%." }, "0601/astro-ph0601519_arXiv.txt": { "abstract": "We present an efficient second order accurate scheme to treat stiff source terms within the framework of higher order Godunov's methods. We employ Duhamel's formula to devise a modified predictor step which accounts for the effects of stiff source terms on the conservative fluxes and recovers the correct isothermal behavior in the limit of an infinite cooling/reaction rate. Source term effects on the conservative quantities are fully accounted for by means of a one-step, second order accurate semi-implicit corrector scheme based on the deferred correction method of Dutt et. al. We demonstrate the accurate, stable and convergent results of the proposed method through a set of benchmark problems for a variety of stiffness conditions and source types. ", "introduction": "We wish to solve the following system of partial differential equations describing a hydrodynamic flow with a stiff (energy) source term \\begin{equation} \\label{hypsys:eq} \\frac{\\partial U}{\\partial t} + \\sum _{d=1}^{D} \\frac{\\partial F_d(U)}{\\partial x_d} = S(U) \\end{equation} where $D$ is the dimensionality of the problem, $U,~F(U),~S(U)$ are the conservative variables, the conservative fluxes and the source term respectively, given by \\begin{equation} U= \\begin{pmatrix} \\rho \\\\ \\rho u_1 \\\\ \\vdots \\\\ \\rho u_D \\\\ \\rho E \\end{pmatrix} ; \\quad F_d(U)= \\begin{pmatrix} \\rho u_d \\\\ \\rho u_1 u_d + p \\,\\delta_{1d} \\\\ \\vdots \\\\ \\rho u_D u_d + p \\,\\delta_{Dd} \\\\ (\\rho E + p) u_d \\end{pmatrix}; \\quad S(U)= \\begin{pmatrix} 0 \\\\ 0 \\\\ \\vdots \\\\ 0 \\\\ \\rho\\Lambda(e,\\rho) \\end{pmatrix}. \\label{eq:ugs} \\end{equation} In the above equations, $\\rho$ is the density, $u_d$ the velocity in the $d$ direction, $E=e+ \\sum_{d=1}^Du_d^2/2$ is the total specific energy with, $e$, the specific internal energy. $\\Lambda(e,\\rho)$ is the term describing the source of specific internal energy. In the following we consider the case of a stiff source term corresponding to an endothermic process, such as occurs in radiative losses. In addition, we restrict our analysis to source types that, at least near equilibrium, behave as a relaxation law. In the stiff case, the characteristic relaxation time scale for $S$ may be much smaller than the CFL time step for the hydrodynamic waves. For that reason, we would like to use a semi-implicit method, treating the stiff source term implicitly, while using an explicit method for the hyperbolic terms. However, the classical analysis of such fast endothermic processes shows that, in the limit as the relaxation time goes to zero, the gas can be described by the compressible flow equations with an isothermal equation of state \\cite{vincentiKruger65}. Pember \\cite{pember93} showed that the use of formally second-order accurate semi-implicit methods such as Strang splitting, or a second-order Godunov predictor-corrector method could lead to a substantial loss of accuracy, due to inconsistencies between the the characteristic tracing step without sources and the effective limiting isothermal behavior. Such inconsistencies between the flux calculation with the limiting isothermal equation of state can lead to dramatic errors particularly at sonic points. Pember proposed various approaches to the problem based on classical relaxation theory. Roe and Hittinger \\cite{roehit01} also addressed the issues raised here in relation to Godunov's method with stiff relaxation. In their approach they split the equations based on a splitting of state space into stiff and non-stiff subspaces of the linearized source term to obtain in the stiff limit formulations similar to ours. However, neither Pember nor Roe and Hittinger did present a complete method that is second-order accurate in both the stiff and non-stiff limits, nor did they discuss the extension to more than one dimension. The problem of hyperbolic system with stiff relaxation has also been considered by other authors in the past mostly for one-dimensional systems and within the framework of Runge-Kutta based methods of lines. In particular Jin \\cite{jin95} designed second order Runge-Kutta type splitting methods with the correct asymptotic limit. Jin and Levermore's \\cite{jinlevermore96} developed a semidiscrete high resolution method which, in order to ensure the correct asymptotic behavior, employes a linear combination of the conservative fluxes for the homogeneous (i.e. without the relaxation term) and equilibrium system. The fluxes are computed with a higher order Godunov's method and the scheme allows for a rapid transition between the stiff and nonstiff regimes. As the authors point out, however, the upwind property of the scheme is not strictly guaranteed for all stiffness conditions. Finally, Caflisch, Jin and Russo \\cite{caflisch97} developed a scheme for hyperbolic systems with relaxation that is uniformly accurate for various ranges of stiffness conditions (see also Ref.~\\cite{jinpareschitoscani98,pareschirusso05}). The aim of this paper is to build a higher order Godunov's method that preserves the properties of robustness and accuracy across a variety of stiffness conditions thus avoiding the problems described in \\cite{pember93}. In particular, in order to preserve higher order accuracy, we aim for a semi-implicit method that corresponds to a standard second-order Godunov method of the appropriate hyperbolic problem for the stiff or non-stiff limits. To this end, we use second-order accurate deferred corrections method of a type presented in \\cite{dugrro00}, to obtain a semi-implicit corrector that is a special case of the algorithms described in \\cite{minion03}, although any implicit L-stable second-order one-step method would be acceptable. The main new idea in our work is contained in our treatment of the predictor step for computing the hyperbolic fluxes, based on the derivation of a local effective dynamics using Duhamel's formula. This leads to an explicit predictor step that corresponds to that for a conventional second-order Godunov method for Eq.~(\\ref{hypsys:eq}) in the limit where the relaxation time is comparable to or greater than the hydrodynamic CFL time step; and to a second-order Godunov method for the isothermal equations in the limit where the relaxation time is much smaller than the hydrodynamic time step. Our approach is similar to that used in \\cite{tcm05} for obtaining a well-behaved numerical method for incompressible viscoelastic flows in both the viscous and elastic limits; however, the details there are quite different than those for the present setting. The paper is organized as follows. In section~\\ref{sdc:se} we describe a second order accurate, semi-implicit corrector method based on the deferred corrections ideas presented in~\\cite{dugrro00,minion03} to be used for the final source term update. In section~\\ref{mgf:se}, based on Duhamel's formula, we work out a modified formulation of Godunov's predictor step and flux calculation suitable for the case of stiff source terms. In section~\\ref{sa:se} we discuss stability issues for our approach, and Section~\\ref{lrnz:se} contains the extension of the method to the case in which the source term depends both on the gas density as well as the internal energy. In section~\\ref{test:se} we test the performance of the code and demonstrate the accuracy of the method in various stiffness conditions. The paper concludes with section~\\ref{concl:se} where the main results of the paper are summarized. ", "conclusions": "\\label{concl:se} We have presented a second order accurate semi-implicit predictor-corrector scheme to treat stiff source terms within the framework of higher order Godunov's methods. Our treatment of the predictor step for computing the hyperbolic fluxes, is based on the derivation of a local effective dynamics using Duhamel's formula. This leads to a conventional second-order Godunov method when the system relaxation time is larger than the time step and to a second-order Godunov method for the isothermal equations in the limit of a stiff source term. Finally, we obtain a semi implicit corrector using a one-step second-order accurate deferred corrections method as suggested in \\cite{dugrro00,minion03}. Our tests indicate that the proposed method is stable, robust and its second order accuracy preserved across a variety of stiffness conditions. We have also discussed the case of a general source term which depends both on $e$ and $\\rho$ and shown that the method is applicable provided that the flow is thermally stable or the non-stiff part of the source term is resolved in time. The additional cost involved in the formulation of our scheme is minimal; all it requires is an estimate of the term $\\Lambda_e$ which in a purely relaxation case is trivial and for a more complicated source term (such as the case of radiative losses) is still minor compared to the estimate of the source term itself. In our implementation the factor $\\alpha(\\gamma-1)+1$ is stored as an additional primitive variable and used as polytropic index in the characteristic analysis in stead of $\\gamma$. \\vskip 1truecm \\leftline{\\bf Acknowledgment} FM is grateful to the Lawrence Berkeley National Laboratory for its hospitality and acknowledges support by the Swiss Institute of Technology through a Zwicky Prize Fellowship. PC was supported by the Mathematical, Information, and Computing Sciences Division of the United States Department of Energy Office of Science under contract number DE-AC02-05CH11231." }, "0601/astro-ph0601033_arXiv.txt": { "abstract": "We determine the disk mass distribution around 336 stars in the young ($\\sim 1$ Myr) Orion Nebula cluster by imaging a $2\\rlap{.}'5 \\times 2\\rlap{.}'5$ region in $\\lambda$3 mm continuum emission with the Owens Valley Millimeter Array. For this sample of 336 stars, we observe 3 mm emission above the 3$\\sigma$ noise level toward ten sources, six of which have also been detected optically in silhouette against the bright nebular background. In addition, we detect 20 objects in 3 mm continuum emission that do not correspond to known near-IR cluster members. Comparisons of our measured fluxes with longer wavelength observations enable rough separation of dust emission from thermal free-free emission, and we find substantial dust emission toward most objects. For the sample of ten objects detected at both 3 mm and near-IR wavelengths, eight exhibit substantial dust emission. Excluding the two high-mass stars ($\\theta^1$OriA and the BN object) and assuming a gas-to-dust ratio of 100, we estimate circumstellar masses ranging from 0.13 to 0.39 M$_\\odot$. For the cluster members not detected at 3 mm, images of individual objects are stacked to constrain the mean 3 mm flux of the ensemble. The average flux is detected at the 3$\\sigma$ confidence level, and implies an average disk mass of 0.005 M$_{\\odot}$, comparable to the minimum mass solar nebula. The percentage of stars in Orion surrounded by disks more massive than $\\sim 0.1$ M$_{\\odot}$ is consistent with the disk mass distribution in Taurus, and we argue that massive disks in Orion do not appear to be truncated through close encounters with high-mass stars. Comparison of the average disk mass and number of massive dusty structures in Orion with similar surveys of the NGC 2024 and IC 348 clusters is used to constrain the evolutionary timescales of massive circumstellar disks in clustered environments. ", "introduction": "Over the last two decades, high resolution millimeter, infrared, and optical images have provided evidence for the existence of circumstellar disks on scales of $\\sim 0.1$--1000 AU around young stars \\citep[e.g.,][]{KS95,DUTREY+96, PADGETT+99,OW96,EISNER+04}. Circumstellar disks are the likely birth-sites for planetary systems, and determining their ubiquity, properties, and lifetimes is crucial for constraining the timescales and mechanisms of planet formation. The mass distribution of protoplanetary disks is especially important since disk mass is related to the mass of planets that may potentially form. The minimum-mass protosolar nebula, $\\sim 0.01$ M$_{\\odot}$ \\citep{WEID+77,HAYASHI81}, is an informative benchmark against which to compare disk masses around young stars, and such comparisons can constrain the number of protoplanetary disks with the potential to form planetary systems like our own solar system. While direct imaging at optical through near-IR wavelengths has provided concrete evidence for a limited number of circumstellar disks \\citep[e.g.,][]{OW96}, and observations of near-IR excess emission have shown statistically that most young stars with ages less than a few million years still possess inner circumstellar disks \\citep[e.g.,][]{STROM+89,HLL01b}, these studies did not constrain the disk mass distribution. To probe the bulk of the disk mass, which resides in cooler, outer disk regions, observations of optically-thin millimeter emission are needed. Several investigators have carried out comprehensive single-dish mm and sub-mm continuum surveys toward regions of star formation comprising loose aggregates of stars: Taurus \\citep{BECKWITH+90,OB95,MA01,AW05}, $\\rho$ Ophiuchi \\citep{AM94,NUERNBERGER+98, MAN98}, Lupus \\citep{NCZ97}, Chamaeleon I \\citep{HENNING+93}, Serpens \\citep{TS98}, and MBM 12 \\citep{ITOH+03,HOGERHEIJDE+02}. In Taurus, 37\\% of stars appear to possess disks more massive than $\\sim 0.01$ M$_{\\odot}$, and the median disk mass is $5 \\times 10^{-3}$ M$_{\\odot}$ \\citep{AW05}. The fraction of massive disks and the median disk mass is comparable in $\\rho$ Ophiuchi \\citep{AM94}. Expanding millimeter continuum surveys to include rich clusters allows the determination of accurate statistics on the frequency and evolution of massive disks as a function of both stellar mass and age. Also, since most stars in the Galaxy form in rich clusters \\citep{LADA+91,LSM93,CARPENTER00,LL03}, understanding disk formation and evolution in cluster environments is a vital component in our general understanding of how stars and planets form. The main challenge to observing rich clusters at mm-wavelengths is that very high angular resolution is required to resolve individual sources and to distinguish compact disk emission from the more extended emission of the molecular cloud. Single-aperture mm-wavelength telescopes lack sufficient angular resolution, and to date, only three rich clusters have been observed with mm-wavelength interferometers: the Orion Nebula cluster \\citep{MLL95,BALLY+98,WAW05}, IC 348 \\citep{CARPENTER02}, and NGC 2024 \\citep{EC03}. The enhanced sensitivity of the most recent observations of Orion has enabled the detection of several massive ($\\ga 0.01$ M$_{\\odot}$) disks \\citep{WAW05}, while upper limits from other surveys range from $\\sim 0.025$--0.17 M$_{\\odot}$ \\citep{MLL95,BALLY+98}. Moreover, extending detection limits by considering as ensembles the large numbers ($\\ga 100$) of young stars included in the surveys of IC 348 and NGC 2024 allowed estimates of mean disk masses of $\\sim 0.002$ and 0.005 M$_\\odot$, respectively \\citep{CARPENTER00,EC03}. While it appears that an average star aged $\\la 1$ Myr still possesses a massive circumstellar disk, more sensitive observations are necessary to detect directly large numbers of massive disks at a range of ages, and thereby constrain the mass distribution and evolutionary timescales. Here, we present a new mm-wavelength interferometric survey of the Orion Nebula cluster (ONC), a young, deeply embedded stellar cluster that includes the bright, massive Trapezium stars. The Trapezium region contains hundreds of stars within several arcminutes, and pre-main-sequence evolutionary models \\citep[e.g.,][]{DM94} fitted to spectroscopic and/or photometric data indicate that most stars are less than approximately one million years old \\citep[e.g.,][]{PROSSER+94,HILLENBRAND97}. Moreover, the standard deviation in the distribution of inferred stellar ages is $\\la 1$ Myr \\citep{HILLENBRAND97}. Our observations thus provide a snapshot of millimeter emission around a large number of roughly coeval young stars. With the large number of stars in the ONC, we can begin to investigate the correlation of disk properties with stellar and/or environmental properties. Previous investigations of near-IR excess emission have explored the dependence of inner disk properties on stellar mass, age, and environment \\citep[e.g.,][]{HILLENBRAND+98,LADA+00}. For example, the fraction of stars in Orion exhibiting near-IR excess emission seems largely independent of stellar age and mass, although there are indications of a paucity of disks around very massive stars \\citep{HILLENBRAND+98,LADA+00}. In addition, the inner disk fraction may decrease at larger cluster radii \\citep{HILLENBRAND+98}. To explore how the properties of the outer disk component correlate with such stellar and environmental properties, millimeter observations of cool, optically-thin dust emission are necessary. Our observations represent an improvement over previous work because our mosaicked image encompasses more than three times as many sources as the previous surveys, enabling an improvement of $\\sqrt{3}$ in estimates of frequency and mean mass of circumstellar disks. Moreover, the comparable sensitivity of this survey with previous observations of IC 348 and NGC 2024 allows a more direct comparison between relatively young (NGC 2024: 0.3 Myr; Meyer 1996; Ali 1996), intermediate (ONC: 1 Myr; Hillenbrand 1997), and old (IC 348: 2 Myr; Luhman et al. 1998; Luhman 1999) clusters, providing constraints on timescales for disk evolution. ", "conclusions": "} We observed a $2\\rlap{.}'5 \\times 2\\rlap{.}'5$ region of the Orion Nebula cluster in $\\lambda$3 mm continuum emission with the Owens Valley Millimeter Array. The mosaic encompassed 336 young stars in the vicinity of the massive Trapezium stars, and constrained the disk mass distribution for this large number of cluster members. We detected 30 objects in 3 mm continuum emission above the 3$\\sigma$ limit, 10 of which correspond with near-IR cluster members. Six of these, in turn, also correspond with optically-detected proplyds. Comparison of our measured fluxes with longer wavelength observations enabled rough separation of emission due to dust and that due to thermal free-free emission, and we found that the 3 mm emission toward 8 objects likely arises, in part, from dust. We argued that with the exception of the massive stars $\\theta^1$OriA and the BN object, sources detected in both 3 mm and near-IR emission are probably young stars surrounded by disks, and we computed circumstellar masses of $0.13$-$0.39$ M$_{\\odot}$ based on observed 3 mm fluxes (these masses are uncertain by at least a factor of three due to uncertainties in converting flux into mass). Since the vast majority ($\\ga 98\\%$) of near-IR cluster members do not possess disks more massive than $\\sim 0.1$ M$_{\\odot}$, we placed constraints on lower-mass disks by considering the ensemble of 326 non-detected, predominantly low-mass stars. For the ensemble, we computed an average disk mass of 0.005 M$_{\\odot}$, which has a statistical significance of 3$\\sigma$ (although the absolute value of the mass contains larger uncertainties related to converting flux into mass). The average disk in the ONC is thus comparable to the minimum mass solar nebula, suggesting that most young ($\\la 1$ Myr) stars in richly clustered environments (and by extension, most stars in the Galaxy) are surrounded by sufficient circumstellar masses to form solar systems like our own. Furthermore, we found that the frequency of massive disks in the ONC appears similar to that in Taurus, perhaps refuting suggestions that disks in Orion are truncated due to close encounters with the massive Trapezium stars. Finally, we compared the disk mass distributions in three clusters of different ages to begin to constrain the evolutionary timescales of massive disks. Although substantial uncertainties remain, it appears that massive disks may evolve significantly on 1-2 Myr timescales. \\noindent{\\bf Acknowledgments.} JAE is currently supported by a Miller Research Fellowship, and acknowledges past support from a Michelson Graduate Research Fellowship. JMC acknowledges support from the Owens Valley Radio Observatory, which is supported by the National Science Foundation through grant AST-9981546. The authors also wish to thank John Bally for providing HST images of the Orion proplyds." }, "0601/astro-ph0601633.txt": { "abstract": "\\spitzer\\ IRAC and MIPS images of the Trifid Nebula (M20) reveal its spectacular appearance in infrared light, highlighting the nebula's special evolutionary stage. The images feature recently-formed massive protostars and numerous young stellar objects, and a single O star that illuminates the surrounding molecular cloud from which it formed, and unveil large-scale, filamentary dark clouds. The hot dust grains show contrasting infrared colors in shells, arcs, bow-shocks and dark cores. Multiple protostars are detected in the infrared, within the cold dust cores of TC3 and TC4, which were previously defined as Class 0. %because they are the cold dust condensations and presence of %molecular outflows. The cold dust continuum cores of TC1 and TC2 contain only one protostar each; the newly discovered infrared source in TC2 is the driving source of the HH399 jet. The \\spitzer\\ color-color diagram allowed us to identify $\\sim$160 young stellar objects (YSOs) and classify them into different evolutionary stages. The diagram also revealed a unique group of YSOs which are bright at 24$\\mu$m but have the spectral energy distribution peaking at 5-8$\\mu$m. Despite expectation that Class 0 sources would be ``starless\" cores, the \\spitzer\\ images, with unprecedented sensitivity, uncover mid-infrared emission from these Class 0 protostars. The mid-infrared detections of Class 0 protostars show that the emission escapes the dense, cold envelope of young protostars. The mid-infrared emission of the protostars can be fit by two temperatures of 150 and 400 K; the hot core region is probably optically thin in the mid-infrared regime, and the size of hot core is much smaller than that of the cold envelope. %the mid-infrared emission %cannot arise from the same location as the mm-wave emission, and %instead must arise from a much smaller region with less intervening %extinction to the central object which drives the bipolar outflows. The presence of multiple protostars within the cold cores of Class 0 objects implies that clustering occurs at this early stage of star formation. The TC3 cluster shows that the most massive star is located at the center of the cluster and at the bottom of the gravitational-potential well. ", "introduction": "The Trifid Nebula (M20), is one of the best-known astrophysical objects: a classical nebula of ionized gas from an O7 star (HD 164492). The nebula glows in red light, trisected by obscuring dust lanes, with a reflection nebula in the north. The Trifid Nebula is a very young \\ion{H}{2} region with an age of $\\sim 3\\times 10^{5}$ years. The Infrared Space Observatory (ISO) and the Hubble Space Telescope (HST) (Cernicharo et al.\\ 1998; Lefloch \\& Cernicharo 2000; Hester et al.\\ 1999) show the Trifid to be a dynamic, ``pre-Orion\" star forming region containing young stellar objects (YSOs) undergoing episodes of violent mass ejection, and protostars (like HH399) losing mass and energy to the nebula in jets. Four massive (17--60 M$_\\odot$) protostellar cores were discovered, with millimeter-wave observations, in the Trifid (Lefloch \\& Cernicharo 2000). A number of YSO candidates were identified using near-infrared (Rho et al. 2001) and X-ray observations (Rho et al. 2004). We adopted a distance of 1.68 kpc for the Trifid Nebula, which was measured by Lynds et al. (1986). There have been two primary tools for studying massive star formation: molecular line emission which traces dense material and outflows (e.g. Shirley et al. 2003), and dust continuum emission which traces cold, dense material (e.g. William et al. 2004). The youngest objects, Class 0 protostars, are those that are detected as condensations in dust continuum maps characterized by very low values of the ratio L$_{bol}$/L$_{submm}$, and show collimated CO outflows or internal-heating sources (Andr\\'e 1994). The Class 0 protostars were believed to be ``starless\" cores, with neither near-infrared nor mid-infrared ($<$20$\\mu$m) emission. In this paper, we present spectacular \\spitzer\\ images of the Trifid Nebula and report $\\sim$160 newly identified young stellar objects (YSOs) using their infrared colors. We found multiple protostars within cold cores and many evolved YSOs are located along the ionization fronts. We illustrate that \\spitzer\\ infrared images provide new, excellent tools for studying massive-star formation in ways that were not previously available. ", "conclusions": "The detection of multiple mid-infrared protostars from the cold dust cores is a new result. How does the mid-infrared emission leak through thick envelopes of material infalling onto the stars? Isn't the mid-infrared emission absorbed by the cold and thick envelopes? We can address where the emission arises, how the dust is heated, and how the emission escapes cold cores. The cold envelope sizes of the TC3, TC4, and TC4b cores are 0.2, 0.2, and 0.16 pc, respectively , which were determined by \\citet{lef00}, using the 50\\% contour in the millimeter flux map. The estimated optical depth is: \\begin{equation} \\tau_{1250} = {F_{1250} \\over {\\Omega \\, B_{\\nu}(T)}} = 1.08\\times 10^{-4} F_{1250} {d_{1.68kpc}^2 \\over R^{2} } \\label{eq1} \\end{equation} \\noindent where F$_{1250}$ is the mm-wave flux, $\\Omega$ is a solid angle of the source size, B$_{\\nu}$(T) is a black body Planck function at a temperature of T, and R is a source radius in unit of pc. The optical depths of the TC3, TC4, and TC4b cores are between 0.7-2.5 $\\times 10^{-3}$ (for a temperature of 22 K). The corresponding optical depths at 8$\\mu$m were determined using $\\tau_{8}$ = $\\tau_{1250} \\times (1250/8)^{\\alpha_1}$ $(8/100)^{\\alpha_2}$, where $\\alpha_1=2$ is the extinction slope between 100-3000$\\mu$m and $\\alpha_2=1$ is the extinction slope between 10-100$\\mu$m (Li \\& Draine 2001); this yields $\\tau_{8 \\mu m}$ $\\sim$ 1--7. The corresponding visual extinction was determined using A$_{8{\\mu}m}$/A$_K$ = 0.5 (Draine 2003); this yields a total extinction of A$_v$ $\\sim$ 20--150. Despite the high extinction of the envelope and the high optical depth of the protostars (estimated using a temperature of 22K) at 8$\\mu$m, apparently we observed mid-IR emission from the early protostars. We note that the optical depths of 150 K and 400 K components are much smaller, more than one magnitude smaller than that of the cold (22 K) envelope. From the SED fits (see Fig. \\ref{sedtc4a}) we estimated that the optical depth of the 8$\\mu$m emitting region is $< 10^{-3}$ for the warm and hot temperature components, which suggests that the accretion region is probably optically thin if it is larger than $\\sim 10$ AU. In addition, the \\spitzer\\ images revealed that the infrared sources of the protostars were point-sources, inferring that the regions emitting 400 K and 150 K would be smaller than the PSF radii of the IRAC and MIPS 24$\\mu$m images, which are 0.007 pc (= 1500 AU) and 0.025 pc (= 5$\\times 10^3$ AU), respectively. %which are 0.014 pc (=3000 AU) and 0.05 pc, respectively. This is consistent with the models of the massive protostars \\citep{oso99}; the cold component (responsible for far-infrared and 1300$\\mu$m emission) would be emitted from a region $>$ $10^4$ AU from the central source, while the 400 K component (responsible for mid-infrared emission) would be from $<$ 1500 AU region. The mid-infrared emission can escape the dense envelope; it suffers a modest mid-infrared extinction but traces the emission directly from the accretion region. Previous ISO observations showed similarly mid-infrared counterparts of a few low mass Class 0 objects in other star forming regions \\citep{cer00}. % although the ISO spatial resolution was limited. New \\spitzer\\ observations of Class 0 objects also detected mid-infrared protostars in L1014 (Young et al. 2004) and Cepheus E \\citep{nor04}. The detections infer that the region of hot core in early protostars is optical thin at the mid-infrared regime, and the size of hot core is smaller than the cold envelope. The mid-infrared emission is powered by accretion of envelope material onto the protostar, and most of the bolometric luminosity at the early stage of the protostar is from accretion. The accretion rate is given by \\cite{mck02,mck03}: $$\\dot{m_{*}}= 4.75\\times10^{-4}\\,\\, \\epsilon_{core}^{1/4} (f_{gas} \\,\\phi_P \\,\\alpha_{vir})^{3/8} \\,\\, \\biggl({m_{*f} \\over {30 M_\\odot}}\\biggr)^{3/4} \\Sigma_{cl}^{3/4} \\biggl({m_{*} \\over m_{*f} }\\biggr)^{1/2} \\,\\, M_{\\odot} \\, yr^{-1} $$ \\begin{equation} \\sim 4.6\\times 10^{-4} \\biggl({m_{*f} \\over {30 M_\\odot}}\\biggr)^{3/4} \\Sigma_{cl}^{3/4} \\biggl({m_{*} \\over m_{*f} }\\biggr)^{1/2} \\,\\, M_{\\odot} \\, yr^{-1} \\label{eq2} \\end{equation} where $\\epsilon_{core}$ is the fraction of the total core mass (M$_{core}$), $\\phi_P$ is the ratio of the core's surface pressure to the mean pressure in the clumps, $f_{gas}$ is the fraction of the cloud's mass that is in gas, as opposed to stars, $\\Sigma_{cl}$ is a mean mass column density, $m_{*}$ (=$\\epsilon_{core}$M$_{core}$, where M$_{core}$ is the total core mass) is the instantaneous mass (the current mass), and $m_{*f}$ is the final mass of the star. We used the values of $\\Sigma_{cl}$ =1, $\\epsilon_{core}$=0.5 and $\\phi_P$=0.663 (see \\cite{mck02} for details). The bolometric luminosity is the sum of the internal and accretion luminosity (L$_{bol}$ = L$_{int}$ + L$_{acc}$). The accretion luminosity is given by \\cite{mck03}: \\begin{equation} L_{acc} \\sim 3.0\\times10^4 \\,\\, \\biggl({f_{acc} \\over 0.5}\\biggr) \\biggl({m_{*f} \\over {30 M_\\odot}}\\biggr)^{1.2} \\Sigma_{cl}^{3/4} \\biggl({m_{*} \\over m_{*f} }\\biggr)^{0.95} \\,\\, \\Sigma_{cl}^{3/4} \\,\\, L_{\\odot} \\end{equation} where ${f_{acc}}$ ($<<$1) is a factor that accounts for the energy advected into the star or used to drive protostellar outflows. The internal luminosity, L$_{int}$, is equal to the luminosity transported by radiation and is determined mostly by the stellar mass (L$_{int}$ $\\propto$ M$_*^{5.5}$ R$_*^{-0.5}$ $\\propto$ M$_{*}$). A simple formula cannot be used for a wide range of masses of stars, because as the mass increases, the relative contribution of the Kramers and electron scattering opacities changes \\citep{nak00}. Therefore, in order to estimate the mass of star from the bolometric luminosity, we directly used Figure 2 of \\citet{mck02}. % and Figure 6 of \\citet{mck03}. The instantaneous masses of TC3A and TC4A are 3-8 M$_{\\odot}$ and 2-5 M$_{\\odot}$ for the bolometric luminosities of $\\sim$ 1700 and 1200 L$_{\\odot}$, respectively, assuming a significant amount of the 1.3mm flux is from Component A. The accreting properties of the TC3 core are similar to those of G34.24+0.1 \\citep{mck03}. We estimated accretion rates using the equation 2 for a grid of final masses of 7.5, 30 and 120 M$_\\odot$. The accretion rates for TC3 and TC4 are 1-5$\\times$10$^{-4}$ M$_{\\odot}$ yr$^{-1}$ and 0.9-4$\\times$10$^{-4}$ M$_{\\odot}$ yr$^{-1}$, respectively. We also estimated the current masses and the accretion rates of TC1 and TC2 using L$_{bol}$ of 600 L$_{\\odot}$; the current masses are 1-3 M$_{\\odot}$ and the accretion rates are 0.8-1.7$\\times$10$^{-4}$ M$_{\\odot}$ yr$^{-1}$. The estimated accretion luminosities infer that high portions of the bolometric luminosities are from accretion luminosities in the early stage of massive-star formation. The timescale of the infalling stage is determined mainly by the conditions in the stars' natal cloud and weakly depends on the mass of stars. In the study of massive-star formation, a fundamental open question is how clusters are formed. Do massive dense cores have internal substructures? Do clumps evolve independently to produce stars, or do they share a common evolutionary process? Are apparent-single stars born single, or are they born in groups and subsequently ejected? While low-mass stars are believed to be produced mainly through accretion (Lada 1991), there are two main scenarios to explain high-mass star formation. One scenario is through accretion, like that of low-mass stars but with higher accretion rates; the other scenario is through formation of high-mass stars through coalescence of lower-mass protostars. We found 5 protostars for each of TC3 and TC4 cores, and the brightest and the most massive star in each cluster is located in the dust continuum peak (see Table 2), implying these systems are possible protoclusters. The brightest protostar appears near the center of the dust continuum peak. The brightest 24$\\mu$m source in the TC3, TC3A, is located in the millimeter peak and at the center of the core. Within the cluster, as the separation of TC3B-TC3C from the largest star of each Component A increases, the mid-infrared luminosity decreases in the case of the TC3 core, as shown in Table 2. This suggests that the brightest and most massive star among 4 protostars is located at the bottom of the gravitational potential well and is located where it formed, although the mid-infrared luminosities may not be directly correlated with the total bolometric luminosities. Accretion from gas or nearby fragmented clumps is found through the accretion of residual gas onto relatively low-mass cores. This picture may be consistent with the mass segregation, in young stellar clusters, and competitive accretion proposed by Bonnell \\& Davies (1998) and Bonnell et al. (2004). However, the brightest 24$\\mu$m source in the TC4 core, TC4A, is slightly off from the center of the millimeter peak, although TC4A is still the largest star in the cluster and the closest star to the dust continuum peak. %As the separation %of TC4B from the largest star, Component A, increases, The mid-infrared luminosity of TC4B is also smaller than that of the TC4A. %the star does not directly decrease in the case of the core of %TC4. It is unknown how these mid-infrared luminosities correlate with the total bolometric luminosities for individual sources. Comparable resolution far-infrared or millimeter observations are required in order to answer this question. Unfortunately, even through the Trifid Nebula is one of the nearest massive-star-forming regions with a distance of only 1.68 kpc, protostars in the Trifid as well as other comparable massive-star-forming regions are not bright enough for such observatories as the submillimeter Array (SMA). Future far-infrared observations with comparable spatial resolution are needed to answer this question." }, "0601/astro-ph0601205_arXiv.txt": { "abstract": "{We present optical and near-infrared observations with Keck of the binary millisecond pulsar PSR~J0751+1807. We detect a faint, red object -- with $R=25.08\\pm0.07$, $B-R=2.5\\pm0.3$, and $R-I=0.90\\pm0.10$ -- at the celestial position of the pulsar and argue that it is the white dwarf companion of the pulsar. The colours are the reddest among all known white dwarfs, and indicate a very low temperature, $T_\\mathrm{eff}\\approx4000$\\,K. This implies that the white dwarf cannot have the relatively thick hydrogen envelope that is expected on evolutionary grounds. Our observations pose two puzzles. First, while the atmosphere was expected to be pure hydrogen, the colours are inconsistent with this composition. Second, given the low temperature, irradiation by the pulsar should be important, but we see no evidence for it. We discuss possible solutions to these puzzles. ", "introduction": "\\label{sec:intro} Among the pulsars in binaries, the largest group, the low-mass binary pulsars, has low-mass white-dwarf companions. Before the companions became white dwarfs, their progenitors filled their Roche lobe and mass was transferred to the neutron stars, thereby spinning them up and decreasing their magnetic fields. Considerations of the end of this stage, where the white dwarf progenitor's envelope becomes too tenuous to be supported further, allow one to make predictions for relations between the orbital period and white dwarf mass, and orbital period and eccentricity (for a review, e.g., \\citealt{pk94,sta04}). Furthermore, after the cessation of mass transfer, two clocks will start ticking at the same time: the neutron star, now visible as a millisecond pulsar, will spin down, while the secondary will contract to a white dwarf and start to cool. Consequently, the spin-down age of the pulsar should equal the cooling age of the white dwarf. From optical observations of white-dwarf companions to millisecond pulsars one can estimate the white-dwarf cooling age and compare it with the pulsar spin-down age. Initial attempts to do this \\citep{hp98a,hp98b,sdb00} revealed a dichotomy in the cooling properties of white dwarfs in the sense that some white dwarf companions to older pulsars have cooled less than those of younger pulsars. In particular, the companions of PSR~J0437$-$4715 \\citep{dbv93,sbb+01} and PSR~B1855+09 \\citep{kbkk00,rt91} have temperatures of about 4000--5000\\,K, with characteristic pulsar ages of 5\\,Gyr. This is in contrast to the companion of PSR~J1012+5307 \\citep{llfn95,kbk96,cgk98}, which has a higher temperature (8600\\,K), while it orbits an older pulsar (8.9\\,Gyr). A likely cause for this dichotomy is the difference in the thickness of the envelope of hydrogen surrounding the helium core of the white dwarf \\citep{ashp96}. After the cessation of mass transfer, the white dwarfs have relatively thick ($\\sim\\!10^{-2}$\\,M$_\\odot$) hydrogen envelopes which are able to sustain residual hydrogen shell-burning, keeping the white dwarf hot and thereby slowing the cooling \\citep{dsbh98}. The shell burning, however, can become unstable and lead to thermal flashes which can reduce the mass of the envelope. White dwarfs with such reduced, relatively thin ($\\la10^{-3}$\\,M$_\\odot$) hydrogen envelopes cannot burn hydrogen and, as a result, cool faster. The transition between thick and thin hydrogen envelopes was predicted to lie near 0.18--0.20\\,M$_\\odot$ (where heavier white dwarfs have thin envelopes; \\citealt{ashp96,seg00,asb01}). Until recently, PSR~J1012+5307, with an orbital period $P_\\mathrm{b}=0.60$\\,d, was the only system for which a thick hydrogen envelope was required to match the two timescales. Given the relation between the white dwarf mass and the orbital period \\citep{jrl87,rpj+95,ts99}, companions in similar or closer orbits should have similar or lower mass, and thus have thick hydrogen envelopes as well. This was confirmed by the recent discovery of two new, nearby, binary millisecond pulsars with orbital periods similar to that of PSR~J1012+5307; PSR~J1909$-$3744 (1.53\\,d, \\citealt{jhb+05}) and PSR~J1738+0333 (0.354\\,d, Jacoby et al., in prep.; see \\citealt{kbjj05} for preliminary results). For both, the temperatures and characteristic ages are similar to those of PSR J1012+5307, and thus one is led to the same need for a thick hydrogen envelope. These discoveries, combined with the thin envelopes inferred for PSR~J0034$-$0534 (1.59\\,d) and binaries with longer periods, suggest that the transition occurs at a mass that corresponds to an orbital period just over 1.5\\,d (\\citealt{kbjj05}). All systems with shorter orbital periods should have thick hydrogen envelopes. The two known millisecond pulsars with white dwarf companions that have shorter orbital periods than PSR~J1012+5307 but do not have optical counterparts, are PSR~J0613$-$0200, with a 1.20\\,d period, and PSR~J0751+1807, which has the shortest orbital period of all binary millisecond pulsars with $M_\\mathrm{c}>0.1$\\,M$_\\odot$ companions, 0.26\\,d \\citep{lzc95}. The latter system is of particular interest because the companion mass has been determined from pulse timing ($M_\\mathrm{WD}=0.19\\pm0.03$\\,M$_\\odot$ at 95\\% confidence; \\citealt{nss+05}), so that one does not have to rely on the theoretical period-mass relationship. Intriguingly, for PSR J0751+1807, optical observations from \\citet{lcf+96} set a limit to the temperature of 9000\\,K, which is only marginally consistent with it having a thick hydrogen envelope. Based on this, \\citet{esa01}, suggested the hydrogen envelope may have been partially lost due to irradiation by the pulsar. The faintness of the companion to PSR~J0751+1807 aroused our curiosity and motivated us to obtain deep observations to test the theoretical ideas discussed above. We describe our observations in Sect.~\\ref{sec:observations}, and use these to determine the temperature, radius and cooling history in Sect.~\\ref{sec:tandr}. In Sect.~\\ref{sec:irradiation}, we investigate irradiation by the pulsar, finding a surprising lack of evidence for any heating. We discuss our results in Sect.~\\ref{sec:discussion}. ", "conclusions": "\\label{sec:discussion} We have optically identified the white dwarf companion of the binary millisecond pulsar PSR~J0751+1807. We find that the companion has the reddest colours of all known millisecond pulsar companions and white dwarfs. These colours indicate that the companion has a very low (ultra-cool) temperature of $T_\\mathrm{eff}\\sim\\!3500-4300$\\,K. Furthermore, the colours suggest that the white dwarf has a pure helium atmosphere, or a helium atmosphere with some hydrogen mixed in, as invoked for the field white dwarf WD~0346+246 which has similar colours \\citep{osh+01,ber01}. Our observations are inconsistent with evolutionary models, from which one would expect a pure hydrogen atmosphere. Indeed, as for other short-period systems, the hydrogen envelope is expected to be thick enough to sustain significant residual hydrogen burning, leading to temperatures far in excess of those observed. It may be that the mass of the envelope was reduced due to shell flashes or irradiation by the pulsar, as was proposed by \\citet{esa01}. However, we see no evidence for irradiation, despite the fact that the pulsar spin-down flux inpinging on the white dwarf is roughly double the observed thermal flux. Clues to what happens might be found from more detailed studies of the spectral energy distribution, or more accurate phase-resolved photometry. Finally, a deeper observation at infrared wavelengths would allow one to distinguish between the different atmosphere compositions for the companion: for a pure helium atmosphere, black-body like colours are expected, while if any hydrogen is present, the infrared flux would be strongly depressed (as is seen for WD 0346+246). With adaptive optics instruments, such observations should be feasible." }, "0601/astro-ph0601496_arXiv.txt": { "abstract": "{ Some common properties of helical magnetic fields in decaying and driven turbulence are discussed. These include mainly the inverse cascade that produces fields on progressively larger scales. Magnetic helicity also restricts the evolution of the large-scale field: the field decays less rapidly than a non-helical field, but it also saturates more slowly, i.e.\\ on a resistive time scale if there are no magnetic helicity fluxes. The former effect is utilized in primordial field scenarios, while the latter is important for successfully explaining astrophysical dynamos that saturate faster than resistively. Dynamo action is argued to be important not only in the galactic dynamo, but also in accretion discs in active galactic nuclei and around protostars, both of which contribute to producing a strong enough seed magnetic field. Although primordial magnetic fields may be too weak to compete with these astrophysical mechanisms, such fields could perhaps still be important in producing polarization effects in the cosmic background radiation. ", "introduction": "Magnetic helicity plays a fundamental role both in primordial theories of galactic magnetism as well as in dynamo theories amplifying and sustaining contemporary galactic fields. Both issues have been reviewed in recent years (Grasso \\& Rubinstein 2001; Widrow 2002; Giovannini 2004; Brandenburg \\& Subramanian 2005a). We will therefore only try to collect the main points relevant to the issues concerning magnetic helicity in galactic and protogalactic magnetism. The main reason magnetic helicity is at all of concern to us is that even in the resistive case the rate of magnetic helicity dissipation asymptotes to zero as the magnetic Reynolds number goes to infinity. This is not the case with magnetic energy dissipation, which remains always important, and does not decrease with increasing magnetic Reynolds number (Galsgaard \\& Nordlund 1996). Therefore the magnetic helicity is nearly conserved at all times. This has serious consequences for the evolution of magnetohydrodynamic (MHD) turbulence, as has been demonstrated by a number of recent simulations when the resolution has been large enough (Brandenburg 2001a; Mininni et al.\\ 2005). At a more descriptive level, magnetic helicity characterizes the degree of field line linkage. As the magnetic field relaxes, its energy decreases, but the linkage stays, at least as much as possible. The field's inability to undo its knots implies also that the field cannot decay freely. This slows down the decay, which is important if a primordial field is to be of any significance at the time of recombination. In the driven case, on the other hand, magnetic helicity is better pictured in terms of writhe and twist (e.g.\\ Longcope \\& Klapper 1997; D\\'emoulin et al.\\ 2002). Writhe refers to the tilt of a flux tube, and we use both expressions synonymously. A cyclonic event tilts individual flux tubes, but as it does so, a corresponding amount of internal twist is necessarily introduced in the tube (Blackman \\& Brandenburg 2003). This is what saturates the dynamo, and this can be a very powerful effect if the small-scale internal twist cannot escape. In this review we discuss both decaying and driven turbulence. The former is relevant for prolonging the decay of a primordial field, while the latter is relevant for understanding how the galactic dynamo saturates and how to enable it to do so faster. ", "conclusions": "Not all magnetic fields will be helical, but if they are, this can have dramatic consequences for their evolution. The effects can be equally dramatic both in decaying and in driven turbulence, as has been highlighted in this review. Although we have not discussed this in the present paper, it should be emphasized that helical large-scale magnetic fields can also be generated in non-stratified shear flows where there is no $\\alpha$ effect, but there can instead be the so-called shear--current of $\\meanWW\\times\\meanJJ$ effect (Rogachevskii \\& Kleeorin 2003, 2004). This effect may also explain the large-scale dynamo action seen in \\Fig{pmean_comp}, where the results without helicity are quite similar to those with helicity (Brandenburg 2005). One-dimensional mean field calculations with the $\\meanWW\\times\\meanJJ$ effect (Brandenburg \\& Subramanian 2005c) show that in this case a magnetic $\\alpha$ effect can be produced that has different signs on the two sides of the midplane. This magnetic $\\alpha$ effect thus contributes to the saturation of the dynamo even if there is no ordinary (kinetic) $\\alpha$ effect. This highlights once more the dramatic effects played by magnetic helicity. Whether or not the primordial magnetic field was really helical remains a big question. If it was, it is likely that an inverse cascade process has produced fields of progressively larger scale. This might lead to observable effects in the cosmic microwave background. Such a field may also be important for seeding the galactic dynamo, but it is important to realize that a variety of astrophysical mechanisms may also produce seed fields just as large. Our estimate for magnetized outflows from AGNs or YSOs assumes that the source remains active for a certain period of time, and that their exhaust goes freely into the ambient medium. Partial evidence for this actually happening lies in the fact that clusters of galaxies are chemically enriched with heavier elements. Given that magnetic fields are intrinsically connected with the outflow, just like the heavier elements in it, it is quite plausible that some degree of magnetic contamination of the cluster must have occurred. In order to produce finally the observed large-scale magnetic fields of galaxies, some more reshaping, amplification, and maintenance against magnetic decay is necessary. Roughly, we expect this to happen just like the mean field dynamo is able to amplify and maintain the field, although it must operate on an already strong enough field. This initial field will still be random and of mixed parity about the midplane (or equator), but there will be some finite degree of quadrupolar field which is the one that is dominant in many galaxies; see Krause \\& Beck (1998) and Brandenburg \\& Urpin (1998) for a related discussion about the importance of seeding the quadrupolar field component. As we have argued above, the catastrophic quenching problem of the dynamo has to be overcome, and this is likely to be the case because of various magnetic and current helicity fluxes operating within the entire dynamo domain. In the context of the solar dynamo, simulations have now begun to demonstrate the dramatic difference made by open boundary conditions, and we hope that a similar demonstration will soon be possible for the galactic dynamo as well. Corresponding mean field calculations have already been performed showing that the catastrophic quenching effect is overcome by an advective flux out of the domain along the vertical direction. In particular, it will be interesting to see whether the shedding of magnetic helicity can actually lead to directly observable effects." }, "0601/astro-ph0601175_arXiv.txt": { "abstract": "{We present results from Chandra observations of the 3C/FR~I sample of low luminosity radio-galaxies. We detected a power-law nuclear component in 12 objects out of the 18 with available data. In 4 galaxies we detected nuclear X-ray absorption at a level of $N_H \\sim (0.2-6) \\times 10^{22}$ cm$^{-2}$. X-ray absorbed sources are associated with the presence of highly inclined dusty disks (or dust filaments projected onto the nuclei) seen in the HST images. This suggests the existence of a flattened X-ray absorber, but of much lower optical depth than in classical obscuring tori. We thus have an un-obstructed view toward most FR~I nuclei while absorption plays only a marginal role in the remaining objects. Three pieces of evidence support an interpretation for a jet origin for the X-ray cores: i) the presence of strong correlations between the nuclear luminosities in the radio, optical and X-ray bands, extending over 4 orders of magnitude and with a much smaller dispersion ($\\sim$ 0.3 dex) when compared to similar trends found for other classes of AGNs, pointing to a common origin for the emission in the three bands; ii) the close similarity of the broad-band spectral indices with the sub-class of BL Lac objects sharing the same range of extended radio-luminosity, in accord with the FR~I/BL~Lacs unified model; iii) the presence of a common luminosity evolution of spectral indices in both FR~I and BL~Lacs. The low luminosities of the X-ray nuclei, regardless of their origin, strengthens the interpretation of low efficiency accretion in low luminosity radio-galaxies. ", "introduction": "By exploring the properties of low luminosity radio-galaxies (LLRG) it is possible to extend the study of active galactic nuclei (AGN), and in particular of the radio-loud sub-population, toward the lowest level of nuclear luminosity. This offers the opportunity to improve our understanding of the mechanism of accretion onto super-massive black holes and of their radiative manifestations. In this respect, it is important to note that LLRG not only represent the bulk of radio-loud active galaxies, due to the steepness of their luminosity function, but also a substantial fraction of the overall galaxies population. In fact LLRG are associated to a fraction as high as 40 \\% of all bright elliptical and lenticular galaxies \\citep{sadler89,auriemma77}. Thus the presence of a low power radio-galaxy represents by far the most common manifestation of nuclear activity in early type galaxies. Low luminosity AGN (LLAGN) also represent a link between the high luminosity AGN and the population of quiescent galaxies, as it is now widely recognized that most (if not all) galaxies host a super-massive black-hole \\citep[e.g.][]{kormendy95}. The comparison of the different manifestations of nuclear activity across the largest possible range of luminosity is then a crucial step to unveil the connections (and the diversities) between active and non-active galaxies. Unfortunately, the study of LLRG has been significantly hampered by the contamination from host galaxy emission which, in most observing bands, dominates the emission from the AGN. To constrain the physical processes at work in these objects it is clearly necessary to disentangle the AGN and host's contributions via spectral decomposition or high resolution imaging. In the last few years, thanks in particular to the Hubble Space Telescope (HST) this has become routinely possible. An example of the insights that can be obtained following this approach comes from the results derived from the analysis of broad-band HST imaging of a sample of LLRG \\citep{chiaberge:ccc}. In the majority of the targets, HST images revealed the presence of unresolved optical nuclei. Fluxes and luminosities of these sources show a tight correlation with the radio cores, extending over four orders of magnitude. This has been interpreted as being due to a common non-thermal emission process in the radio and optical band, i.e. that we are seeing the optical emission from the base of a relativistic jet. The low luminosity of LLRG optical nuclei, and the possible dominance of the emission related to out-flowing material, indicates a low level of accretion and/or a low radiative efficiency with respect to classical, more luminous AGN. The high fraction ($\\sim$ 85 \\%) of objects with detected optical nuclear sources suggests a general lack of obscuring molecular tori, a further distinction with respect to other classes of AGN. The tenuous torus structure, when considered together with the low accretion rate, the low mass of the compact emission line regions (10 - 10$^3 M_{\\sun}$) and the limits to the mass of the Broad Line Region ($M_{BLR} < 10^{-2} M_{\\sun}$), indicates that a general paucity of gas in the innermost regions of LLRG emerges as the main characterizing difference from more powerful AGN \\citep{capetti:cccriga}. The advent of Chandra provides us with the unique opportunity to extend the study of LLRG to the X-ray band since its high spatial resolution enables us to isolate any low power nuclear source associated to LLRG. Furthermore, with respect to the optical data, the spectral capabilities of Chandra allow us to study directly the spectral behaviour of the X-ray nuclei. This can be used to quantify the effects of local absorption as well as the slope of their high energy emission as to better constrain the emission processes at work in these sources, particularly when combined in a multi-wavelength analysis using also optical and radio observations. This approach was already followed by several authors in the past \\citep[e.g.][]{fr1sed,hardcastle00,trussoni03,hardcastle03} but clearly, the superior capabilities of Chandra warrant to re-explore this issue in much greater depth. Here we present results obtained from the analysis of Chandra data for the sample of LLRG drawn from the 3C catalogue of radio-sources \\citep{bennett62}, more specifically those with a morphological classification as FR~I \\citep{fanaroff74}. Considering LLRG from the 3C sample, the same studied by \\citet{chiaberge:ccc} represents an obvious choice. It is the best studied sample of radio-loud galaxies in existence with a vast suite of ground and spaced based observations for comparison at essentially all accessible wavelengths. Other samples of LLRG have been considered in the literature, but, not surprisingly, the coverage provided by Chandra observations for these samples is severely incomplete. For example, only about 10 \\% of B2 sample of LLRG \\citep{colla75} has been observed with Chandra (to be compared to the $\\sim$ 55 \\% for the 3C/FR~I sample), and the analysis is further obstructed by the unknown selection biases introduced in the choice of the individual targets. Similarly we preferred to focus on the Chandra data alone, to provide the highest possible level of uniformity in the analysis. The paper is organized as follows: in Sec. \\ref{obs} we present the observations and the data reduction that lead to the results described in Sec. \\ref{results} and in particular to the detection of X-ray nuclei in most FR~I. In Sec. \\ref{absorption} we explore the effects of the nuclear X-ray absorption, by also relating it to the presence of dust features seen in the HST images. The origin of the X-ray nuclei found in LLRG is explored in Sec. \\ref{jet}. The results are discussed and summarized in Sec. \\ref{summary}. We adopted $\\rm{H}_o=75 $ km s$^{-1} $Mpc$^{-1}$ and $q_0=0.5$. ", "conclusions": "" }, "0601/astro-ph0601669_arXiv.txt": { "abstract": " ", "introduction": "The structure and evolution of dark matter halos is widely discussed in the recent years (e.g. \\cite{BM}) A large number of papers has been devoted to the investigation of spatial density profiles of dark matter halos, $\\rho(r)$. both numerical \\cite{Navarro, Moore05} and analytical \\cite{LMKW}. Recently several papers appeared that described general properties of solutions of Jeans equation \\cite{DM,AWBBD}. The Jeans equation for a sperically symmetric self-gravitating system reads as follows: \\begin{equation} \\label{Jeans} \\frac{d}{dr} (\\rho\\,{\\sigma_r}^2) + \\frac{2\\beta}{r} (\\rho\\,{\\sigma_r}^2) + \\rho\\frac{G\\,M(r)}{r^2} = 0 \\,. \\end{equation} Here ${\\sigma_r}^2$ and ${\\sigma_t}^2$ are the radial and tangential velocity dispersion, $\\beta = 1-\\frac{{\\sigma_t}^2}{2\\,{\\sigma_r}^2}$ is the Binney's anisotropy parameter (systems with $0<\\beta<1$ are radially anisotropic, with $\\beta<0$ -- tangentially anisotropic, and $\\beta=0$ is the isotropic case). An additional assumption is the following empirical property of dark matter halos: the generalized phase-space-{\\it like} density of dark matter particles in a halo is a power-law in radius: \\begin{equation} \\label{gpsd} \\rho(r)/{\\sigma_r}^\\epsilon(r) \\propto r^{-\\alpha} \\,, \\end{equation} which is established both in N-body calculations \\cite{TN} and some analytical models \\cite{AWBBD}. For a closed set of equations one also needs to specify $\\beta$. Here two different approaches may be adopted: in \\cite{AWBBD} isotropy was assumed ($\\beta=0$), and in \\cite{DM} the following relationship between $\\beta$ and the logarithmic slope of density profile $\\gamma = d \\ln\\rho(r)/d r$ was taken: \\begin{equation} \\label{betagammarelation} \\beta = \\beta_0 + b(\\gamma-\\gamma_0) \\,. \\end{equation} This kind of relation has been proposed in \\cite{HM} based on empirical analysis of different numerically treated situations. These investigations demonstrated that under these assumptions there exist only few possibilities for a self-consistent halo model. The most realistic is the set of models with density profiles having inner and outer power-law asymptotes (so-called Zhao $\\alpha\\beta\\gamma$ models \\cite{Zhao}). However, this approach, being very fruitful, does not describe explicitly the distribution function. It operates only with its second moment with respect to velocity, i.e., the velocity dispersion. The most complete description of halo structure can be made in terms of phase-space distribution function $f(r,v)$. One attempt to deal with full velocity distribution function (VDF) was made in \\cite{Tsallis}, where an Eddington inversion formula was applied to obtain VDF for power-law density profiles. However, this study was limited to isotropic case. An approach based on certain physical assumptions regarding the form of the distribution function seems more promising. It avoids postulating empirical assumptions like (\\ref{gpsd}, \\ref{betagammarelation}), which, nevertheless, still hold in many cases. It is convenient to express the distribution function in terms of adiabatic invariants during slow evolution of the gravitational potential. This allows to describe easily the process of adiabatic contraction of dark matter halo caused by condensation of baryonic matter in its centre, which occurs during the formation of a galaxy. This approach is adopted in the present paper. The paper is organized as follows. In the second section we describe two models of inner halo structure and explain physical arguments for the underlying assumptions concerning the dark matter distribution function. In the third section we calculate the velocity anisotropy coefficient. Furthermore, in the fourth section we consider the halo response to the adiabatic compression and compare it with other studies. Finally, the conclusions are presented. ", "conclusions": " \\begin{enumerate} \\item The more radially biased is the velocity, the less compressed is the halo. \\item The shallower is initial density profile (less value of $\\gamma$), the less is the degree of compression compared to that of Blumenthal's method. Standard model of adiabatic contraction systematically overestimates the effect of contraction. \\item The model predicts a moderate enhancement (2 to 4 times) of dark matter mass in the bulge. The shallower is the initial profile, the greater the increase of dark matter mass. This result significantly increases the estimates of possible annihilation radiation flux from the center of the Galaxy \\cite{GS, Merritt, ZV}. \\end{enumerate} I am grateful to Maxim Zelnikov for helpful discussion on the topic of the paper, to Steen Hansen for pointing out some interesting issues and to anonymous referee for valuable comments and proposals. This work was supported by Landau Foundation (For\\-schungs\\-zent\\-rum J\\\"ulich) and Russian Fund for Basic Research (project nos. 01-02-17829, 03-02-06745)." }, "0601/astro-ph0601343_arXiv.txt": { "abstract": "{We solve a weakly singular integral equation by Laplace transformation over a finite interval of $\\mathbb{R}$. The equation is transformed into a Cauchy integral equation, whose resolution amounts to solving two Fredholm integral equations of the second kind with regular kernels. This classical scheme is used to clarify the emergence of the auxiliary functions expressing the solution of the problem. There are four such functions, two of them being classical ones. This problem is encountered while studying the propagation of light in strongly scattering media such as stellar atmospheres.\\\\ Key words: Weakly singular integral equation, finite Laplace transform, sectionally analytic function, radiation transfer theory, stellar atmospheres. ", "introduction": "The integral form of the equation describing the radiative transfer of energy in a static, plane-parallel stellar atmosphere is \\cite{ahuesetal2002a} \\begin{equation} \\label{eq1} S(a,b,\\tau)=S_0(a,b,\\tau)+a\\int_{0}^{b}K(\\tau-\\tau')S(a,b,\\tau')d \\tau', \\end{equation} where $S$ is the source function of the radiation field and $S_0$ describes the radiation of the primary (internal or external) sources. These functions depend on the two parameters of the problem: the albedo $a\\in]0,1[$, which characterizes the scattering properties of the stellar plasma, and the optical thickness $b>0$ of the atmosphere. They also depend on the optical depth $\\tau\\in[0,b]$, which is the space variable. Equation (\\ref{eq1}) means that the radiation field at level $\\tau$ is the sum of the {\\it{direct}} field from the primary sources, and the {\\it{diffuse}} field having scattered at least once.\\\\ In the simplest scattering process conceivable - i.e., a monochromatic and isotropic one - the kernel of the integral equation (\\ref{eq1}) is the function \\begin{equation} \\label{eq2} K(\\tau):\\,=\\frac{1}{2}E_1(\\vert \\tau \\vert)\\quad(\\tau\\in \\mathbb{R}^*), \\end{equation} where $E_1$ is the first exponential integral function \\begin{equation} \\label{eq3} E_1(\\tau):\\,=\\int_{0}^{1}\\exp(-\\tau/x)\\frac{dx}{x}\\quad(\\tau > 0). \\end{equation} Since $E_1(\\vert \\tau \\vert) \\sim - \\ln(\\vert \\tau \\vert)$ when $\\vert \\tau \\vert \\to 0^+$, the kernel of the integral equation (\\ref{eq1}) is weakly singular on its diagonal. The free term $S_0$ includes the thermal emission of the stellar plasma - of the form $(1-a)B^*(\\tau)$, where $B^*$ is a known function - and the contribution $aJ_0^{ext}(b,\\tau)$ of the external sources via the boundary conditions: see the introduction of \\cite{ahuesetal2002a}. Hence $S_0(a,b,\\tau)=(1-a)B^*(\\tau)+aJ_0^{ext}(b,\\tau)$. In a homogeneous and isothermal atmosphere assumed to be in local thermodynamic equilibrium, the function $B^*$ coincides with the Planck function $B(T)$ at the (constant) temperature $T$. Moreover, $J_0^{ext}=0$ in the absence of external sources and thus \\begin{equation} \\label{eq4} S_0(a,b,\\tau)=(1-a)B(T), \\end{equation} which shows that $S_0$ is independent of $\\tau$ in this model. The solution $S$ to the problem (\\ref{eq1}) is then \\begin{equation} \\label{eq5} S(a,b,\\tau)=(1-a)B(T)Q(a,b,\\tau), \\end{equation} where $Q$ solves the following integral equation: \\begin{equation} \\label{eq6} Q(a,b,\\tau)=1+a\\int_{0}^{b}K(\\tau-\\tau')Q(a,b,\\tau')d \\tau'. \\end{equation} It is proved in \\cite{ahuesetal2002b} that the space $C^0([0,b])$ of the continuous functions from $[0,b]$ to $\\mathbb{R}$ is invariant under the operator \\begin{equation} \\label{eq7} \\Lambda\\;:\\;f\\;\\;\\rightarrow\\;\\;\\Lambda f(\\tau):\\,= \\int_{0}^{b}K(\\tau-\\tau')f(\\tau')d\\tau', \\end{equation} with norm \\begin{equation} \\label{eq8} \\|\\Lambda\\|_{\\infty}= \\int_{0}^{b/2}E_1(\\tau)d\\tau=1-E_2(b/2). \\end{equation} Here, $E_2(\\tau):=\\int_{0}^{1}\\exp(-\\tau/x)dx$ is the exponential integral function of order 2. Equation (\\ref{eq6}), which can be written in the form \\begin{equation} \\label{eq9} Q=1+a \\Lambda Q, \\end{equation} has therefore a unique solution in $C^0([0,b])$ provided that $a<1$ or $b<+\\infty$.\\\\ This problem is a basic one in stellar atmospheres theory, and more generally in transport theory. It describes in the simplest way the multiple scattering of some type of particles (here photons) on scattering centers distributed uniformly in a slab of finite thickness, a very simple 1D-configuration. Its applications in astrophysics and neutronics - among other fields - are presented in \\cite{rutily2002}. It is important to solve this problem very accurately, thus providing a benchmark to validate the numerical solutions of integral equations of the form (\\ref{eq1}).\\\\ Physicists and astrophysicists have developed many methods for solving integral equations of the form (\\ref{eq1}) with a convolution kernel defined by (\\ref{eq2}) \\cite{rutily-bergeat1994}. The main steps for solving the prototype equation (\\ref{eq6}) are summarized in a recent article \\cite{chevallier-rutily2004}, which contains accurate tables of the function $(1-a)Q$ for different values of the parameters $a$ and $b$. While reading this paper, one is struck by the complexity of the ``classical'' solution to Eq. (\\ref{eq6}). It requires the introduction of many intricate auxiliary functions introduced in the literature over more than thirty years. The reader quickly loses the thread of the solution, which reduces his chances of exploiting it for solving problems of the more general form (\\ref{eq1}).\\\\ The aim of the present article is to get around this difficulty by solving Eq. (\\ref{eq6}) straightforwardly, introducing as few auxiliary functions as possible to express its solution. These functions are briefly studied in Appendix A. The method is based on the finite Laplace transform, which reduces the problem (\\ref{eq6}) to solving a Cauchy integral equation over $[-1, +1]$, which in turn can be transformed into two Fredholm integral equations over [0, 1]. This approach has been developed in transport theory after the publication in 1953 of the first English translation of Muskhelishvili's monograph {\\it Singular integral equations} \\cite{muskhelishvili1992}: see, e.g., \\cite{busbridge1955}, \\cite{mullikin1964} and \\cite{carlstedt-mullikin1966}. It can be considered as an extension of the Wiener-Hopf method \\cite{busbridge1960} for solving integral equations of the form (\\ref{eq1}) with $b<\\infty$. Both methods are characterized by an intensive use of the theory of (sectionally) analytic functions, which allows solution of Eq. (\\ref{eq1}) in a concise manner. This is obvious when comparing the solution we derive here to the classical solution of the particular problem (\\ref{eq6}), which does not use any calculation in the complex plane. The former method clarifies the origin and the role of the auxiliary functions expressing the solution to a problem of the form (\\ref{eq1}). Since these functions are independent of the source term $S_0$, they are ``universal'' for a given scattering kernel.\\\\ The remainder of this article is organized as follows: in Sec. 2, the finite Laplace transform of the $Q$-function is calculated on the basis of some recent developments on Cauchy integral equations \\cite{rutily-bergeat2002}-\\cite{rutily-chevallier-bergeat2004}. Then the Laplace transform is inverted and the solution achieved in \\cite{chevallier-rutily2004} is concisely retrieved with the help of the theorem of residues (Sec. 3). It involves two functions $F_+$ and $F_-$ with remarkable properties, as shown in Appendix B. The difficulties arising from the numerical evaluation of the latter functions are investigated in \\cite{chevallier-rutily2004}. ", "conclusions": "Using the finite Laplace transform, we derived in a concise way the expression (\\ref{eq35}) of the solution to Eq. (\\ref{eq6}), which is appropriate for numerical evaluation \\cite{chevallier-rutily2004}. Our main objective was to come directly to this expression, with emphasis on the manner in which the four auxiliary functions of the problem were generated. They are ({\\it{i}}) the function $T=T(a,z)$, which characterizes the solution to Eq. (\\ref{eq6}) in an infinite medium (the range $[0,b]$ is replaced by $\\mathbb{R}$), ({\\it{ii}}) the function $H=H(a,z)$ which expresses its solution in a half-space ($b=\\infty$), and ({\\it{iii}}) the functions $\\zeta_{\\pm}=\\zeta_{\\pm}(a,b,z)$ that complete the previous ones in the finite case ($b<\\infty$). The main properties of these functions are summarized in Appendix A. We note that the $T$-function can be expressed in terms of elementary transcendental functions [Eq. (\\ref{eqA3})], the $H$-function is defined {\\it{explicitly}} by an integral on $[0,1]$ [Eq. (\\ref{eqA10})], and the functions $\\zeta_{\\pm}$ are defined {\\it{implicitly}} as the solutions to Fredholm integral equations of the second kind [Eq. (\\ref{eqA13})]. It seems that the problem (\\ref{eq6}) has no exact closed-form solution for $b<\\infty$.\\\\ The approach of this article is appropriate for solving integral equations of the form (\\ref{eq1}), with a kernel defined by (\\ref{eq2}) and any free term. It also applies to convolution kernels defined by functions more general than (\\ref{eq2}), for example of the form \\begin{equation} \\label{eq37} K(\\tau):=\\int_{I}\\Psi(x)\\exp(-\\vert \\tau \\vert /x)dx, \\end{equation} where $I$ is an interval of $\\mathbb{R}$ and $\\Psi$ a function ensuring the existence of the integral. This class of kernels models scattering processes more complex than the one considered in this article, which corresponds to $I=\\,]0,1]$ and $\\Psi(x)=(1/2)(1/x)$. \\appendix \\renewcommand{\\theequation}{\\Alph{section}\\arabic{equation}} \\setcounter{equation}{0} \\setcounter{section}{1}" }, "0601/astro-ph0601357_arXiv.txt": { "abstract": "The large-scale magnetic field of our Galaxy can be probed in three dimensions using Faraday rotation of pulsar signals. We report on the determination of 223 rotation measures from polarization observations of relatively distant southern pulsars made using the Parkes radio telescope. Combined with previously published observations these data give clear evidence for large-scale counterclockwise fields (viewed from the north Galactic pole) in the spiral arms interior to the Sun and weaker evidence for a counterclockwise field in the Perseus arm. However, in interarm regions, including the Solar neighbourhood, we present evidence that suggests that large-scale fields are clockwise. We propose that the large-scale Galactic magnetic field has a bisymmetric structure with reversals on the boundaries of the spiral arms. Streaming motions associated with spiral density waves can directly generate such a structure from an initial inwardly directed radial field. Large-scale fields increase toward the Galactic Center, with a mean value of about 2~$\\mu$G in the Solar neighbourhood and 4~$\\mu$G at a Galactocentric radius of 3 kpc. ", "introduction": "A diffuse magnetic field exists on all scales in our Galaxy. This field can be detected through observations of Zeeman splitting of spectral lines, of polarized thermal emission from dust at mm, sub-mm or infrared wavelengths, of optical starlight polarization due to anisotropic scattering by magnetically-aligned dust grains, of radio synchrotron emission, and of Faraday rotation of polarized radio sources \\citep[see][for a review]{hw02}. The first two approaches have been used to detect respectively the line-of-sight strength and the transverse orientation of magnetic fields in molecular clouds \\citep[e.g.][]{cru99,ncr+03,fram03}. Starlight polarization can be used to delineate the orientation of the transverse magnetic field in the interstellar medium within 2 or 3 kpc of the Sun. Careful analysis of such data show that the local field is mainly parallel to the Galactic plane and follows local spiral arms \\citep[e.g.][]{hei96}. Since we live near the edge of the Galactic disk, we cannot have a face-on view of the global magnetic field structure in our Galaxy through polarized synchrotron emission, as is possible for nearby spiral galaxies \\citep[e.g.][]{bbm+96}. Polarization observations of synchrotron continuum radiation from the Galactic disk \\citep[e.g.][]{rfr+02} give the transverse direction of the field in the emission region and some indication of its strength. Large-angular-scale features are seen emerging from the Galactic disk, for example, the North Polar Spur \\citep[e.g.][]{jfr87,dhjs97,drrf99,rfr+02}, and the vertical filaments near the Galactic center \\citep{hsg+92,dhr+98}. There are also many small-angular-scale structures resulting from diffuse polarized emission at different distances which are modified by foreground Faraday screens \\citep{gdm+01,ul02,hkd03a}. Faraday rotation of linearly polarized radiation from pulsars and extragalactic radio sources is a powerful probe of the diffuse magnetic field in the Galaxy \\citep[e.g.,][]{sk80,sf83,ls89,rk89,hq94,hmbb97,id98,fsss01}. Faraday rotation gives a measure of the line-of-sight component of the magnetic field. Extragalactic sources have the advantage of large numbers but pulsars have the advantage of being spread through the Galaxy at approximately known distances, allowing direct three-dimensional mapping of the field. Pulsars also give a direct estimate of the strength of the field through normalisation by the dispersion measure (DM). The rotation measure (RM) is defined by \\begin{equation} \\phi = {\\rm RM}\\; \\lambda^2 \\end{equation} where $\\phi$ is the position angle in radians of linearly polarised radiation relative to its infinite-frequency ($\\lambda = 0$) value and $\\lambda$ is its wavelength (in m). For a pulsar at distance $D$ (in pc), the RM (in rad~m$^{-2}$) is given by \\begin{equation} {\\rm RM} = 0.810 \\int_{0}^{D} n_e {\\bf B} \\cdot d{\\bf l}, \\end{equation} where $n_e$ is the electron density in cm$^{-3}$, ${\\bf B}$ is the vector magnetic field in $\\mu$G and $d {\\bf l}$ is an elemental vector along the line of sight toward us (positive RMs correspond to fields directed toward us) in pc. With \\begin{equation} {\\rm DM}=\\int_{0}^{D} n_e d l, \\end{equation} we obtain a direct estimate of the field strength weighted by the local free electron density \\begin{equation}\\label{eq_B} \\langle B_{||} \\rangle = \\frac{\\int_{0}^{D} n_e {\\bf B} \\cdot d{\\bf l} }{\\int_{0}^{D} n_e d l } = 1.232 \\; \\frac{\\rm RM}{\\rm DM}. \\label{eq-B} \\end{equation} where RM and DM are in their usual units of rad m$^{-2}$ and cm$^{-3}$ pc and $B_{||}$ is in $\\mu$G. \\citet{man72,man74} was first to systematically measure a number of pulsar RMs and to use them to investigate the large-scale Galactic magnetic field. He concluded that the local uniform field is directed toward Galactic longitude $l\\sim90\\degr$. \\citet{tn80} modeled the magnetic field configuration to fit the 48 pulsar RMs available that time, and they confirmed the local field direction and also found evidence for a field reversal near the Carina--Sagittarius arm. After the large pulsar RM dataset was published by \\citet{hl87}, \\citet{ls89} used 185 pulsar RMs to study the Galactic magnetic field and confirmed the field reversal found by \\citet{tn80}. \\citet{rl94} observed 27 RMs of distant pulsars in the first Galactic quadrant and provided evidence for a clockwise field (viewed from the north Galactic pole) near the Crux--Scutum arm. These field directions were recently re-examined by \\citet{wck+04} using revised pulsar distances from the NE2001 electron density model \\citep{cl02} together with their 17 new RMs, finding evidence for several field reversals both exterior and interior to the Solar circle. \\citet{hmq99} observed 54 RMs and tentatively identified a counterclockwise field near the Norma arm, which was later confirmed by \\citet{hmlq02} using the RM data discussed in this paper. More recently, \\citet{val05} has reanalysed the available pulsar RM data and interpreted it in terms of an overall clockwise field with a counterclockwise ring of width $\\sim 1$ kpc and radius $\\sim 5$ kpc centered on the Galactic Center. Pulsar RMs have also been used to study the small-scale random magnetic fields in the Galaxy. Some pairs of pulsars, which are close in sky position and have similar DMs, have very different RMs, indicating an irregular field structure on scales about 100~pc \\citep{ls89}. Some of these irregularities may result from HII regions in the line of sight to the pulsar \\citep{mwkj03}. \\citet{rk89} fitted the single-cell-size model for the residuals of pulsar RMs after the RM contribution of the proposed large-scale ring-field structure was subtracted and obtained a strength for the random field $B_r\\sim5\\mu$G. \\citet{os93} analyzed the difference of RMs and DMs of pulsar pairs and concluded that $B_r\\sim 4-6\\mu$G independent of cell-size in the range of 10 -- 100~pc. In fact the random fields exist on all scales. \\citet{hfm04} have found the power-law distribution for magnetic field fluctuations as $E_B(k)= C \\ (k / {\\rm kpc^{-1}})^{-0.37\\pm0.10}$ at scales from $1/k=$ 0.5~kpc to 15~kpc,with $C= (6.8\\pm0.8)\\ 10^{-13} {\\rm erg \\ cm^{-3} \\ kpc}$, corresponding to an rms field of $\\sim6\\mu$G in the scale range. The interstellar magnetic field is stronger at smaller scales and may be strongest at the scales of energy injection by supernova explosions and stellar winds (1 -- 10~pc). The Parkes multibeam survey has discovered a large number of low-latitude and relatively distant pulsars \\citep{mlc+01,mhl+02,kbm+03,hfs+04}, providing a unique opportunity to probe the diffuse magnetic field in a substantial fraction of the Galactic disk with much improved spatial resolution. In addition, improved estimates of pulsar distances are available from the NE2001 electron density model \\citep{cl02}. In this paper, we adopt a distance of the Sun from the Galactic center of $R_{\\odot}=8.5$~kpc. We have used the Parkes 64-m telescope of the Australia Telescope National Facility to observe the polarization properties of 270 pulsars, most of which were discovered in the Parkes Multibeam Pulsar Survey. After processing, we obtained 223 pulsar RMs which we present in \\S2. All available pulsar RM data have been used to reveal magnetic field directions along the spiral arms and in interarm regions, as presented in \\S3. The field strength and its Galactocentric radial dependence are analysed in \\S4. Our model for the large-scale Galactic field is discussed in \\S5 and concluding remarks are in \\S6. ", "conclusions": "Pulsars have unique advantages as probes of the large-scale Galactic magnetic field. Their distribution throughout the Galaxy at approximately known distances allows a true three-dimensional mapping of the large-scale field structure. Furthermore, combined with the measured DMs, pulsar RMs give us a direct measure of the mean line-of-sight field strength along the path, weighted by the local electron density. We have measured RMs for 223 pulsars, most of which lie in the fourth and first Galactic quadrants and are relatively distant. These new measurements enable us to investigate the structure of the Galactic magnetic field over a much larger region than was previously possible. Clear evidence is found for coherent large-scale fields aligned with the spiral-arm structure of the Galaxy. In all of the inner arms (Norma, Crux-Scutum, Carina-Sagittarius) there is strong evidence that the large-scale fields are counterclockwise when viewed from the north Galactic pole. Weaker evidence also suggests that a counterclockwise field is present in the Perseus arm. On the other hand, at least in the local region and in the inner Galaxy in the fourth quadrant, there is good evidence that the fields in interarm regions are similarly coherent, but clockwise in orientation. Evidence is also presented that large-scale magnetic fields are stronger in the inner part of the Galaxy and that fields in interarm regions are weaker than those in spiral arms. We therefore propose a bisymmetric model for the large-scale Galactic magnetic field with reversals on arm-interarm boundaries so that all arm fields are counterclockwise and all interarm fields are clockwise. This model for the Galactic magnetic field is appealing in its simplicity. It receives strong support from an objective analysis of mean line-of-sight fields near tangential points of an equiangular spiral. However, it clearly needs to be backed up by further observational work and by modelling of the effects of streaming motions on various initial field configurations. It is possible that the mode of initial field formation is not critical in determining the dominant present-day structure. Further rotation measure observations, especially for interarm regions and especially in the first Galactic quadrant, would be especially valuable." }, "0601/astro-ph0601161_arXiv.txt": { "abstract": "Resonant relaxation (RR) of orbital angular momenta occurs near massive black holes (MBHs) where the potential is spherical and stellar orbits are nearly Keplerian and so do not precess significantly. The resulting \\textit{coherent} torques efficiently change the magnitude of the angular momenta and rotate the orbital inclination in all directions. As a result, many of the tightly bound stars very near the MBH are rapidly destroyed by falling into the MBH on low-angular momentum orbits, while the orbits of the remaining stars are efficiently randomized. We solve numerically the Fokker-Planck equation in energy for the steady state distribution of a single mass population with a RR sink term. We find that the steady state current of stars, which sustains the accelerated drainage close to the MBH, can be $\\lesssim\\!10$ larger than that due to \\textit{non-coherent} 2-body relaxation alone. RR mostly affects tightly bound stars, and so it increases only moderately the total tidal disruption rate, which is dominated by stars originating from less bound orbits farther away. We show that the event rate of gravitational wave (GW) emission from inspiraling stars, originating much closer to the MBH, is dominated by RR dynamics. The GW event rate depends on the uncertain efficiency of RR. The efficiency indicated by the few available simulations implies rates $\\lesssim\\!10$ times higher than those predicted by 2-body relaxation, which would improve the prospects of detecting such events by future GW detectors, such as \\textit{LISA.} However, a higher, but still plausible RR efficiency can lead to the drainage of all tightly bound stars and strong suppression of GW events from inspiraling stars. We apply our results to the Galactic MBH, and show that the observed dynamical properties of stars there are consistent with RR. ", "introduction": "Galactic nuclei with massive black holes (MBHs) are stellar systems with relaxation times that are often shorter than the age of the Universe. In that case the distribution function (DF) of the system may approach a steady state. This steady state is determined by the boundary conditions (an inner sink, such as the tidal radius of the MBH, and an outer source at the interface with the host galaxy) and by the mutual interactions between the stars themselves. The nature of the {}``microscopic'' interactions between the stars determines the rate at which the system relaxes to its steady state. With the notable exception of $N$-body simulations (e.g. Baumgardt, Makino \\& Ebisuzaki \\cite{Baum04a}, \\cite{Baum04b}; Preto, Merritt \\& Spurzem \\cite{P04}; Merritt \\& Szell \\cite{MS05}), analyses of the evolution of the DF near a MBH have almost exclusively relied on the assumption that the mechanism through which stars exchange angular momentum and energy is dominated by \\textit{uncorrelated two-body interactions} (Chandrasekhar \\cite{Ch43}; see Binney \\& Tremaine \\cite{BT87} for a more recent discussion). This assumption is made in Fokker-Planck models (e.g. Bahcall \\& Wolf \\cite{BW76} {[}hereafter BW76{]}; Bahcall \\& Wolf \\cite{BW77} {[}hereafter BW77{]}; Cohn \\& Kulsrud \\cite{CK78}; Murphy, Cohn \\& Durisen \\cite{MCD91}), where the microscopic interactions are expressed by the diffusion coefficients, and in Monte Carlo simulations (e.g. Shapiro \\& Marchant \\cite{SM79}; Marchant \\& Shapiro \\cite{MS79}, \\cite{MS80}; Freitag \\& Benz \\cite{FB01}, \\cite{FB02}). Stars around MBHs are described as moving in the smooth average potential of the MBH and the stars, and the scattering by the fluctuating part of the potential is modeled as a hyperbolic Keplerian interaction between a passing star and a test star. The scattering effects accumulate non-coherently in a random-walk fashion. The (non-resonant) relaxation time $T_{\\mathrm{NR}}$ can be defined as the time $T_{\\mE}$ it takes for the negative specific energy $\\mE\\!\\equiv\\!-v^{2}/2-\\varphi$ of a typical star (hereafter {}``energy'') to change by order unity. This is also the time $T_{J}$ it takes for its specific angular momentum $J$ (hereafter {}``angular momentum'') to change by an amount of order $J_{c}(\\mE)$, the maximal (circular orbit) angular momentum for that energy% \\footnote{\\label{ft:TJ}Throughout this paper, {}``angular momentum relaxation time'' means the time it takes until $J$ is changed by order $J_{c}$, rather than by order $J$. The time-scale for changes by order $J\\!\\mE)\\Ms]/\\sqrt{2\\mE}$, where $\\Mbh$ is the MBH mass, $\\Ms$ is the stellar mass and $N(>\\!\\mE)$ is the number of stars with orbital energies above $\\mE$ (more bound than $\\mE$). On Keplerian orbits $J_{c}\\!=\\!\\sqrt{G\\Mbh a}$, where $a$ is the semi-major axis. When relaxation is dominated by uncorrelated two-body interactions, the {}``non-resonant'' relaxation time $T_{\\textrm{NR}}$ of stars of mass $\\Ms$ can be written in the Keplerian regime as \\begin{equation} T_{\\textrm{NR}}=A_{\\Lambda}\\left({\\frac{\\Mbh}{\\Ms}}\\right)^{2}{\\frac{P(a)}{N(10$, comparable to those estimated for AGN and much higher than the commonly-assumed values for X-ray binaries of 2--5. Jet power constraints do not in most cases rule out such high Lorentz factors. The upper limits on the opening angles show no evidence for smaller Lorentz factors in the steady jets of Cygnus X-1 and GRS\\,1915+105. In those sources in which deceleration has been observed (notably XTE\\,J1550--564 and Cygnus X-3), some confinement of the jets must be occurring, and we briefly discuss possible confinement mechanisms. It is however possible that all the jets could be confined, in which case the requirement for high bulk Lorentz factors can be relaxed. ", "introduction": "Proper motions of X-ray binary (XRB) jets have often been used to place limits on the jet Lorentz factors. \\citet{Fen03} recently argued that it was in fact impossible to do more than place a lower limit on the Lorentz factors of the jets from two-sided jet proper motions. For the persistent, continuous jets observed to exist in the low/hard X-ray states of black hole candidates, Gallo, Fender \\& Pooley (2003) found a universal correlation between the X-ray and radio fluxes of the sources, and used the scatter about this relation to constrain the Lorentz factors of such jets to $\\lesssim 2$. However, \\citet{Hei04} argued that the scatter about such a relation could not be used to constrain the mean Lorentz factor of the jets, but rather only the width of the Lorentz factor distribution. Other arguments, such as those based on jet power requirements, must be used to determine the absolute values of the jet Lorentz factors. Bulk jet flow velocities close to $c$, the speed of light, have been inferred in many XRB systems \\citep[e.g.][]{Mir94,Hje95,Fen04}, whereas the transverse expansion speeds have not yet been reliably measured. To date, there are few reported detections of XRB jets resolved perpendicular to the jet axis. This places strong upper limits on the opening angles of the jets (often less than a few degrees; see Table \\ref{tab:obs}), and hence on the transverse expansion speeds. While jets can in principle undergo transverse expansion at a significant fraction of $c$, time dilation effects associated with the bulk motion would reduce the apparent opening angle in the observer's frame. The magnitude of this effect would be determined by the bulk Lorentz factor of the flow. This raises the possibility of using the observed opening angle of a freely-expanding jet to constrain its Lorentz factor. Alternatively, if the Lorentz factors thus derived were incompatible with values deduced from independent methods, a strong argument could be made for jet confinement out to large (parsec) scales in such Galactic sources. In this paper, we first develop the formalism for deducing the Lorentz factor of a freely-expanding jet given a measurement of the opening angle and the inclination angle of the jet axis to the line of sight. In \\S\\,\\ref{sec:constraints} we compare constraints on the Lorentz factors derived from opening angle considerations with those from other methods. Transient and steady jets are compared in \\S\\,\\ref{sec:low-hard}, and the derived X-ray binary jet Lorentz factors are compared to those seen in AGN in \\S\\,\\ref{sec:AGN}. We discuss possible mechanisms for jet confinement and a method of using lightcurves to test confinement in \\S\\,\\ref{sec:confinement}. A summary of the observed properties of the individual sources is given in Appendix \\ref{sec:details}. ", "conclusions": "If XRB jets are not confined, but are expanding freely, it is possible to constrain their Lorentz factors from measurements of the jet opening angles. The small opening angles we observe are in this case a consequence of the transverse Doppler effect slowing the apparent expansion speed in the observer's frame. From the upper limits to the opening angles quoted in the literature, the Lorentz factors thus derived are significantly more relativistic (with a mean Lorentz factor $>10$) than is commonly assumed, and possibly no less so than AGN jets. However, if the jets we observe do indeed have Lorentz factors in the commonly-assumed range of 2--5, then we can quantify the degree of confinement. The lateral expansion speed perpendicular to the jet axis must then in some cases be $\\lesssim 0.15c$. In most cases, we cannot exclude the possibility that the jets are unconfined from jet power constraints, nor from measurements of the proper motions of knots in XRB jets. From the distances $d<R_{\\rm c}$ and the other $M_h(q)]= \\mathrm{\\Psi}(>q)$, where $\\mathrm{\\Phi}(>M_h)$ is the number density of galactic halos larger that $M_h$ and $\\mathrm{\\Psi}(>q)$ is the number density of galaxies with either stellar mass greater than $M_{\\mathrm{star}}$ or luminosity greater than $L_r$. The results are in good agreement with $M_h/L_r$ ratios inferred through X-ray mapping of the gravitational potential and through gravitational lensing. Both relations exhibit a double-power law shape with a break around $M_{h,{\\mathrm{break}}}\\approx 3\\times 10^{11}\\, M_{\\odot}$, corresponding to $M_{\\mathrm{star},\\mathrm{break}} \\approx 1.2\\times 10^{10}\\, M_{\\odot}$ and to an absolute magnitude $M_{r,{\\mathrm{break}}}\\approx -19.6$. A transition at about the same magnitude in the galaxy properties has been evidenced by Kauffmann et al. (2003) and Baldry et al. (2004). An additional interesting outcome of our analysis is that the $M_{\\mathrm{star}}$--$M_h$ relation is already established at redshift $z\\approx 1.7$, in line with the theoretical expectation of the anti-hierarchical baryon collapse scenario (Granato et al. 2004). Applying the same technique to the local velocity dispersion function of galaxies and to the black hole mass function, we have also computed the $\\sigma$--$M_h$ and $M_{\\rm BH}$--$M_h$ relationships. The former is quite close to a single power law $\\sigma \\propto M_h^{1/3}$. The latter is again a double power law breaking approximately at $M_{h,{\\mathrm{break}}}$, corresponding to $M_{\\rm BH}\\sim 9\\times 10^{6}\\, M_{\\odot}$. The associated velocity dispersion, $\\sigma \\simeq 88$ km s$^{-1}$, is very close to the first estimate of the critical velocity dispersion for the gas removal by SN explosions given by Dekel \\& Silk (1986), who found a critical halo velocity $V_{\\mathrm{crit}}\\sim 120$ km s$^{-1}$, corresponding to a critical velocity dispersion $\\sigma_{\\mathrm{crit}}\\sim 80\\,\\hbox{km}\\,\\hbox{s}^{-1}$. As a test of our results, we combined the $M_{\\rm BH}$--$M_h$ relation [eq.~(\\ref{eq|MbhMh})] with the $\\sigma$--$M_h$ relation [eq.~(\\ref{eq|sigmaMh})]; as shown by the lower panel of Fig.~\\ref{fig|MbhMh}, the resulting $M_{\\rm BH}$--$\\sigma$ relation is consistent with the observational data. The relationships we have obtained are model-independent and can be interpreted in terms of feedback effects by supernovae and AGNs in galactic structures. We also presented a simple feedback model, which nicely reproduces the approximate proportionalities $M_{\\rm BH}\\propto M_{\\mathrm{star}}\\propto M_h$ observed in the high mass range, and the break of these relationships at $M_{h,{\\mathrm{break}}}\\approx 3\\times 10^{11}\\, M_{\\odot}$. At low masses, the $M_{\\mathrm{star}}$--$M_h$ relation derived here ($M_{\\mathrm{star}}\\propto M_h^{3.1}$) is steeper than that yielded by the model ($M_{\\mathrm{star}}\\propto M_h^{5/3}$), and would imply that only a tiny fraction of the baryons initially associated with the halo remains within it in the form of stars (and we know that the gas does not add much to the baryon content of galaxies). On the other hand, if the amount of stars formed is so low, for a standard Salpeter IMF the energy injected by SN explosions is insufficient to expel the residual gas if the baryon fraction is close to the cosmic value. Thus, if the slope of the $M_{\\mathrm{star}}$--$M_h$ relation really is as steep as its face value suggests, we must conclude that either the initial baryon fraction in low-mass galaxies was substantially lower than the cosmic value (due, e.g., to a pre-heating of the intergalactic medium hampering the infall of baryons into shallow potential wells) or that the SN feedback in these objects was substantially stronger than in more massive galaxies. As discussed in \\S$\\,$\\ref{sect:ghmf}, however, the uncertainties on the low luminosity portion of the LF are large enough to allow for a flatter slope, closer to the model prediction and almost sufficient to grant the gas removal by SN feedback. The errors shown in the figures mostly reflect uncertainties in the $M/L$ ratio. We must not forget however, other error sources. For example, Fig.~\\ref{fig|MsLrMh}$b$ shows that different choices for the GHMF yield a systematic difference in the results, that, at the high-mass end, become comparable to the scatter considered in the same figure. Further uncertainties come from estimates of the GSMF; these are illustrated, in Figs.~\\ref{fig|MsLrMh}, ~\\ref{fig|Ratio}, ~\\ref{fig|Efficiency}, and ~\\ref{fig|feedback}, by comparisons with results obtained using the GSMF by Bell et al. (2003). Nevertheless, our approach provides results consistent with observations, and have comparable uncertainties. Moreover, since our approach bypasses any assumption on the DM profile, it could provide a valuable tool to discriminate among the different models of DM mass distribution in galaxies. Our analysis has shown that the relationships presented above bear the imprint of the processes ruling the galaxy formation and evolution. Models should eventually comply with them." }, "0601/astro-ph0601094_arXiv.txt": { "abstract": "In order to examine the ``giant impact hypothesis'' for the Moon formation, we run the first grid-based, high-resolution hydrodynamic simulations for an impact between proto-Earth and a proto-planet. The spatial resolution for the impact-generated disk is greatly improved from previous particle-based simulations. This allows us to explore fine structures of a circumterrestrial debris disk and its long-term evolution. We find that in order to form a debris disk from which a lunar-sized satellite can be accumulated, the impact must result in a disk of mostly liquid or solid debris, where pressure is not effective, well before the accumulation process starts. If the debris is dominated by vapor gas, strong spiral shocks are generated, and therefore the circumterrestrial disk cannot survive more than several days. This suggests that there could be an appropriate mass range for terrestrial planets to harbor a large moon as a result of giant impacts, since vaporization during an impact depends on the impact energy. ", "introduction": "In the current standard scenario of planet formation, the final stage of assemblage of terrestrial planets is collisions among proto-planets of about Mars mass \\citep{koku98,kort00}. It is widely accepted that at this final assemblage stage, the Moon is formed from the circumterrestrial debris disk generated by an off-set impact of the proto-Earth with a Mars-sized proto-planet, which is known as the ``Giant Impact (GI) hypothesis'' \\citep{hart75,came76}. The initial phase of the GI scenario, a giant impact and the formation of the circumterrestrial debris disk, has been studied by a series of hydrodynamic simulations \\citep{benz86,came00,canu01,canu04}. N-body simulations of the accumulation of the Moon from the debris disk, which is the last phase of GI scenario, revealed that a single large moon is formed just outside the Roche limit, at a distance of about three to four times the Earth's radius, within several months \\citep{ida97,koku00}. It is believed that this GI scenario explains a number of mysteries concerning the origin of the Moon \\citep{came00}: Why is the Moon so large compared to satellites of other planets? Why is the Moon deficient in iron and volatiles compared to the Earth? Why does the Earth-Moon system have large angular momentum? Almost all hydrodynamic simulations of GI in past decades used the Smoothed Particle Hydrodynamic (SPH) method, which is a particle-based Lagrangian scheme \\citep{lucy77,ging77}. In the SPH method, the numerical accuracy, in other words, the resolution, is determined by the number of SPH particles. The latest simulations \\citep{canu01,canu04} used $10^4-10^5$ SPH particles, which is an improvement of 1-2 orders of magnitude compared to simulations in the previous two decades \\citep{benz86,came00}. It is apparent, however, that the SPH method is not the best scheme to simulate an impact between two proto-planets. Firstly, SPH is not suitable to deal with the strong shocks and shear motion produced by the off-set impact. Moreover, since it is basically the Lagrangian scheme, the current SPH does not have resolutions fine enough for diffuse regions. This is critical problem particularly for GI, because the debris disk consists of only a few \\% of the total mass. As a result, even for a simulation with $10^5$ SPH particles, only a few $10^3$ SPH particles are used to represent the debris disk, and thus the fine structure of the disk is not resolved accurately. In the SPH formalism, the spatial resolution is determined by the `kernel size', which is variable with the local density. For a diffuse circumterrestrial disk, this kernel size can be as large as the radius of the disk itself \\citep[see Figure 1 of ][]{canu01}. This is the main reason that evolution of the debris is followed for very short period ($<1$ day) after the impact \\citep{canu01}. Therefore the post-impact evolution of the debris, which is a key for the GI hypothesis, is not well understood. Alternatively, one can use grid-based Eulerian methods to overcome the above problems. \\citet{melo89} have written the only preliminary study on the GI by the grid-based method to date. However, they neglected self-gravity of the planets and debris, and the evolution is followed only for less than 1 hour around the impact. Therefore, it cannot be compared to the recent SPH simulations. In this paper, we present the first three-dimensional hydrodynamic simulations of the giant impact followed by formation of a debris disk, taking into account the self-gravity, using a high-accuracy Eulerian-grid scheme. Our aim is to clarify the long-term (more than 100 hours after the impact) evolution of the debris with the highest numerical accuracy used in the previous simulations for GI. We pay attention especially to effects of pressure in the debris on formation of the large moon, rather than exploring the large parameter space for the mass ratio or orbits of the proto-planets. The previous SPH simulations with $10^4-10^5$ particles would be fine for studying the behavior of GI for the initial several hours, and they suggest that the fraction of mass orbiting the proto-Earth just after the impact is mainly determined by the orbital parameters. On the other hand, the long-term behavior of the debris, and therefore the final mass of the moon depends on the evolution of the debris disk. It is expected that the pressure in the low-density debris should affect the hydrodynamic nature of the disk, e.g., by generating shocks. Pressure in the debris is determined by the thermodynamics and the phase-change of the material after the impact, which is not fully understood for GI \\citep{stev87}. For example, if a substantial fraction of the material is in a hot vapor form after the impact, the thermal pressure dominates dynamics of the disk, but if liquid or solid is a major component of the debris, its behavior could be very different. Since there is no practical numerical codes to treat directly two-phase flow and phase-change, we run hydrodynamic simulations assuming two extreme equations of state, which approximately represent vapor or liquid-dominated materials. The paper is organized as follows. In \\S 2, we briefly describe our numerical method, equations of state, and initial conditions. We show the numerical results in \\S 3. We give conclusions and discuss an implication on a necessary condition of formation of large satellites for the Earth-type planets in \\S 4. ", "conclusions": "We run for the first time three-dimensional grid-base hydrodynamic simulations of the giant impact between proto-Earth and a proto-planet taking into account self-gravity. We assume two types of the equation of state that phenomenologically represent hot gaseous material or liquid/solid material. Our numerical experiments suggest that in order to form a lunar-sized satellite from the circumterrestrial debris disk produced by the giant impact, the most fraction of the debris should not be in a pressure-dominated phase (e.g., hot vapor gas). Otherwise the subsequent disk evolution results in forming only a small satellite because of the fast angular momentum transfer associated with spiral shocks prior to the satellite accretion stage. The time-scale of the angular momentum transfer is an order of 10 days, therefore the accumulation to form a large satellite should be much faster than this. Yet our results may not rule out a possibility forming a satellite as large as the Moon from a pressure-dominated disk, if the disk mass is much larger than the current lunar mass. This might be the case for collisions with a massive impactor, but this causes another problem on a fraction of vapor in the debris (see discussion below). In agreement with the previous SPH simulations \\citep[e.g.,][]{canu01, canu04}, we find that the predicted lunar mass at $t\\simeq 10$ hours after the collision does not strongly depend on choice of EOS and numerical methods for the same orbital parameters. This means that the early phase of GI is dominated by a gravitational process, not by thermodynamical processes. On the other hand, the late phase of GI, i.e., evolution of the debris disk, is sensitive to EOS, especially the pressure in the disk. In order to clarify this difference, the spatial resolutions of the previous SPH simulations were apparently not fine enough. Our results give an important implication on a necessary condition for formation of large satellites for the Earth-type planets. The present results suggest that one of the important keys for GI scenario is the fraction of the vaporized debris after the giant impact. If the kinetic energy of the collision is much larger than the latent heat of the major component of the proto-planet, most of the proto-planet could be vaporized by the impact. If this is the case, the disk evolution would not lead to formation of a large moon as explained above. On the other hand, if the impact velocity is slower than a critical value, a large fraction of the material can be in a liquid or solid phase, and shocks do not dominate the angular momentum transfer in the debris disk. In this case, the post-impact evolution of the proto-planet could be similar to that in model B in our experiments, and therefore a single large moon could be formed. By assuming the impact is head-on and comparing the latent heat of SiO$_2$ \\citep{stev87} and the kinetic energy of the impact, we can roughly estimate the critical impact velocity as $\\simeq $15 km s$^{-1}$, which is slightly larger than the surface escape velocity of the Earth. In the grazing impact between proto-Earth and a Mars-sized proto-planet, shocks propagating in the planets involve vaporization, and its velocity is probably comparable (or smaller) to this critical value, and therefore it would be natural to postulate that most of the debris mass is in a liquid phase. This argument leads to an interesting suggestion: if the impact velocity is larger than the critical velocity, in other words, the mass of the planet is larger than, say, a few Earth mass, the giant impact never results in forming a large satellite. Finally, one should recall that a correct phase-change even for the major component, such as SiO$_2$, during GI followed by formation of a debris disk is still unknown \\citep[e.g.][]{stev87}. For example, the post-impact expansion could produce more vapor in the debris material, and this may change the hydrodynamic property of the circumterrestrial disk. The final fate, namely whether the planet has satellites, and how large they are, depends on the nature of the phase-change. The thermo-dynamical process during catastrophic impact between the proto-planets is still too complicated to be explored using current numerical techniques including the SPH and grid-based methods. This ultimately requires a self-consistent numerical scheme to simulate two-phase flow taking into account the realistic thermodynamical processes." }, "0601/astro-ph0601607_arXiv.txt": { "abstract": "{Search for specific chemical signatures of intermittent dissipation of turbulence in diffuse molecular clouds. We observed \\HCOp(1-0) lines and the two lowest rotational transitions of \\twCO\\ and \\thCO\\ with an exceptional signal-to-noise ratio in the translucent environment of low mass dense cores, where turbulence dissipation is expected to take place. Some of the observed positions belong to a new kind of small scale structures identified in CO(1-0) maps of these environments as the locus of non-Gaussian velocity shears in the statistics of their turbulent velocity field \\ie\\ singular regions generated by the intermittent dissipation of turbulence. We report the detection of broad \\HCOp(1-0) lines ($10 {\\rm mK}< \\Tas < 0.5$ K). The interpretation of 10 of the \\HCOp\\ velocity components is conducted in conjunction with that of the associated optically thin \\thCO\\ emission. The derived \\HCOp\\ column densities span a broad range, $10^{11}< N(\\HCOp)/\\Delta v <4 \\times 10^{12}$ \\cq/\\kms, and the inferred \\HCOp\\ abundances, $2 \\times 10^{-10}