{ "0311/astro-ph0311385_arXiv.txt": { "abstract": "Magnetic fields in galaxy clusters can be investigated using a variety of techniques. Recent studies including radio halos, Inverse Compton hard X-ray emissions and Faraday rotation measure, are briefly outlined. A numerical approach for investigating cluster magnetic fields strength and structure is presented. It consists of producing simulated rotation measure, radio halo images, and radio halo polarization, obtained from 3-dimensional multi-scale cluster magnetic field models, and comparing with observations. ", "introduction": "A complete description of the astrophysical processes in cluster of galaxies requires knowledge of magnetic fields. The most detailed evidence for this component comes from the radio observations:\\\\ $\\bullet$ Some clusters of galaxies exhibit diffuse non-thermal synchrotron radio halos, associated with the intra-cluster medium, which extend up to Mega-parsec scales. Using minimum energy assumptions, it is possible to estimate an equipartition magnetic field strength averaged over the entire halo volume. These estimates give equipartition magnetic field strengths of $\\simeq$0.1 to 1 $\\mu$G (e.g. Bacchi et al. 2003).\\\\ $\\bullet$ In a few cases, clusters containing a radio halo show an hard X-ray excess emission. This emission could be interpreted in terms of Inverse Compton scattering of the cosmic microwave background photons with the relativistic electrons responsible for the radio halo emission. In this case, the measurements of the magnetic field strength (e.g. Fusco-Femiano et al. 1999, Rephaeli et al. 1999) inferred from the ratio of the radio to X-ray luminosities are consistent with the equipartition estimates.\\\\ $\\bullet$ Indirect measurements of the magnetic field intensity can also be determined in conjunction with X-ray observations of the hot gas, through the study of the Faraday Rotation Measure (RM) of radio sources located inside or behind clusters. By using a simple analytical approach, magnetic fields of $\\sim$$5-30$ $\\mu$G, have been found in cooling flow clusters (e.g. Allen et al. 2001; Taylor \\& Perley 1993) where extremely high RMs have been revealed. On the other hand, significant magnetic fields have also been detected in clusters without cooling flows: the RM measurements of polarized radio sources through the hot intra-cluster medium leads to a magnetic field of $2-8$ $\\mu$G which fluctuate on scales as small as $2-15$ kpc. (e.g. Feretti et al. 1995, Feretti et al. 1999, Clarke et al. 2001, Govoni et al. 2001, Taylor et al. 2001, Eilek \\& Owen 2002). The magnetic field strength obtained by RM studies is therefore higher than the value derived from the radio halo data and from Inverse-Compton X-ray studies. However, as pointed out by Carilli \\& Taylor (2002) and references therein, all the aforementioned techniques are based on several assumptions. For example, the observed RMs have been interpreted, until now, in terms of simple analytical models which consider single-scale magnetic fields. On the other hand, magneto-hydrodynamic cosmological simulations (Dolag et al. 2002) suggest that cluster magnetic fields may span a wide range of spatial scales with a strength that decreases with distance from the cluster center. We developed a numerical approach for investigating the strength and structure of cluster magnetic fields. It consists of comparing simulated rotation measure, radio halo images, and radio halo polarization, obtained from 3-dimensional multi-scale cluster magnetic field models, with observations (Murgia et al. 2003, submitted). ", "conclusions": "The numerical approach presented here demonstrates how the dispersion and mean of the RM measured in radio galaxies embedded in a cluster of galaxies can be used to constrain not only the strength but also the power spectrum slope of the intra-cluster magnetic fields. Moreover, the study of the polarization properties of a large scale radio halo, if it is present in a cluster, can be used to improve the estimates based on the RM analysis.\\\\ \\noindent {\\it Acknowledgment.}~ We thank L.Feretti, G.Giovannini, D.Dallacasa, R.Fanti, G.B. Taylor and K.Dolag for their collaboration." }, "0311/astro-ph0311499_arXiv.txt": { "abstract": "We present a general approach to the modelling of the brightness and polarization structures of adiabatic, decelerating relativistic jets, based on the formalism of \\citet{MS}. We compare the predictions of adiabatic jet models with deep, high-resolution observations of the radio jets in the FR\\,I radio galaxy 3C\\,31. Adiabatic models require coupling between the variations of velocity, magnetic field and particle density. They are therefore more tightly constrained than the models previously presented for 3C\\,31 by \\citet{LB02a}. We show that adiabatic models provide a poorer description of the data in two crucial respects: they cannot reproduce the observed magnetic-field structures in detail, and they also predict too steep a brightness decline along the jets for plausible variations of the jet velocity. We find that the innermost regions of the jets show the strongest evidence for non-adiabatic behaviour, and that the adiabatic models provide progressively better descriptions of the jet emission at larger distances from the galactic nucleus. We briefly discuss physical processes which might contribute to this non-adiabatic behaviour. In particular, we develop a parameterized description of distributed particle injection, which we fit to the observed total intensities. We show that particles are preferentially injected where bright X-ray emission is observed, and where we infer that the jets are over-pressured. ", "introduction": "\\label{Introduction} This paper is one of a series whose aim is to develop quantitative models of jets in low-luminosity (FR\\,I; \\citealt{FR74}) radio galaxies on the assumption that they are decelerating relativistic flows. Our hypothesis is that the jets are close enough to being intrinsically symmetrical, axisymmetric and antiparallel that the observed differences between them are dominated by the effects of relativistic aberration. This hypothesis is motivated by the results of \\citet{LPdRF}, who presented a statistical study of jets in the B2 sample of radio galaxies and showed that the observed correlations between fractional core flux density and side-to-side asymmetries in intensity and width are consistent with jet deceleration and the presence of transverse velocity gradients. They inferred that the jets slow from $\\approx$0.9$c$ where they start to expand rapidly to $\\approx$0.1$c$ over distances $\\sim$1 to 10\\,kpc and that the deceleration scale is an increasing function of jet power. In \\citet[][hereafter LB]{LB02a}, we demonstrated that an intrinsically symmetrical, relativistic jet model provides an excellent description of the total intensity and linear polarization observed at 8.4\\,GHz from the jets of the FR\\,I radio galaxy 3C\\,31. We were able to estimate the angle to the line of sight and the three-dimensional distributions of velocity, emissivity and magnetic-field structure. By combining this kinematic model with a description of the external gas density and pressure derived from {\\em Chandra} observations \\citep{Hard02} and using conservation of particles, energy and momentum, we demonstrated that the jet deceleration could be produced by entrainment of thermal matter and we derived the spatial variations of pressure, density and entrainment rate \\citep{LB02b}. The {\\em free models} developed by LB for 3C\\,31 were designed to fit the observed images without embodying specific preconceptions about the (poorly known) internal physics. In those models, we adopted simple and arbitrary functional forms for the spatial variations of velocity, synchrotron emissivity and field ordering and allowed the emissivity and field ordering to vary separately as smooth functions of position. To proceed beyond such purely empirical descriptions of the jets, we must make further assumptions about processes that affect the components of the modelled emissivity -- the relativistic particle energy spectrum and the strength of the magnetic field. The separation of the emissivity into its components is ill-determined unless inverse-Compton emission can be detected from the synchrotron-emitting electrons, in which case the particle density and field strength can be determined independently. Unfortunately, inverse-Compton emission from the inner jets in 3C\\,31 is too weak to detect \\citep{Hard02}. There are, as yet, no prescriptive theories for dissipative processes such as particle acceleration and field amplification or reconnection in conditions appropriate to FR\\,I radio jets. The energy-loss processes for the radiating particles can be quantified, however. It is inevitable that the particles will suffer adiabatic losses as the jets expand. We argue that synchrotron and inverse-Compton losses are negligible by comparison for electrons radiating at 8.4\\,GHz in 3C\\,31, as the jets have accurately power-law spectra with indices close to 0.5 \\citep{LPBFGMP04} and show no evidence of spectral curvature until much higher frequencies \\citep{Hard02}. It is therefore worthwhile to compare the observations with models in which the radiating particles are accelerated before entering the region of interest and then lose energy only by the adiabatic mechanism while the magnetic field is frozen into and convected passively with the flow, which is assumed to be laminar. Following conventional usage, we refer to such models as {\\em adiabatic}. There are analytical solutions for the adiabatic evolution of emissivity and surface brightness with jet radius and velocity if the magnetic field is exactly parallel or perpendicular to the flow direction. Models of this kind were first considered by \\citet{Bur79}, who showed that the decrease of brightness with distance from the nucleus would be much steeper than he observed in 3C\\,31 if the jets have constant velocity. \\citet{Fan82} pointed out that the brightness would decline more slowly with distance in a decelerating jet. This idea underlies the turbulent jet models of \\citet{ Bic84,Bic86}. A relativistic generalisation was developed by \\citet{Bau97} to model the jets in 3C\\,264 and was applied by \\citet {Fer99} and \\citet{ Bondi00} to other FR\\,I jets. These treatments all assumed that there is no velocity gradient across the jet, and that the magnetic field is exactly parallel or perpendicular to the flow. While this approach is self-consistent, it cannot be applied if there is velocity shear in a direction perpendicular to any component of the field. In LB, we developed a numerical approach to modelling a relativistic jet with a more general magnetic field structure. This approach allowed us to use the variations of total intensity and linear polarization as independent constraints on the jet velocity field. We concluded that the magnetic structure and velocity field in 3C\\,31 are indeed significantly more complex than those assumed by \\citet{Bau97}. Nevertheless, we compared our models and the data for 3C\\,31 with their analytical solutions. We showed that: \\begin{enumerate} \\item the adiabatic approximation is qualitatively inconsistent with the variations of brightness and polarization along the first $\\approx$3 kpc of the jets, but \\item further from the nucleus, the observed variations are closer to those expected if the adiabatic approximation holds. \\end{enumerate} This comparison motivated us to develop a more general approach to modelling of adiabatic, relativistic jets, which we present in this paper. Our approach allows us to calculate how the brightness and polarization structure evolve along an adiabatic jet from {\\it prescribed initial conditions} (specified as profiles across the jet), given the jet geometry and more complex magnetic structures and velocity fields of the type inferred for 3C\\,31 by LB. Detailed comparison of the new adiabatic models with the free models of LB, which fit the observations better, can then diagnose whether and where other physical processes, such as particle acceleration, may be significant in the jets. Section~\\ref{ad-approx} reviews our assumptions and the previous analytical solutions, and then outlines our calculation of synchrotron emission from adiabatic flows. Section~\\ref{Obs-mod} describes our approach to modelling of adiabatic jets; it briefly recapitulates material from LB before discussing new aspects specific to the present study. Section~\\ref{Outer-region} applies our adiabatic models to the outer regions of the jets in 3C\\,31 and shows that they can give a fair description of the VLA observations of these regions. Section~\\ref{Whole-jet} confirms that the adiabatic models fail to describe the inner jet regions and critiques the adiabatic hypothesis in the light of this result; it also discusses the extent to which distributed particle injection can bring the adiabatic models into better agreement with the data. Section~\\ref{Conclusions} summarizes our conclusions. ", "conclusions": "\\label{Conclusions} We have investigated the hypothesis that the radio jets in 3C\\,31 can be modelled as adiabatic, decelerating, relativistic flows. Our technique has several advantages over previous work in: modelling linear polarization as well as total intensity; including the effects of velocity shear; taking account of anisotropic emission in the rest frame and fitting to two-dimensional images rather than to longitudinal profiles. We have found that optimized adiabatic models give a fair description of the observed brightness and polarization distributions in the outer parts of the modelled region. Their fit to the data is inferior to that of the free models of LB, but is obtained with many fewer free parameters. The adiabatic models cannot describe the inner or flaring regions of the jets in 3C\\,31, however: the predicted distributions of total intensity and linear polarization are inconsistent with those observed. In the innermost region, the jets are clearly non-adiabatic unless the emission comes primarily from a part of the jet volume which is not expanding with distance from the nucleus (we do not resolve the region transversely). In the flaring region, a much higher initial velocity is required than is allowed by our measurements of jet/counter-jet sidedness ratio. The emission in this region is well resolved, and comes from the entire jet volume. We have shown that a modified adiabatic model can still be fitted to the total intensity in this region if we add distributed injection of relativistic particles which then evolve adiabatically; the region where these particles must be injected is also one where there is independent evidence for recent particle acceleration from the detection of X-ray synchrotron radiation \\citep{Hard02} and of a local over-pressure in the jet from dynamical arguments \\citep{LB02b}. While the total intensity distributions in the jet and counter-jet of 3C\\,31 can be reproduced satisfactorily by decelerating adiabatic jet models with particle injection in the flaring region, the polarization data (degree of linear polarization and apparent magnetic field direction) cannot. The apparent magnetic field configuration contains several features that are qualitatively incompatible with adiabatic evolution of all the field components in laminar-flow models, even in the presence of a velocity shear. We infer that the departures from adiabatic conditions in 3C\\,31 must also include deviations either from the flux-freezing description of the magnetic fields or from a laminar, axisymmetric velocity field. The free models of LB achieved better consistency with the observed polarization structure by allowing the field to become roughly isotropic at the edges of the jet in the flaring region. This could be achieved by a turbulent velocity component even if the field is frozen into the flow, in which case the resulting shear would also contribute to the enhancement in emissivity required to fit the adiabatic models in the flaring region. Processes such as dynamo action or field-line reconnection might also be significant. More detailed analysis of the particle injection process inferred for the flaring region would benefit from more sensitive X-ray, optical and infra-red data and higher-resolution radio observations of this region, i.e.\\ from deeper exposures with {\\em Chandra} and {\\em HST} and from deep higher-resolution polarimetry with the EVLA. Improved knowledge of the intensity and apparent magnetic field structures within this region might also assist development of models for the magnetic field microphysics in this region of the 3C\\,31 jet, which all of our analysis suggests is crucial in determining the deceleration dynamics of the jet, and its brightness and polarization properties further from the nucleus." }, "0311/astro-ph0311516_arXiv.txt": { "abstract": "Numerical calculations and linear theory of radiation magnetohydrodynamic flows indicate that the photon bubble and magnetorotational instability (MRI) may produce large density inhomogeneities in radiation pressure supported media. We study the effects of the photon bubble instability on accretion disk spectra using 2-D Monte Carlo (MC) and 1-D Feautrier radiative transfer calculations on a snapshot of a 2-D numerical simulation domain. We find an enhancement in the thermalization of the MC spectra over that of the Feautrier calculation. In the inner-most regions of these disks, the turbulent magnetic pressure may greatly exceed that of the gas. It is then possible for bulk turbulent Alfv\\'enic motions driven by the MRI to exceed the thermal velocity making turbulent Comptonization the dominant radiative process. We estimate the spectral distortion due to turbulent Comptonization utilizing a 1-D MC calculation. ", "introduction": "The standard assumption central to most accretion disk models is that the turbulence responsible for angular momentum transport and gravitational energy release can be parameterized by an ``$\\alpha$-viscosity'' prescription (Shakura \\& Sunyaev 1973). The hypothesis that turbulence drives accretion has been borne out by the identification of the magnetorotational instability (MRI, Balbus \\& Hawley 1991) as a generic source for magnetohydrodynamic turbulence in differentially rotating flows. However, it is not yet known how accurately $\\alpha$-disk models describe real accretion disks. It is possible that both the structure and emitted spectrum of a disk in which the MRI is active may deviate significantly from the predictions of ad hoc $\\alpha$-disk models. Numerical simulations of the radiation magnetohydrodynamic (RMHD) equations (Turner et al. 2003) show that MRI turbulence can produce a factor of 20 variation in density. In addition to the MRI, local linear analyses of the RMHD equations demonstrate that a radiatively driven photon bubble instability exists in radiation dominated accretion disks (Gammie 1998). This instability is present as long as the magnetic field is finite, but the growth rate for the instability is greatest when the magnetic pressure exceeds the gas pressure (Blaes \\& Socrates 2003). Numerical RMHD calculations of radiation pressure supported disks which neglect shear (Turner et al., in preparation) show that the non-linear development of the photon bubble instability does indeed produce large density variations. In section 2 we discuss the effects of these density inhomogeneities on the emergent spectrum. Fluid motions associated with the MRI may also influence the emission. MRI turbulence is fundamentally magnetic in nature with a characteristic velocity $v_w$ given by the Alfv\\'en speed $v_A$ which is less than the local sound speed $c_s$. Close to the hole, where the radiation pressure is dominant, $c_s$ is given by the radiation sound speed which is larger than the sound speed of the gas. Thus, it is possible for the MRI turbulence to be {\\it highly supersonic relative to the gas} and Compton scattering off of the turbulent eddies may produce a larger spectral modification than thermal Comptonization (Socrates, Davis, \\& Blaes 2003, hereafter SDB03). In section 3, we summarize these results. ", "conclusions": "Large density fluctuations due to the photon bubbles instability may alter the spectrum emitted in radiation pressure dominated accretion disks. These results are based on calculations which neglect shear. If shear is not neglected the MRI may affect the photon bubble instability, possibly altering the disk structure. However, for sufficiently large $v_A$, the MRI may be compressible in radiation dominated regions and may produce significant inhomogeneities on its own (Turner et al. 2003). We also analyze the interactions of photons with turbulent eddies and show that for sufficiently large turbulent stresses, a prominent spectral component in the far UV and X-ray bands emerges. Observations of accreting systems may provide a means to directly probe the mechanism responsible for angular momentum transport in these sources." }, "0311/astro-ph0311270_arXiv.txt": { "abstract": "We investigate self-similar dynamical processes in an isothermal self-gravitational fluid with spherical symmetry. In reference to earlier complementary solution results of Larson, Penston, Shu, Hunter and Whitworth \\& Summers, we further explore the `semi-complete solution space' from an initial instant $t\\rightarrow 0^{+}$ to a final stage $t\\rightarrow +\\infty$. These similarity solutions can describe and accommodate physical processes of radial inflow, core collapse, oscillations and envelope expansion (namely, outflow or wind) or contraction as well as shocks. In particular, we present new classes of self-similar solutions, referred to as `envelope expansion with core collapse' (EECC) solutions, that are featured by an interior core collapse and an exterior envelope expansion concurrently. The interior collapse towards the central core approaches a free-fall state as the radius $r\\rightarrow 0$, while the exterior envelope expansion gradually approaches a constant radial flow speed as $r\\rightarrow +\\infty$. There exists at least one spherical stagnation surface of zero flow speed that separates the core collapse and the envelope outflow and that travels outward at constant speed, either subsonically or supersonically, in a self-similar manner. Without crossing the sonic critical line where the travel speed of nonlinear disturbances relative to the radial flow is equal to the sound speed, there exist a continuous band of infinitely many EECC solutions with only one supersonic stagnation point as well as a continuous band of infinitely many similarity solutions for `envelope contraction with core collpase' (ECCC) without stagnation point. Crossing the sonic critical line twice analytically, there exist infinitely many discrete EECC solutions with one or more subsonic stagnation points. Such discrete EECC similarity solutions generally allow radial oscillations in the subsonic region between the central core collapse and the outer envelope expansion. In addition, we obtained complementary discrete ECCC similarity solutions that cross the sonic critical line twice with subsonic oscillations. In all these discrete solutions, subsonic spherical stagnation surfaces resulting from similarity oscillations travel outward at constant yet different speeds in a self-similar manner. With specified initial boundary or shock conditions, it is possible to construct an infinite number of such EECC similarity solutions, which are conceptually applicable to various astrophysical problems involving gravitational collapses and outflows. We mention potential applications of EECC similarity solutions to the formation process of proto-planetary nebulae connecting the AGB phase and the planetary nebula phase, to H{\\sevenrm II} clouds surrounding star formation regions, and to a certain evolution phase of galaxy clusters. ", "introduction": "Hydrodynamical processes of a self-gravitational spherical gas have been investigated from different perspectives and in various contexts of astrophysical or cosmological problems on entirely different temporal and spatial scales. As an important example, solutions to the model problem of spherical gravitational collapse provide physical insights for understanding processes of star formation and outflows (Ebert 1955; Bonner 1956; Hayashi 1966; Hunter 1967; Spitzer 1968; Larson 1973; Mestel 1974). Another important field is the formation and evolution of galaxy clusters that contain massive dark matter haloes and hot gases with temperatures of $\\sim 10^7-10^8$K (e.g. Fabian 1994; Sarazin 1988; Gunn \\& Gott 1972; Bertschinger 1985; Fillmore \\& Goldreich 1984; Navarro, Frenk \\& White 1996). Detailed numerical computations to solve the fully coupled nonlinear partial differential hydrodynamic equations show that solutions, sufficiently away from initial and boundary conditions, appear to evolve into self-similar behaviours (e.g. Sedov 1959; Landau \\& Lifshtiz 1959). Over past several decades, researchers have studied self-similar solutions under spherical symmetry and isothermal conditions. Simultaneously and independently, Penston (1969a, b) and Larson (1969a, b) found one discrete self-similar infall solution without involving a central core mass, the so-called LP-solution referred to in the literature, on the basis of their numerical calculations of spherically symmetric gravitational collapse for star formation. In addition to this LP-solution, Shu (1977) derived other types of self-similar solutions, including the so-called `expansion-wave collapse solution' (with a discontinuous derivative at the sonic critical point), which leads to the physical scenario of `inside-out collapse' in the context of molecular cloud and/or star formation (e.g. Shu, Adams \\& Lizano 1987). Shu (1977) showed clearly the existence of the sonic critical line across which two types of eigensolutions may be possibly specified. Following Shu's analysis, Hunter (1977) found a class of discrete `complete solutions' by extending the time domain way back to the pre-catastrophic period starting from $t\\rightarrow -\\infty$ and by applying a zero-speed boundary condition there (see Fig. 6). One may regard the LP-solution as a special member of Hunter's class and Shu's expansion-wave collapse solution turns out to be a limiting case of Hunter's solution class [see solution parameters (11) and figs. 1 and 2 of Hunter (1977)]. In terms of the `semi-complete solution' structure (see Fig. 6), the `complete' LP-solution and each of `complete' solutions $a$, $b$, $c$, $d$ and so forth of Hunter's (1977) can be split into two branches (i.e. one inflow and one outflow) at $x\\equiv r/(at)\\rightarrow +\\infty$. In the `semi-complete' perspective, Hunter's solutions $a$, $b$, $c$, $d$ and so forth involve radial oscillations with increasing number of nodes or stagnation points in the subsonic region with either envelope expansion or envelope contraction but without initial onset of central core collapse (see Fig. 6). Within the same overall framework of Shu (1977) and Hunter (1977), Whitworth \\& Summers (1985, hereafter WS) introduced more elaborate mathematical techniques to construct similarity solutions across the sonic critical line with weak discontinuities (Lazarus 1981; Hunter 1986; Ori \\& Piran 1988; Boily \\& Lynden-Bell 1995). Whitworth \\& Summers noted the existence of two types of sonic points (namely, saddle and nodal points defined later) in the problem (see Jordan \\& Smith 1977 for details) and carefully examined stability properties of numerical integration directions. They realized the distinctly different properties of Shu's two types of eigensolutions at the sonic critical line and referred to the two types as `primary' and `secondary' directions\\footnote{For the two eigensolutions crossing the nodal sonic critical point and in terms of absolute values, the one of larger flow speed gradient is along the primary direction and the one of smaller flow speed gradient is along the secondary direction.} in the case of nodal points (Jordan \\& Smith 1977). By allowing one solution to pass the nodal sonic point along the secondary direction and an infinite number of solutions to pass it tangentially along the primary direction as well as locally linear combinations of the two eigensolutions, WS constructed two-parameter continuum bands of solutions in addition to the previous discrete solutions. Regarding the analysis of WS, Hunter (1986) promptly pointed out that the solutions of WS involve weak discontinuities across the sonic critical line in that physical variables and all their first derivatives in space and time are continuous\\footnote{This does not include those cases when a solution on one side is along the primary direction and that on the other side is along the secondary direction.} and derivatives of some higher orders may also be continuous. Moreover, Hunter (1986) suggested that the solutions of WS may be unstable against small perturbations as numerical computations did not show tendency towards any of such solutions except for the LP-solution and therefore might not be acceptable physically. Ori \\& Piran (1988) proceeded to verify Hunter's conjecture and derived a necessary stability criterion for `complete solutions'. Lynden-Bell \\& Lemos (1988) and Lemos \\& Lynden-Bell (1989a, 1989b) studied self-similar solutions for a cold fluid, not only in the Newtonian regime but also in the regime of general relativity. Foster \\& Chevalier (1993) studied the gravitational collapse of an isothermal sphere by hydrodynamic simulations. They recovered the LP-solution in the central region where a core forms and the self-similar solutions of Shu (1977) when the ratio of initial outer cloud radius to core radius is $\\gsim 20$. Boily \\& Lynden-Bell (1995) studied the self-similar solutions for radial collapse and accretion of radiative gas in the polytropic approximation. Hanawa \\& Nakayama (1997) also investigated the stability problem of the complete self-similar solutions in a normal mode analysis and concluded that only the LP-solution is stable. The most recent work of Harada, Maeda \\& Semelin (2003) studied spherical collapse of an isothermal gas fluid through numerical simulations and showed a critical behaviour for the Newtonian collapse that turns out to follow the self-similar form of the first member (i.e. solution $a$) of Hunter's solutions (Hunter 1977). Lai \\& Goldreich (2000) studied the growth of nonspherical perturbations in the collapse of a self-gravitating spherical gas cloud. They found through numerical computations that nonspherical perturbations damp in the subsonic region but grow in the supersonic region with asymptotic scaling relations for their growths. Lai \\& Goldreich mentioned potential applications to core-collapse of supernova explosions (Goldreich \\& Weber 1980; Yahil 1983), where the asymmetric density perturbation may lead to asymmetric shock propagation and breakout, giving rise to asymmetry in the explosion and a kick to the new-borne neutron star. Subsequently, Lai (2000) presented a global stability analysis for a self-similar gravitational collapse of a polytropic gas. In addition to the shock-free self-similar solutions, Tsai \\& Hsu (1995) found a similarity shock solution in reference to Shu's expansion-wave collapse solution through both numerical and analytical analyses for the formation of low-mass stars. Their results were recently expanded by Shu, Lizano, Galli, Cant\\'o \\& Laughlin (2002) in the context of self-similar `champagne flows' driven by a shock in H{\\sevenrm II} regions. One shock solution found by Tsai \\& Hsu (1995) turns out to be the limiting case of many self-similar shocked LP-solutions found by Shu et al. (2002). The perspectives of complete (Hunter 1977, 1986) and semi-complete (Shu 1977; WS) similarity solutions are both valid with proper physical interpretations and with the corresponding identification of an initial moment. In view of the invariance under the time-reversal transformation of self-gravitational fluid equations $(1)-(3)$ below, that is, $t\\rightarrow -t$, $\\rho\\rightarrow\\rho$, $u\\rightarrow -u$, the complete and semi-complete similarity solutions are closely related to each other. For example, the complete LP-solution (fig. 1 of Hunter 1977) describes a self-similar collapse process starting from the pre-catastrophic instant $t\\rightarrow-\\infty$ with a zero flow speed and a density distribution, passing through the sonic critical point once, gaining a finite radial infall speed at $t=0$, and approaching eventually a free-fall state involving a central core collapse as $t\\rightarrow +\\infty$. However, the complete LP-solution (Hunter 1977) can be broken into two branches of similarity solution in the semi-complete space. One branch is only partially shown in the lower-left corner of Shu's fig. 2 and is described by equations $(A7)$ and $(A8)$ in Shu's appendix (see also Fig. 5). This branch corresponds to a self-similar outflow without core mass but with a finite central density decreasing with time $t$ in a scaling of $t^{-2}$. The other branch corresponds a constant radial inflow speed at $x\\rightarrow +\\infty$ and a core collapse at $x\\rightarrow 0^{+}$ without crossing the sonic critical line. To some extent, Shu's criticism on the utility of the semi-complete LP-solution may be compromised by joining the other branch of the semi-complete LP-solution of infall-core-collapse solution (Hunter 1977) or by inserting a shock to match with an outward `breeze' for `champagne flows' in H{\\sevenrm II} regions (Tsai \\& Hsu 1995; Shu et al. 2002). With the complementarity of complete and semi-complete perspectives for self-similar solutions, we re-visit this classic problem in the semi-complete solution space with the strong motivation of finding similarity solutions for concurrent processes of central core collapse and envelope expansion or contraction that may or may not involve subsonic radial oscillations and/or shocks across the sonic critical line. Specifically, we have learned Shu's expansion-wave collapse solution as the limit of a class of core collapse solutions with zero infall speed at $x\\rightarrow +\\infty$. We have also noted that Hunter's complete collapse solutions, including the LP-solution, can be used to describe radial outflows with or without subsonic radial oscillations in the semi-complete perspective. It is then physically plausible to derive self-similar solutions for concurrent envelope expansion with core collapse (EECC) that may or may not involve subsonic radial oscillations and/or shocks. If one allows solutions that pass through the sonic critical line with weak discontinuities (Lazarus 1981), then the mathematical solution space can be further expanded enormously (WS). This paper gives a positive account of constructing EECC similarity solutions that are distinctly different from known similarity solutions. With proper selections and adaptations, we believe that these EECC similarity solutions, due to their generality and simplicity, provide an important basic conceptual framework for a wide range of astrophysical problems involving gravitational collapses, outflows, contractions and subsonic radial oscillations etc. We shall mention a few potential applications, including the dynamical process of forming a proto-planetary nebula that evolves from the asymptotic giant branch (AGB) phase to the planetary nebula phase with a central hot proto white dwarf. Specifics will be described in forthcoming papers. Several authors have approached the same set of basic isothermal fluid equations with different notations, choice of variables and conventions of presenting their results. It is sometimes mind-boggling to find correspondences among their solutions and results of analysis. Without further complications, we shall more or less adopt Shu's notations and present our results in a similar manner. The basic equations and the self-similar transformation are summarized in Section 2. In contrast to Shu's focus on the first quadrant of $-v$ versus $x$, we pay equal attention to both first and fourth quadrants for $-v$ versus $x$ presentation as we are interested in both outflows and collapses. In order to place our work in a proper perspective in reference to known results, we have explored the entire solution structure, including the counterparts of those solutions found by previous authors, in Section 3. Physical interpretations for the obtained solutions are described in Section 4. We summarize and discuss our results in Section 5. Some technical details and procedures can be found in the Appendix. ", "conclusions": "In this paper, we investigated the spherical self-similar solutions for self-gravitating isothermal flows in the `semi-complete solution space' extending from the initial instant $x\\rightarrow +\\infty$ to the final instant $x\\rightarrow 0^{+}$. The relevant similarity solutions are obtained and classified to compare with previous solutions and analyses [Larson 1969a; Penston 1969a; Shu 1977; Hunter 1977; Whitworth \\& Summers 1985 (WS)]. The novel class of infinitely many similarity solutions obtained in this paper, i.e., the EECC self-similar solutions and the envelope contraction with core collapse (ECCC) self-similar solutions (e.g. solution labelled 2 in Figs. 10 and 11 that passes the sonic critical line twice with specific parameters $V=-0.766$, $A=1.209$ and $m_0=5.171\\times 10^{-4}$), is constructed in the `semi-complete solution space'. This is to be compared with the similarity solutions derived in the `complete solution space' first introduced by Hunter (1977). By the invariance property under time reversal in this classical problem, these new self-similar solutions may be formally extended to the `complete solution space' by properly joining two separate branches in the first and fourth quadrants of the `semi-complete solution space' but with diverging asymptotic behaviours of $v(x)$ and $\\alpha(x)$ at $t\\rightarrow -\\infty$. Similarity solutions without crossing the sonic critical line as shown in Figs. 2 and 3 also carry ECCC and EECC features, respectively, and are much easier to construct technically in the isothermal approximation; in a polytropic gas with various laws of radiative losses, some qualitatively similar features were mentioned in subsection 8.2 of Boily \\& Lynden-Bell (1995). Regarding the isothermal approximation, the core collapsing (CC) portion may be justified qualitatively on the ground that the gas would heat up by adiabatic compressions. In the presence of a sufficient amount of radiative agents in a relatively high density region involving more frequent collisions, such radiative cooling processes may sustain an approximate condition of quasi-constant temperature. To sustain an isothermal condition for the envelope expansion (EE) portion, some heating mechanisms are necessary to compensate the inevitable cooling resulting from adiabatic expansions. Naturally, radiations from gas in the CC portion can be absorbed by gas in the EE portion to heat up the latter to some extent (depending on the escape efficiency of photons). In the astrophysical context of star formation, a gas cloud can be submerged in an intense electromagnetic radiation field when the nuclear reaction has been ignited in the central protostar. In such a situation, the large-scale gas cloud dynamics of contraction and expansion may be regarded as quasi-isothermal. Similar conditions may be applicable in the emergence of a planetary nebula system as well as in a certain evolution phase for hot gas in a galaxy cluster. Although not fully understood in details, there are other possible heating processes that may happen in the EE portion such as micro-turbulence, flow instabilities, shocks, wave dampings and magnetic fields etc. With these considerations in mind, it is worthwile to note that the isothermal assumption is a special case of the more general polytropic approximation (Cheng 1978; Goldreich \\& Weber 1980; Bouquet et al. 1985; Yahil 1983; Suto \\& Silk 1988; Maeda et al. 2002; Harada et al. 2003). Asymptotic solutions parallel to those of $(16)-(18)$ as well as eigensolutions across the sonic critical line can be readily derived. Therefore, EECC self-similar solutions in the polytropic approximation can be obtained with combined numerical and analytical analyses. One major difference is that, in the polytropic approximation, the asymptotic solution takes the form of \\begin{equation} v\\rightarrow Vx^{(1-\\Gamma)/(2-\\Gamma)}\\ , \\end{equation} \\begin{equation} \\alpha\\rightarrow Ax^{-2/(2-\\Gamma)}\\ , \\end{equation} \\begin{equation} m\\rightarrow Dx^{(4-3\\Gamma)/(2-\\Gamma)}\\ , \\end{equation} where $x\\propto rt^{(\\Gamma -2)}$ is the independent similarity variable, $\\Gamma$ is the polytropic index, and $V$, $A$ and $D$ are three constant coefficients to the leading order. For $1<\\Gamma<2$, the radial flow speed (either outflows or inflows) vanishes as $x\\rightarrow +\\infty$. For a finite radial range, the basic concept of EECC similarity solutions advanced in this paper remains valid. Additional similarity solution features such as shocks, subsonic radial oscillations, central core collapse etc. are also expected. Polytropic EECC similarity solutions should have a wider range of astrophysical applications with, at least, one more degree of flexibility (i.e. $\\Gamma$ parameter). Conceptually, the EECC similarity solutions provide a simple model scenario for the concurrence of inner infalls (including central core collapses) and outer envelope outflows (including asymptotic winds) with the separation surface (or stagnation point) travelling outward at a constant speed. In any self-gravitating spherical fluid system, the emergence of an EECC similarity phase depends upon allowed physical conditions that can be reached via a certain set of initial/boundary conditions of density and velocity at some time ago through other processes that were usually not self-similar. For applications of the EECC similarity solutions, as well as other similarity solutions, we have noted several potential areas. The first one is the dynamical evolution phase connecting the AGB or post-AGB phase and the proto planetary nebula (PN) phase, where the central core collapse to form a proto white dwarf and the envelope wind to create various PN morphologies are expected to happen simultaneously in a timescale of $\\sim 10^{3}$ yrs (e.g. Balick \\& Frank 2002). By this process, a late-type star can continue to lose its envelope mass while forming a compact object at the collapsing core. After the self-similar phase, the system continues to evolve in a detached manner with the core and envelope separated. Shocks can be introduced either at the end of the AGB phase (Kwok 1982, 1993) when the faster envelope expansion during the EECC phase catches up with the slower pre-existing wind or when infalling materials bounced back from the central compact object. Much as Tsai \\& Hsu (1995) and Shu et al. (2002) introduced shocks to the similarity solutions in star formation and `champagne flows' of H{\\sevenrm II} regions, one can construct EECC similarity solutions with shocks in the context of planetary nebulae. The solutions other than the EECC ones (including their time-reversal counterparts) have been used in star-formation as well as other gravitational collapse and flow problems. For astrophysical systems of even larger scales such as clusters of galaxies (e.g. Sarazin 1988; Fabian 1994), the similarity solutions can be valuable for understanding a certain phase of their evolution. Specifically, if their evolution involves a self-similar phase of central collapse (e.g. Gunn \\& Gott 1972; Fillmore \\& Goldreich 1984; Bertschinger 1985; Navarro et al. 1997), then isothermal similarity solutions (17) and (18) seem to suggest two possible classes of galaxy clusters that emit X-rays through hot gases virialized in gravitational potential wells, namely, those with steep gravitational potential wells and thus extremely high X-ray core luminosities, and those with relatively smooth and shallow gravitational potential wells and thus normal X-ray core luminosities. Finally, we note that the stability of similarity solutions is of interest in various contexts. As several authors have declared (Ori \\& Piran 1988; Hanawa \\& Nakayama 1997), some self-similar solutions may not be stable when one performs a normal mode analysis. The instability growth rates are different for various types of similarity solutions in general. At this moment, the stability problem for the EECC similarity solution remains open for future research. In some cases, depending on the growth rates, an unstable EECC similarity solution might be of interest to disrupt a self-similar phase within proper timescales. For example, we do not expect a self-similar phase to last forever, especially for the ephemeral phase of $\\sim 10^3$ yrs between the end of AGB phase and the PN phase (Kwok 1982, 1993; Balick \\& Frank 2002)." }, "0311/astro-ph0311046_arXiv.txt": { "abstract": "The $\\simeq$ 100 kg highly radiopure NaI(Tl) set-up of the DAMA project (DAMA/NaI) took data over seven annual cycles up to July 2002 and has achieved results on various rare processes. Its main aim has actually been the exploitation of the model independent WIMP annual modulation signature. After this conference the total exposure, collected during the seven annual cycles, was released. This cumulative exposure (107731 kg $\\times$ day) has given a model independent evidence for the presence of a Dark Matter particle component in the galactic halo at 6.3 $\\sigma$ C.L.; this main result is summarised here. Some of the many possible corollary model dependent quests for the candidate particle are mentioned. At present, after about five years of new developments, a second generation low background set-up (DAMA/LIBRA with a mass of $\\simeq$ 250 kg NaI(Tl)) was built and is taking data since March 2003. New R\\&D efforts toward a possible NaI(Tl) ton set-up, we proposed in 1996, have been funded and started in 2003. ", "introduction": "The DAMA project has been proposed by the italian group to INFN and firstly funded in 1990 \\cite{1}; in 1992 the chinese colleagues joined the project. DAMA is an observatory for rare processes based on the development and use of various kinds of radiopure scintillators. Several low background set-ups have been realised; the main ones are: i) DAMA/NaI ($\\simeq$ 100 kg of radiopure NaI(Tl)), which took data underground over seven annual cycles and was put out of operation in July 2002; ii) DAMA/LXe ($\\simeq$ 6.5 kg liquid Xenon), which has been upgraded various times and is in operation; iii) DAMA/R\\&D, which is a set-up devoted to measurements on prototypes/small--scale--experiments and is in operation; iv) the new second generation set-up DAMA/LIBRA ($\\simeq$ 250 kg; more radiopure NaI(Tl)) in operation since March 2003. Moreover, in the framework of devoted R\\&D for higher radiopure detectors and PMTs, sample measurements are regularly carried out by means of the low background DAMA/Ge detector, installed deep underground since $\\gsim$ 10 years and, in some cases, by means of Ispra facilities. Recent results from DAMA/LXe are presented in these proceedings elsewhere \\cite{2}, while in ref. \\cite{3} some recent results achieved with DAMA/R\\&D can be found. The DAMA/NaI set-up and its performances have been described in details in ref. \\cite{Nim98}. Since then some upgrading has been carried out; in particular, in summer 2000 the electronic chain and data acquisition system were completely substituted as reported in ref. \\cite{RNC}. As mentioned, DAMA/NaI has taken data over seven annual cycles up to July 2002 \\cite{RNC}. The first data release and publication of DAMA/NaI data occurred in 1996 \\cite{Psd96} for an exposure of 4123 kg $\\times$ day (DAMA/NaI-0) analysed by using the pulse shape discrimination (PSD) technique in NaI(Tl) in order to discriminate between electromagnetic background and recoils\\footnote{The NaI(Tl) Dark Matter experiments have discrimination capability as largely stated in literature by several groups and at present exploited as main goal by the NAIAD experiment. In Zeplin I the pulse shape discrimination technique has been employed in LXe as well (sometimes this rejection procedure has been indicated as rejection by {\\it timing}), similarly as done and published by DAMA/LXe in 1998. Moreover, we remind that the WIMP annual modulation signature acts (as pointed out since the 80's by \\cite{Freese}) itself as a powerful background rejection procedure and that it can be effectively exploited only in large mass experiments such as so far DAMA/NaI and DAMA/LIBRA.}. Investigation of possible diurnal effects has been carried out as well \\cite{Diu}. The main aim of the experiment was actually the investigation of the presence of a Dark Matter particle component in the galactic halo by means of the model independent WIMP annual modulation signature, which has been deeply investigated over seven annual cycles (107731 kg $\\times$ day total exposure). DAMA/NaI has also obtained results on the investigation of some exotic Dark Matter candidates (such as neutral SIMPs, neutral nuclearities and Q-balls) \\cite{exotic}, has searched for solar axions \\cite{Dax} and has given results on several other rare processes such as e.g. spontaneous emission of protons in $^{23}$Na and $^{127}$I with violation of the Pauli exclusion principle \\cite{35}, nuclear level excitation of $^{23}$Na and $^{127}$I during charge-non-conserving processes \\cite{cnc}, electron stability and non-paulian transitions in Iodine atoms by L-shell\\cite{Ela99}. In the following, the result obtained by DAMA/NaI in the investigation of the WIMP annual modulation signature over the seven annual cycles will be briefly summarised. This cumulative 107731 kg $\\times$ day exposure has given a model independent evidence for the presence of a Dark Matter particle component in the galactic halo at 6.3 $\\sigma$ C.L.; this main result and corollary interpretations in some of the possible different model scenarios have been discussed in experimental and theoretical details in ref. \\cite{RNC} and have firstly been presented after this Conference in a seminar at Gran Sasso National Laboratory (LNGS) at end of July 2003. The new second generation larger mass and higher radiopure NaI(Tl) set-up, named DAMA/LIBRA, is now in operation and will be briefly introduced. ", "conclusions": "DAMA/NaI has been a pioneer experiment running at LNGS for about a decade and investigating as first the WIMP annual modulation signature with suitable exposed mass, sensitivity and control of the running parameters. During seven independent experiments of one year each one, it has pointed out the presence of a modulation satisfying the many peculiarities of a WIMP induced effect, reaching an evidence at 6.3 $\\sigma$ C.L.. No other experiment has so far been in position to give a result directly comparable in a model independent way with that of DAMA/NaI. As a corollary, it has also pointed out the complexity of the quest for a WIMP candidate because of the present poor knowledge on the many astrophysical, nuclear and particle physics aspects. After the completion, on July 2002, of the DAMA/NaI data taking, the second generation DAMA/LIBRA set-up ($\\simeq$ 250 kg mass) has been installed and preliminarily put in operation, as a result of continuous efforts by the DAMA collaboration toward the creation of ultimate radiopure NaI(Tl). DAMA/LIBRA, having a larger exposed mass and an higher overall radiopurity, will significantly contribute in the incoming years to the further understanding of the field. Presently, new R\\&D efforts toward a possible NaI(Tl) ton set-up have been funded and related works have already been started." }, "0311/astro-ph0311100_arXiv.txt": { "abstract": "We present a simple technique to estimate mass-to-light ($M/L$) ratios of stellar populations based on two broadband photometry measurements, i.e.\\ a color-$M/L$ relation. We apply the color-$M/L$ relation to galaxy rotation curves, using a large set of galaxies that span a great range in Hubble type, luminosity and scale size and that have accurately measured H{\\small I} and/or H$\\alpha$ rotation curves. Using the color-$M/L$ relation, we construct stellar mass models of the galaxies and derive the dark matter contribution to the rotation curves. We compare our dark matter rotation curves with adiabatically contracted Navarro, Frenk, \\& White (1997, NFW hereafter) dark matter halos. We find that before adiabatic contraction most high surface brightness galaxies and some low surface brightness galaxies are well fit by a NFW dark matter profile. However, after adiabatic contraction, most galaxies are poorly fit in the central few kpc. The observed angular momentum distribution in the baryonic component is poorly matched by $\\Lambda$CDM model predictions, indicating that the angular momentum distribution is not conserved during the galaxy assembly process. We find that in most galaxies the dark matter distribution can be derived by scaling up the H{\\small I} gas contribution. However, we find no consistent value for the scaling factor among all the galaxies. ", "introduction": "To constrain the stellar mass distribution in our sample galaxies, we use the color-$M/L$ relationships for mixed stellar populations of Bell \\& de Jong (2001). In this paper we showed that for almost any reasonable star formation history and consistent chemical evolution of a galaxy, the stellar population models predict a strong correlation between the optical color of the population and its $M/L$. This was shown to be robust against gas inflow and outflow, different star formation prescriptions and mild starbursts (Fig.\\,1a). \\begin{figure} \\mbox{ \\epsfysize=5.8cm \\epsfbox[74 355 216 513]{dejong_r_fig1a.eps} \\epsfysize=5.8cm \\hspace{1.cm} \\epsfbox[188 370 410 597]{dejong_r_fig1b.eps} } \\caption{ {\\bf Left (a):} Bell \\& de Jong (2001) model stellar population $M/L$ ratios in the $B$, $I$ and $K$-band versus $B$--$R$ color (points) and the fitted slopes (lines). {\\bf Right (b):} The maximum disk $M/L_K$ ratios for the Verheijen (1997) galaxies (upper limit symbols) versus $B$--$R$ color. The solid line shows the stellar population model predictions for a Salpeter IMF (solid line) and a renormalized Salpeter IMF (dotted line).} \\end{figure} The one major assumption we had to make was that all galaxies have similar stellar Initial Mass Functions (IMFs), independent of their other properties. The IMF used in our color-$M/L$ analysis determines the slope of the relation, but not the zero-point. This reflects the fact that the high mass end of the IMF determines the luminosity and the color of a population, but the low mass end sets its mass, shifting the relation in Fig.\\,1a up and down. We used the maximum disk $M/L_K$ limits of the Ursa Major galaxies observed by Verheijen (1997) to constrain the IMF. In Fig.\\,1b we show the maximum allowed $M/L_K$ ratios for these galaxies (hence the upper limit symbols) versus their $B$--$R$ color. We compare these to the predictions of our models for a standard Salpeter IMF (solid diagonal line). If our models are correct, all $M/L$ ratios of galaxies should scatter around this line. Clearly, this is in conflict with observations, as some galaxy upper limits are below the line, and thus using the $M/L$ of our model would over-predict the rotation curve of these galaxies. Hence, we use a normalization that is lower by 0.3 dex (dotted line). In principle, the normalization could be much lower, but there are many indications that in particular high surface brightness galaxies are close to maximum disk (see e.g.\\ Weiner, these proceedings) and therefore we chose a normalization that has these galaxies close to maximum disk. If our color-$M/L$ relation is correct and all true $M/L$ values are near this line, clearly many galaxies are substantially sub-maximal, with some galaxies in the top part of the diagram being sub-maximal by a factor of 10. We use here the updated color-$M/L$ relations of Bell et al.~(2003), which were calibrated with SDSS and 2MASS data. ", "conclusions": "Stellar population synthesis models predict a strong color-$M/L$ relation. We use this correlation to model rotation curves of galaxies and determine their DM distribution. Some of the main conclusions of our investigation are:\\\\ - There are large degeneracies in the fitted NFW parameters, even with the use of stellar $M/L$ constraints.\\\\ - With the chosen color-$M/L$ normalization, adiabatically contracted halos not only over-predict central rotation speeds of dwarf galaxies, but also those of high surface brightness galaxies.\\\\ - The angular momentum distribution in the observed baryons differs from $\\Lambda$CDM DM halo predictions, suggesting that detailed angular momentum is not conserved during the galaxy assembly process.\\\\ - H{\\small I} rotation curves cannot consistently be scaled to produce the DM signature." }, "0311/astro-ph0311322_arXiv.txt": { "abstract": "The radial velocities of the 13 globular clusters in the Large Magellanic Cloud have a dispersion of 28 km s$^{-1}$ relative to the HI rotation curve of the LMC, compared to a dispersion of 30 km s$^{-1}$ with regard to the mean globular cluster velocity. This shows that, contrary to a suggestion by Schommer et al. (1992), one cannot yet rule out the possibility that the LMC globular clusters formed in a pressure supported halo, rather than in rotating disk. The globular clusters in the LMC may therefore, after all, exhibit a relationship between age and kinematics that is similar to that of the clusters in M33. ", "introduction": "In a pioneering investigation of the clusters in the Large Magellanic Cloud, Freeman, Illingworfth \\& Oemler (1983) surmised that ``[T]here is no old, kinematical halo population among the clusters of the LMC.'' At that time radial velocities were only available for nine old clusters in the Large Cloud. Presently 13 globular clusters are known to be associated with the Large Magellanic Cloud (Schommer 1991, Schommer et al. 1992, Suntzeff 1992, van den Bergh 2000, p. 104). From their discussion of the kinematics of old clusters in the LMC Schommer et al. (1992) concluded that: ``The oldest clusters still present an enigma; they do not have the kinematics of an isothermal, or slowly rotating, pressure-supported halo. These objects rotate with an amplitude comparable to that of the younger disk''. Taken at face value this conclusion would appear to indicate that the oldest clusters in the Large Cloud formed quite differently from those in M33 (Schommer et al. 1991, Chandar et al. 2002), in which the oldest (globular) clusters are observed to have halo kinematics, whereas younger (open) clusters exhibit disk-like motions. The view that late-type galaxies such as M33 (M V = -18.9) and the LMC (M V = -18.5), which have similar uminosities, had very different evolutionary histories would not fit comfortably with most current views of galaxy evolution. This prompts one to ask how secure is the conclusion that the oldest massive clusters in the Large Cloud formed in a disk, rather than in the LMC halo? ", "conclusions": "By lumping old and intermediate-age LMC clusters together Schommer et al. (1992) found that these clusters exhibited significant rotation. However, the data plotted in Figure 1 show that the oldest (SWB class VII) clusters do not exhibit a significant rotational component. In this respect the Large Cloud cluster system resembles M33 in which Chandar et al. find that 85 $\\pm$ 5 percent of the oldest clusters have halo kinematics. It is therefore concluded that worries about the apparent difference between the kinematics of the oldest population components in M33 and in the LMC (van den Bergh 2000, p. 83) were probably premature. I thank Ken Freeman for pointing out to me that the number of old LMC globular clusters might not be large enough to prove the conclusion that these clusters have disk kinematics. Thanks are also due to Peter Stetson for statistical advice. \\clearpage" }, "0311/astro-ph0311608_arXiv.txt": { "abstract": "We report infrared spectroscopy of the unusual eclipsing pre-main sequence object KH~15D, obtained using NIRSPEC on Keck II. During eclipse, observations using low spectral resolution ($\\lambda/\\delta\\lambda \\sim 1000$) reveal the presence of prominent molecular hydrogen emission in 5 lines near $2~\\mu$m. The relative line strengths are consistent with thermal excitation at $T \\sim 2800 \\pm300K$. Observations out of eclipse, at both low and high spectral resolution ($\\lambda/\\delta\\lambda \\sim 2 \\times 10^{4}$), show reduced contrast with the stellar continuum. The change in contrast for the strongest line, 1-0~S(1), is consistent with an approximately constant emission line superposed on a variable stellar continuum. Emission in the 1-0~S(1) line is observed to extend by $\\gtrsim 4$ arc-sec both east and west of the stellar point spread function ($\\gtrsim 3000$ AU). Observed at high spectral resolution, the velocity and intensity structure of the 1-0~S(1) profile are both asymmetric. East of the stellar PSF (by $1.1 - 2.3$ arc-sec) the emission is blueshifted ($-63$ km sec$^{-1}$), and has significantly greater intensity than the marginally redshifted component ($+2$ km sec$^{-1}$, $\\sim$ consistent with zero) which dominates west of the stellar PSF. The spatial extent of the emission, and the excitation temperature, suggest shock-excitation of ambient gas by a bipolar outflow from the star and/or disk. However, it is difficult to account for the observed radial velocity unless the outflow axis is inclined significantly to the plane of the sky. ", "introduction": "KH~15D is a weak-lined T Tauri star (K7V) in NGC 2264 ($d=760$ pc), which shows deep (3.5 mag) periodic eclipses, each lasting for about 20 days, a large fraction of the 48.36 day period \\citep{her02,hh03}. The eclipses occur abruptly, indistinguishable from obscuration by a `knife edge' \\citep{her02}. During eclipse, the lack of reddening in the spectrum \\citep{ham01}, and polarization which varies only weakly with wavelength \\citep{agol03}, imply that the obscuring matter must be comprised of large particles. A consensus view to date is that the eclipses are caused by structure in a protoplanetary disk, fortuitously edge-on to our line of sight. Candidate structures include a density wave \\citep{her02}, a vortex \\citep{bv03}, or a warp in the disk induced by planet-disk interactions \\citep{agol03,winn03}. The considerable recent interest in KH~15D derives from the diagnostic potential inherent in the eclipse geometry, and the serendipity of finding a protoplanetary system whose structure suggests that planetary formation is actively occurring. Although the eclipses in KH~15D are evidently caused by large solid particles, observations of the gaseous component may help to deduce the nature of the system. \\citet{ham03} observed double-peaked atomic hydrogen emission lines, indicative of a bipolar outflow. In this {\\it Letter} we report the spectroscopic discovery of $2~\\mu$m molecular hydrogen lines in emission, also showing a double-peaked structure indicative of outflow. This H$_2$ emission has been independently discovered by \\citet{tok03}, by imaging in a narrow-band filter. Molecular hydrogen emission is commonly observed from pre-main sequence objects \\citep{bach96}, where it can be excited by fluorescence \\citep{black87} or by shocks \\citep{sb82}, and has been well-studied in many cases. H$_2$ observations should therefore help to relate KH~15D to other pre-main sequence systems. ", "conclusions": "\\citet{tok03} have imaged KH~15D in a narrow band filter centered on the 1-0 S(1) line. They find a prominent jet starting in close coincidence with the star, and extending many arc-sec to the north. They have considered whether the jet is physically associated with KH~15D or is a chance superposition (H$_2$ emission is common in star-forming regions like NGC 2264). Independent of our spectroscopy, they conclude that the association of the jet with KH~15D is probably physical. We find that: 1) our low resolution spectrum during eclipse shows bright S(1) emission centered on the stellar position to $\\sim 0.2$ arc-sec (Sec. 3.1), and 2) our high resolution spectrum resolves the line into two velocity components, whose relative intensities change aburptly at the stellar position (Sec 3.2). Therefore our results strongly support the conclusion that S(1) emission is physically associated with KH~15D. Recently, quiescent H$_2$ 1-0~S(1) emission has been detected near several T-Tauri stars \\citep{wein00, bary03}. This quiescent emission shows narrow line profiles, just as we see in KH~15D. We considered the possibility that the KH~15D emission originates from quiescent gas, with Doppler shifts resulting from Keplerian orbital motion. We cannot exclude that some fraction of the emission, especially in the red component, arises from quiescent emission. Nevertheless, because the KH~15D emission is spatially extended and shows evidence of outflow, it seems more closely related to the shock-heated H$_2$ which has been observed in hundreds of young stellar objects, including classical T-Tauri stars \\citep{bach96, her97}. Interpretation of our KH~15D observations requires knowing the stellar radial velocity, and the heliocentric radial velocity we adopt for the system ($+10$ km sec$^{-1}$) may be in error if the star is a binary. Nevertheless, the double-peaked profiles we observe in the close vicinity of the star suggest outflow. However, a simple picture of a symmetric bipolar jet emergent normal to a single edge-on disk is hard to reconcile with our observations, and with the \\citet{tok03} imaging. In such a simple interpretation, the line-of-sight shock velocity in the 1-0 S(1) line would have to be at least half of the difference between the red- and blueshifted components, i.e. $\\gtrsim 30$ km sec$^{-1}$. For a disk inclination of $\\sim 84^{\\circ}$ \\citep{ham03}, the flow would be inclined by only $\\sim 6^{\\circ}$ to the plane of the sky, so the total shock velocity would have to exceed $\\sim 250$ km sec$^{-1}$. This is well above the velocity where shocked H$_2$ is observed in other objects (H$_2$ becomes dissociated at high shock velocities, \\citet{sb82}). Therefore the outflow revealed in molecular hydrogen is likely to be inclined significantly to the plane of the sky. Similar arguments would not apply to lines formed very close to the star, where a bipolar outflow may not be collimated. However, the emission we observe is spatially resolved from the star, occurring at distances where bipolar outflows are often highly collimated to a single axis. \\citet{tok03} find only one bright jet of line emission extending to the north. Since our slit was oriented primarily east-west (PA $=66^{\\circ}$), we have observed approximately (but not precisely) perpendicular to the jet axis. Our spectra will reveal the flow pattern in the jet to the extent that the total angle between our slit and the velocity vector differs from ${\\pi}/2$. We believe that our brightest emission component (blueshifted `east' spectrum in Figure~2) samples the \\citet{tok03} jet, whereas the red components are consistent with either a possible fainter counter-jet (not yet detected by imaging), or quiescent emission near the stellar velocity. The extent and complexity of H$_2$ emission from KH~15D is reminiscent of T~Tauri itself, a known binary system. Consistent with its binary nature, T~Tauri shows two nearly perpendicular outflow systems \\citep{her97}. Our observations are not yet sufficient, in terms of spatial sampling, to say whether KH~15D shows a single, or multiple, velocity systems. Mapping spectroscopy of KH~15D in the 1-0~S(1) line is needed to clarify the kinematics of this interesting object." }, "0311/astro-ph0311052_arXiv.txt": { "abstract": "We describe the first results from two observational projects aimed at measuring the amount and spatial distribution of dark matter in distant early-type galaxies (E/S0s) and clusters of galaxies. At the galaxy scale, the Lenses Structure and Dynamics (LSD) Survey is gathering kinematic data for distant (up to $z\\sim1$) E/S0s that are gravitational lenses. A joint lensing and dynamical analysis constrains the fraction of dark matter within the Einstein radius, the mass-to-light ratio of the stellar component, and the total slope of the mass density profile. These properties and their evolution with redshift are briefly discussed in terms of the formation and evolution of E/S0 galaxies and measurement of the Hubble Constant from gravitational time delay systems. At the cluster scale -- after careful removal of the stellar component with a joint lensing and dynamical analysis -- systems with giant radial arcs can be used to measure precisely the inner slope of the dark matter halo. An HST search for radial arcs and the analysis of a first sample are briefly discussed in terms of the universal dark matter halos predicted by CDM simulations. ", "introduction": "Decades after the discovery of dark matter around spiral galaxies little is known about dark matter in early-type galaxies. Mass tracers at large radii (such as stellar kinematics, kinematics of globular clusters and of planetary nebulae, and X-ray halos) generally indicate that a constant mass-to-light ratio cannot reproduce the observations, although there are typically large uncertainties and a wide variety of behavior is seen. The main source of uncertainty in interpreting kinematic measures is that the derived mass profile depends on the assumed orbital structure. This is commonly referred to as the mass-anisotropy degeneracy. The situation is even more uncertain outside the local Universe, because traditional kinematic tracers at large radii are generally not feasible at cosmological distances. However, additional constraints can be gathered by looking at early-type galaxies that are gravitational lenses. The configuration of multiple images provides information on the mass distribution of the lens; first and foremost a robust measurement of the mass enclosed within the Einstein Radius (typically larger than the effective radius), {\\it independent} of its kinematic status. The Lenses Structure and Dynamics (LSD) Survey (Koopmans \\& Treu 2002, 2003; Treu \\& Koopmans 2002a,2003; hereafter KT) is obtaining internal kinematics for a sample of of 11 relatively isolated E/S0 lens galaxies at $z=0.04-1.01$. A joint lensing and dynamical analysis is used to break the mass-anisotropy degeneracy and determine the amount and distribution of luminous and dark matter. The first results from this project are briefly summarized and discussed in Section~2. At the cluster scale, the presence of dark matter has been known for a long time (Zwicky 1937). The distribution of dark matter has been probed over a wide range of scales using X-ray analysis, dynamical studies and gravitational lensing, showing that dark matter dominates in mass. The wealth of mass tracers in clusters -- in particular the existence of giant radial arcs -- makes them an ideal laboratory to test the existence of universal mass density profiles predicted by cold dark matter (CDM) simulations. Combining a lensing analysis of clusters with giant arcs with a dynamical analysis of the kinematics of the brightest cluster galaxy (BCG), we can measure very accurately the mass density profile within $\\sim100$ kpc, disentangle the stellar and dark matter, and pinpoint the inner logarithmic slope of the dark matter halo. This has motivated a search of suitable systems with a central dominant BCG and giant arcs through the entire HST-WFPC2 archive (Sand et al.\\ 2004, in preparation), and the spectroscopic follow-up at Keck (Sand, Treu \\& Ellis 2002; Sand et al.\\ 2003). The first results from this project are briefly summarized and discussed in Section 3. ", "conclusions": "" }, "0311/astro-ph0311578_arXiv.txt": { "abstract": "We present the high resolution X-ray spectrum of the BL Lac object PKS~2155-304 taken with the RGS units onboard {\\em XMM-Newton} in November 2000. We detect a O~{\\sc vii} K$\\alpha$ resonant absorption line from warm/hot local gas at 21.59 \\AA \\/ ($\\sim 4.5$~$\\sigma$ detection). The line profile is possibly double peaked. We do not confirm the strong 20.02 \\AA \\/ absorption line seen with {\\em Chandra} and interpreted as $z \\sim 0.05$ O{\\sc viii} K$\\alpha$. A $3$~$\\sigma$ upper limit of 14 m\\AA \\/ on the equivalent width is set. We also detect the $\\sim 23.5$~\\AA \\/ interstellar O{\\sc i} 1s$\\rightarrow$2p line and derive a factor $\\leq 1.5$ subsolar O/H ratio in the ISM along \\pks \\/ line of sight. ", "introduction": "The observed baryon density at $z > 2$ \\citep[e.g.][]{rau98,wei97} agrees well with Standard Big-Bang nucleosynthesis predictions, when combined with observed light-element ratios \\citep{bur98}. At lower redshift, however, less than $1/3$ of the baryons observed at $z > 2$ have been detected so far \\citep{fuk98}. According to simulations for the formation of structures in the Universe, most of such baryons would be located, at the present epoch, in low density intergalactic gas, which has been shock-heated to temperatures of $\\sim 10^{5} - 10^{7}$~K \\citep[i.e. the warm-hot intergalactic medium WHIM, e.g.][]{hel98}. The most efficient way to detect the presence of WHIM is through resonant absorption lines from highly ionized metals (e.g. O{\\sc vi}, O{\\sc vii}, O{\\sc viii}, Ne~{\\sc ix}) imprinted in the far UV (FUV) and soft X-ray spectra of background sources \\citep{ald94, mul96, hel98, per98, fan00}. However, severe instrumental limitations have hampered so far the detection of conspicuous amount of WHIM. The Far Ultraviolet Spectroscopic Explorer ({\\em FUSE}) allows the OVI doublet ($\\lambda = 1031.93$ and $\\lambda=1037.62$ \\AA ) to be detected only up to z$\\sim 0.2$. More importantly OVI dominates the relative abundance distribution of O in shock-heated gas in pure collisional ionization equilibrium, only at relatively low temperatures, $T\\sim (1-5) \\times 10^{5}$~K, and so it tracks just the low-temperature tail of the WHIM distribution \\citep[e.g.][]{fan00}. Soft X-rays are far more promising. The strongest resonant transitions from O{\\sc vii}, Ne{\\sc ix} (both helium-like ions, and therefore highly stable) and O{\\sc viii}, all fall in the soft X-ray band. These ions dominate the O and Ne relative abundance distributions of both collisionaly ionized gas, and mixed, photoionized and collisionaly ionized gas, over a broad range of temperatures, between $T \\sim 5\\times 10^5$ K and $T \\sim 10^7$ K, where the WHIM distribution peaks \\citep{dav01,fan00}. The major fraction of the WHIM should then be visible in the X-rays. However, despite the large relative abundances of ``X-ray'' ions in the WHIM, the current sensitivity and resolution of X-ray spectrometers has allowed so far only the strongest (EW$\\geq 10$ m\\AA) of these systems to be detected, and only against spectra of very bright background sources (Nicastro et al. 2002; Mathur, Weimberg \\& Chen 2003; Fang et al. 2002; Fang, Weimberg \\& Canizares 2003; with {\\em Chandra} and Rasmussen, Kahn \\& Paerels 2003 with {\\em XMM-Newton}; for a recent review see Paerels \\& Kahn 2003). Three out of these six detections have been made against the spectrum of the bright ($F_{\\rm 2-10 keV} \\simeq 2 \\times 10^{-11} - 5 \\times 10^{-10}$ erg cm$^{-2}$ s$^{-1}$), nearby \\citep[z=0.116]{fal93}, blazar PKS~2155-304 \\citep{nic02, fan02}. A high quality ($\\geq 700$ counts per resolution element at the relevant wavelengths) {\\em Chandra} (HRCS/LETG) observation of PKS~2155-304, revealed the existence of O{\\sc vii}, O{\\sc viii} and Ne{\\sc ix} absorption lines at $z \\simeq 0$, identified with a WHIM system, pervading our Local Group \\citep{nic02}. Lower quality {\\em Chandra} (ACIS/LETG) spectra of PKS~2155-304 confirmed the above findings \\citep{fan02}, and also show a line at 20.02 \\AA, that \\citet{fan02} identify as the O{\\sc viii} WHIM counterpart of a known Ly$\\alpha$ and O{\\sc vi} system at $z \\sim 0.05$, where a concentration of galaxies is seen \\citep[][and references therein]{mar88,shu98,shu03}. However it is not clear why such a feature is not detected in the higher quality HRCS/LETG spectrum published by \\citet{nic02}. Additional data are needed to clarify this issue. In this paper we present the analysis of the high resolution {\\em XMM-Newton} Reflection Grating Spectrometers (RGS hereinafter) spectra of \\pks .\\\\ The structure of the paper is as follows: in \\S~\\ref{sec_obs} we report on the RGS data reduction and analysis. The spectral fitting and comparison with previous measurements is given in \\S~\\ref{sec_fits}. The discussion of the detected features is presented in \\S~\\ref{sec_disc}. ", "conclusions": "\\label{sec_disc} At $z \\sim 0$ we confirm the existence of a O{\\sc vii} K$\\alpha$ absorption feature, but we do not detect features of higher ionization elements (O{\\sc viii}, Ne{\\sc ix} and Ne{\\sc x}) seen by {\\em Chandra} \\citep{nic02}. This lack of detection is likely to be related to the limited number of photons (e.g. $\\sim 15300$ photons between 19.5 and 20.5 \\AA \\/, i.e. $\\sim 100$ photons per 0.0065 \\AA \\/ resolution element) collected in the {\\em XMM-Newton} observation, when the source was $\\sim 2-3$ times fainter than in \\citet{nic02}. The O{\\sc vii} K$\\alpha$ appears to be unsaturated and the curve of growth presented in \\citep{nic02} suggests an O{\\sc vii} column density of $\\sim 10^{16}$ cm$^{-2}$. If the weak evidence of a double peak in the profile of the O{\\sc vii} K$\\alpha$ line is considered real, such shape may indicate that we are starting to detect the different sheets of the local filament. The $3$~$\\sigma$ upper limit on the O{\\sc viii} Ly$\\alpha$ line EW, computed fixing the line FWHM to the O{\\sc vii} K$\\alpha$ FWHM value, is $12.61$ m\\AA , consistent with the value of $9.0^{+2.6}_{-2.7}$ m\\AA \\/ found by \\citet{ras03} in the RGS spectrum of \\pks . Under the assumption of unsaturated lines the EW ratios between different ions of the same element depend on the gas temperature and density. Using the ratio between the upper limit on the EW of O~{\\sc viii} Ly$\\alpha$ to the O{\\sc vii} K$\\alpha$ (best fit value $-2$~$\\sigma$), we obtain an upper limit on the gas temperature of $\\sim 2.5-3.5 \\times 10^6$~K, for a gas density of $10^{-6}$- 1 atom cm$^{-3}$ \\citep[see Fig.~5 in][]{nic02}. These values are consistent with the temperature range predicted for the WHIM. It is not possible to discriminate with the present data whether the O{\\sc vii} features are produced by a WHIM, as proposed by \\citet{nic02} or by radiatively cooling gas within our Galaxy, as proposes by \\citet{hec02}. In the {\\em FUSE} spectrum of \\pks \\/ \\citet{nic02} found two unsaturated O{\\sc vi} absorption lines: a narrow component (FWHM = $106 \\pm 9$ km s$^{-1}$) at $cz = 36 \\pm 6$ km s$^{-1}$ with EW = $(2.1 \\pm 0.2) \\times 10^{-3}$~eV, probably related to a cloud in the Galactic disk, and a broad component (FWHM = $158 \\pm 26$ km s$^{-1}$) at $cz = -135 \\pm 14$ km s$^{-1}$ with EW = $(1.6 \\pm 0.4) \\times 10^{-3}$~eV possibly associated with a WHIM filament. Using our O{\\sc vii} K$\\alpha$ EW and the EW of the O{\\sc vi} broad component from \\citet{nic02} we derive a temperature range of $\\sim (4.5 -200) \\times 10^{5}$~K for a gas density of $\\sim 10^{-6}$ cm$^{-3}$, which restricts to $\\sim (4.5 - 25) \\times 10^{5}$~K once the upper limit from the O{\\sc viii} is considered. We note, however, that the O{\\sc vii} absorption features at local redshift have been found in the spectra of other 4 AGNs (3C~273 \\citet{fan03}; MRK~421 \\citet{nic01,cag03}; NGC~4593 \\citet{mck03} and NGC~3783 \\citet{kas02}) and might be associated to the high velocity O{\\sc vi} lines (sampling a gas at $T\\sim 10^5-10^6$ K) recently seen in the spectra of bright AGNs by FUSE \\citep{sem02, nic02b}. If this is the case, the ``local'' O{\\sc vii} K$\\alpha$ would indicate the presence of WHIM within the Local Group \\citep{nic02b}. The line at 23.5 \\AA \\/ is the interstellar O{\\sc i} 1s$\\rightarrow$2p transition \\citep[e.g.][and references therein]{mcl98}. Assuming a neutral ISM gas, no consistent velocity broadening and unsaturated lines, the comparison of the measured EW with the curve of growth derived using \\citet{sto97} cross sections implies $\\sim 2 \\times 10^{16}$ O atoms cm$^{-2}$. The H absorbing column density in \\pks \\/ direction is $1.36 \\times 10^{20}$ cm$^{-2}$ \\citep{loc95}, which brings to a O/H ratio of $1.4 \\times 10^{-4}$. This value is $\\sim 6$ times smaller than in the Solar system \\citep[$8.5 \\times 10^{-4}$][]{and89}, but consistent with the ISM elemental abundances in the Magellanic Clouds \\citep[e.g.][]{rus90}. Assuming that the line EW is enhanced by velocity broadening would further decrease the O/H ratio. If we adopt the theoretical cross-sections of \\citep{mcl98}, as in \\citet[][(see their Fig.~3)]{dev03}, we derive $\\sim 7 \\times 10^{16}$ O atoms cm$^{-2}$, i.e. a O/H ratio of $\\sim 5 \\times 10^{-4}$, still $\\sim 1.5$ times lower than the Solar value. We note that \\citet{dev03} O{\\sc i} 1s$\\rightarrow$2p EW in \\pks \\/ direction, derived from a coaddition of RGS observations for a total of 346~ks, is $15 \\pm 3$ m\\AA \\/ consistent with our measurement ($17.57^{+1.21}_{-5.81}$ m\\AA ) and the line curve of growth presented in the paper takes saturation effects into account, but still there is indication of a subsolar O/H ratio. We do not confirm the O{\\sc viii} K$\\alpha$ line by \\citet{fan02}. Note that the lack of detection of WHIM outside our Local Group is consistent with the expectations of theoretical models. In fact only one O{\\sc viii} absorption line with EW$>3$ m\\AA \\/ (i.e. detectable by the present X-ray satellites) in the spectrum of a random background source at $z\\sim 0.3$ is predicted \\citep[e.g.][]{hel98}. The non-detection of WHIM up to $z =0.116$ therefore does not pose any problem. Our {\\em XMM-Newton} \\pks \\/ spectrum shows a hint ($1.4$~$\\sigma$ detection; EW $\\sim 5.56$~m\\AA ) of a 22.7 \\AA \\/ feature, seen by \\citet{dev03} in the {\\em XMM-Newton} spectra of \\pks \\/ and MRK~421, and in the {\\em Chandra} spectrum of \\pks . Since they do not find evidence of such line in the {\\em XMM-Newton} spectra of Sco~X-1 and 4U~0614$+$091, they exclude a possible instrumental origin of the line and tentatively identify it with O{\\sc iv} absorption from a local WHIM filament. Higher statistics and other time spaced observations are needed to investigate the reality of the 22.7 \\AA \\/ absorption feature." }, "0311/astro-ph0311264_arXiv.txt": { "abstract": "We present the results of a photometric and spectroscopic survey of the globular cluster system of \\cena\\ (Centaurus~A), a galaxy whose proximity makes it an important target for early-type galaxy studies. We imaged three fields in {\\it UBVRI} that extend 50 and 30~kpc along the major and minor axes, respectively. We used both color and size information to develop efficient selection criteria for differentiating between star clusters and foreground stars. In total, we obtained new velocities for 138 globular clusters, nearly tripling the number of known clusters, and bringing the confirmed total in \\cena\\ to 215. We present a full catalog of all known GCs, with their positions, photometry, and velocities. In addition, we present catalogs of other objects observed, such as foreground stars, background galaxies, three Galactic white dwarfs, seven background QSOs, and 52 optical counterparts to known X-ray point sources. We also report an observation of the cluster G169, in which we confirm the existence of a bright emission line object. This object, however, is unlikely to be a planetary nebula, but may be a supernova remnant. ", "introduction": "Systems of globular clusters are a nearly ubiquitous feature of all nearby galaxies. Globular clusters (GCs), which are typically old with sub-solar metallicities, are the most visible remnants of intense star formation that occurred in a galaxy's distant past. As single-age, single-metallicity stellar populations, GCs are ideal for studying the fossil remains of a galaxy's star formation and metal-enrichment history. The past decade has seen a rapid growth in the study of extragalactic GC systems. The availability of the Hubble Space Telescope (HST) in particular has enabled the study of GC systems in galaxies well beyond the Local Group. One of the more striking results of these studies is the frequency with which the metallicity distributions of these GC systems are bimodal (e.g. Larsen \\etal 2001; Kundu \\& Whitmore 2001). Different scenarios of galaxy formation (or at least of globular cluster system formation) have been proposed or adapted to explain the observed metallicity distributions and other properties of GC systems: mergers of spiral galaxies (Ashman \\& Zepf 1992), multiple {\\it in situ} star formation epochs (Forbes, Brodie, \\& Grillmair 1997), dissipationless hierarchical merging of protogalactic clumps (C\\^{o}t\\'{e}, Marzke, \\& West 1998), and more generalized hierarchical merging (Beasley \\etal 2002). The question that all of these scenarios address largely concerns the nature and time frame of merging or gas dissipation. Elements of all of these scenarios are supported by various studies of different facets of galaxy formation and evolution. For example, the supernova winds that some use to explain the suppression of star formation in dwarf galaxies at high redshift (Dekel \\& Silk 1986) has also been proposed as a possible mechanism for the truncation of metal-poor GC formation at early times (Beasley \\etal 2002). Similarly, hierarchical merging models are often used to explain the mass assembly of present-day galaxies. While most spheroid formation may have occurred in the past (at redshifts $z > 2$), present-day examples of recent merger remnants may give us a window onto this distant epoch. Locally, there is compelling evidence that some ellipticals have recently interacted or merged with another galaxy. Many ellipticals possess large-scale disks of gas and dust. Faint structure in the form of loops, shells, ripples, and tails are especially visible in the outer regions of ellipticals (Malin \\& Carter 1983), and are presumably the aftermaths of a recent interaction. One such galaxy, and perhaps the best candidate for study, is the nearby elliptical, \\cena. We have chosen to examine of the stellar content of \\cena\\ in order to further elucidate its formation history and gain insight on the formation of other ellipticals. Our survey of the planetary nebula (PN) system in \\cena's outer halo represents the kinematics of the field star population, and is presented in a separate paper (Peng, Ford, \\& Freeman 2004). In this paper, we describe our photometric and spectroscopic survey for globular clusters, and present the resulting catalogs of objects. \\vspace{0.3cm} ", "conclusions": "We conducted an optical imaging and spectroscopic survey for globular clusters across $\\sim 1\\degr$ of sky around \\cena. Using \\ubvri\\ photometry, size, and morphological information, we developed an efficient algorithm for selecting likely GCs candidates for spectroscopic follow-up---a necessity given the large number of foreground stars. We obtained radial velocities for over 400 objects. Of these, we identified 102 previously unknown GCs, confirmed the nature of 24 GCs from Rejkuba (2001), 6 GCs from HCH99, 4 GCs from MAGJM96, and 2 from HHH86, providing 138 new GC velocities. We also obtained new velocities for 25 previously confirmed GCs from HGHH92 and HHH86. The total number of confirmed GCs in \\cena\\ is now 215. We present a spectrum of HGHH-G169, which shows an interesting split emission line feature. It is unlikely to be a planetary nebula, as MR02 suggest, but it could be a supernova remnant, and certainly merits a deeper spectrum." }, "0311/astro-ph0311028_arXiv.txt": { "abstract": "In response to the claim by Dziembowski {\\em et al.}\\ (2001) that the solar radius decreases with magnetic activity at the rate of 1.5 km yr$^{-1}$, we consider the theoretical question whether a radius variation is expected with the solar cycle. If the radius variation is caused by the magnetic pressure of toroidal flux tubes at the bottom of the convection zone, then the dynamo model of Nandy and Choudhuri (2002) would suggest a radius decrease with magnetic activity, in contrast to other dynamo models which would suggest a radius increase. However, the radius decrease is estimated to be only of the order of hundreds of metres. ", "introduction": "By analyzing the f-mode frequency data from MDI for the period 1996--2001, Dziembowski {\\em et al.}\\ (2001) claimed that the Sun shrinks in radius from the solar minimum to the solar maximum, at the rate of about 1.5 km per year. If correct, then this will certainly be a most dramatic discovery in solar physics. However, this claim has not yet been supported by other researchers, who often present conflicting results. Antia (2003) provides a review of the subject and concludes that there is no compelling evidence for solar radius change. It may be noted that some earlier studies (N\\\"oel, 1997; Emilio {\\em et al.},\\ 2000) reported much larger radius variation (up to 700 km) based on other (non-helioseismic) methods. It is clear that more work is needed before a definitive conclusion can be reached. The aim of this paper is to discuss whether on theoretical grounds we expect a solar radius variation with solar cycle. Other than some early work by Spruit (1982, 1994), we have not been able to find any other theoretical analysis of this problem. In the light of recent developments of solar physics, this theoretical question certainly needs revisiting. There are two ways in which a magnetic field present in the solar interior may lead to an increase of solar radius. \\begin{enumerate} \\item The pressure of the magnetic field embedded in the gas may cause the gas to expand, thereby inflating the size of the Sun. \\item The inhibition of convection by the magnetic field may cause a piling up of heat below the magnetic field. This would result in heating of the surrounding regions, leading to an expansion. \\end{enumerate} From both of these arguments, it appears that the radius of the Sun should increase when more magnetic fields are present. Since we normally think of the solar maximum as being the time when the Sun has more magnetic fields in the interior, we would expect the Sun to be larger in size during the solar maximum. Shrinking of the Sun with increasing magnetic activity, therefore, would seem improbable at the first sight. We shall argue that the recent solar dynamo model of Nandy and Choudhuri (2002), in principle, allows for the possibility of shrinking of the Sun during the solar maximum. The important theoretical question is whether the theoretically expected radius decrease agrees with the claim of Dziembowski {\\em et al.}\\ (2001). Spruit (1982, 1994) studied the effect of heat blocking by magnetic fields and concluded that the total radius variation cannot be more than 0.14 km. We mainly look at the other effect of magnetic pressure and present order-of-magnitude estimates, suggesting that the radius variation, for reasonable assumptions, cannot be as large as what Dziembowski {\\em et al.}\\ (2001). We have also made a mathematical formulation of the problem, to go beyond order-of-magnitude estimates. To our surprise, we found that the mathematical problem is much more difficult to solve than what we anticipated. Although our mathematical analysis so far has not yielded very definite results, we present a brief account of it to help future researchers who may wish to analyze this problem in greater depth. Section 2 addresses the question whether we expect the radius to increase or decrease during the solar maximum. An order-of-magnitude estimate is presented in Section~3, showing that a much smaller variation of radius is expected than what is reported by Dziembowski {\\em et al.} \\ (2001). Our incomplete mathematical analysis is given in Section~4. Finally, our conclusions are summarized in Section~5. ", "conclusions": "Within the currently understood framework of solar MHD, we have not been able to provide a theoretical explanation of the controversial claim made by Dziembowski {\\em et al.}\\ (2001) that the Sun shrinks with increasing magnetic activity at the rate of 1.5 km yr$^{-1}$. Spruit (1982, 1994) estimated the radius change due to heat blocking by sunspots near the solar surface. Not only the radius variation was found to be much smaller, the radius is expected to increase with the solar maximum. We considered the possibility if the excess magnetic pressure of flux tubes near the base of the SCZ can give us a different result. Even though traditional dynamo models regard the solar maximum as the time when the magnetic flux in the solar interior peaks and would suggest a radius increase at the time of the solar maximum, we pointed out the model of Nandy and Choudhuri (2002) provides a different scenario. Although this model would predict a decrease of the solar radius with increasing magnetic activity (i.e.\\ with increasing loss of magnetic flux from the storage region), various reasonable assumptions give a radius decrease rate about one order of magnitude smaller than what is claimed by Dziembowski {\\em et al}.\\ (2001). It is true that our estimate is based on a very rough calculation. However, we believe that it is still an over-estimate rather than an under-estimate. Only if the magnetic field at the base of the SCZ was concentrated to values as high as $10^6$ G, the radius shrinkage would have been as large as reported by Dziembowski {\\em et al.}\\ (2001). There is no theoretical reason to expect a magnetic field as strong as $10^6$ G at the base of the SCZ. In fact, it is not easy to generate even a $10^5$ G magnetic field there (Choudhuri, 2003). We conclude that either the claim of Dziembowski {\\em et al.}\\ (2001) is incorrect, or else we do not understand some basic physics of the solar cycle." }, "0311/astro-ph0311502_arXiv.txt": { "abstract": "We report further results from a 191~ks \\emph{Chandra} observation of the core of the Perseus cluster, Abell 426. The emission-weighted temperature and abundance structure is mapped detail. There are temperature variations down to $\\sim 1 \\kpc$ in the brightest regions. Globally, the strongest X-ray surface brightness features appear to be caused by temperature changes. Density and temperature changes conspire to give approximate azimuthal balance in pressure showing that the gas is in hydrostatic equilibrium. Si, S, Ar, Ca, Fe and Ni abundance profiles rise inward from about 100~kpc, peaking at about 30--40~kpc. Most of these abundances drop inwards of the peak, but Ne shows a central peak, all of which may be explained by resonance scattering. There is no evidence for a widespread additional cooler temperature component in the cluster with a temperature greater than a factor of two from the local temperature. There is however evidence for a widespread hard component which may be nonthermal. The temperature and abundance of gas in the cluster is observed to be correlated in a manner similar to that found between clusters. ", "introduction": "The Perseus cluster, Abell 426, is the brightest cluster of galaxies in the sky in the X-ray band. The most prominent features in the core of the cluster are the cavities in the X-ray emitting gas (B\\\"ohringer et al 1993), associated with the bubbles blown by the central source 3C~84 (Pedlar et al 1990). The coolest X-ray gas in the cluster lies on the rims of these lobes (Fabian et al 2000, 2003a). A swirl of cool gas occupies the inner 70 kpc radius of the cluster (Churazov et al 2000; Fabian et al 2000; Schmidt, Fabian \\& Sanders 2002; Churazov et al 2003a; Fabian et al 2003a), winding around the core in an anti-clockwise direction. The X-ray emission contains a number of quasi-periodic fluctuations in emission, which are interpreted as sound waves propagating in the intracluster medium (ICM), driven by the active nucleus (Fabian et al 2003a). In addition there appears to be a weak shock to the NE of the core, about 10 kpc from the bubble rim. The abundance in the cluster rises towards the centre, peaking around $0.7\\Zsun$ about 40-50 kpc from the core, and dropping back down in the centre (Schmidt et al 2002; Churazov et al 2003a). We previously observed the Perseus cluster with \\emph{Chandra} (Fabian et al 2000; Schmidt et al 2002) for about~25 ks. The initial results from a much longer 191.2~ks \\emph{Chandra} observation were reported in Fabian et al (2003a), showing evidence for the presence of shocks and ripples in the core of the cluster. Here we present further results from this observation, in particular the temperature and abundance structures in this object, and tests for the presence of multiphase gas. As this is the deepest \\emph{Chandra} observation of a bright nearby diffuse object, our results are affected by systematic uncertainties which are not visible in shorter observations. We do not discuss the high-velocity system (HVS) in depth here (see Gillmon, Sanders \\& Fabian 2003), nor the relationship between the H$\\alpha$ nebulosity and the X-ray emission (see Fabian et al 2003b). The Perseus cluster is at a redshift of 0.0183. We assume that $H_0 = 70 \\kmpspMpc$; therefore 1 kpc corresponds to about 2.7 arcsec. ", "conclusions": "We have studied a 191~ks \\emph{Chandra} image of the core of the Perseus cluster for small-scale temperature and abundance structures. The obvious X-ray surface brightness features are seen to be due to temperature variations, with corresponding density variations ensuring approximate azimuthal pressure balance. Such variations are seen on scales down to above 2.5~arcsec ($\\sim1$~kpc) in the brightest regions. We have made extensive searches for multi-temperature gas components. There is little evidence in the deprojected spectra for any widespread components at a temperature more than a factor of two away from the local temperature. There is some evidence for a widespread hot component (at 16~keV) in the projected data. This emission could alternatively be due to an extended nonthermal component. The solution to the cooling flow problem in the Perseus cluster must produce a fairly uniform mass distribution of gas from 2.5 to 8~keV, with little gas below~2 keV. The abundances of Si, S, Ar, Ca, Fe and Ni rise inward from about 100~kpc, peaking at about 30--40~kpc. Most of these abundances level out or drop again inward of the peak. The extent of the drop is unclear, but it is plausibly explained by resonance scattering. O and Mg are more uniform across the core and Ne shows a central peak. The overall abundance pattern is similar to that found by others using XMM-Newton data of M87 in the Virgo cluster, except that S/Si exceeds unity in the Perseus cluster core. The abundance peaks are likely due to Type Ia supernovae in the central galaxy, NGC\\,1275." }, "0311/astro-ph0311444_arXiv.txt": { "abstract": "The Extended Self-Similarity (ESS) of cosmic microwave background (CMB) radiation has been studied using recent data obtained by the space-craft based Wilkinson Microwave Anisotropy Probe. Using the ESS and the high angular scale resolution (arcminutes) of the data it is shown that the CMB temperature space {\\it increments} exhibit considerable and systematic declination from Gaussianity for high order moments at the small angular scales. Moreover, the CMB space increment ESS exponents have remarkably close values to the ESS exponents observed in turbulence (in magnetohydrodynamic turbulence). ", "introduction": " ", "conclusions": "" }, "0311/astro-ph0311391_arXiv.txt": { "abstract": "At the end of their lives low mass stars such as our Sun lose most of their mass. The resulting planetary nebulae show a wide variety of shapes, from spherical to highly bipolar. According to the generalized interacting stellar winds model, these shapes are due to an interaction between a very fast tenuous outflow, and a denser environment left over from an earlier slow phase of mass loss. Previous analytical and numerical work shows that this mechanism can explain cylindrically symmetric nebulae very well. However, many circumstellar nebulae have a multipolar or point-symmetric shape. With two-dimensional calculations, Icke showed that these seemingly enigmatic forms can be easily reproduced by a two-wind model in which the confining disk is warped, as is expected to occur in irradiated disks. Here, we present the extension to fully three-dimensional adaptive mesh refinement simulations of such an interaction. ", "introduction": "In the final phases of stellar evolution, low mass stars, such as our Sun, first swell up and shed a dense, cool wind in the asymptotic giant branch (AGB) phase. This episode is followed by a fast, tenuous wind that is driven by the exposed stellar core, the future white dwarf. The planetary nebulae (PNe) resulting from this expulsion phase come in a wide variety of shapes, from spherical to highly bipolar. Some even have a multipolar or point-symmetric shape. Balick \\citep{1987AJ.....94..671B} proposed that such forms arise due to an interaction between a slow disk-shaped inner AGB nebula and the fast `last gasp' of the star. Analytical \\citep{1988A&A...202..177I, 1989AJ.....97..462I} and numerical \\citep{1989ApJ...339..268S, 1991A&A...251..369I, 1991A&A...252..718M} work shows that this {\\it generalized interacting stellar winds} (GISW) mechanism works very well (for an up-to-date review, see \\citet{2002ARA&A..40..439B}). Several scenarios for obtaining a disk around a PN exist, and it is in general assumed in the models that the shape of the dense gas around the star is a disk or a toroid. Icke \\citep{2003A&A...405L..11I} proposed that the point-symmetric shapes observed in a number of PNe are formed in an interaction between a spherical stellar wind and a {\\it warped} disk. It is prossible to produce such a warp around a single star, through the combined effects of irradiation and cooling \\citep[e.g.][]{1996MNRAS.281..357P, 1996ApJ...472..582M}. Whereas Icke's computations were restricted to a two-dimensional proof-of-principle, we now present a first series of fully three-dimensional hydrodynamic computations of such a wind-disk interaction using the technique of adaptive mesh refinement (AMR). ", "conclusions": "Our computations show that the wind-disk interaction model in which the confining disk is warped results in a wide variety of point-symmetric shapes. Nebulae that show 'punched holes', such as NGC~7027 \\citep{2002A&A...384..603C}, are readily accommodated in this model. Other candidates for our model are probably NGC~6537 and, to a lesser extent, NGC~7026, in which the inner disk is still visible, while the outer nebula shows clear point-symmetric structures. Also, the shapes of the (proto-)PNe He~2-155 and M~4-18 (and maybe He~2-47) can be explained with our model. As a further application, large-scale explosions in non-planar disks, such as might occur in active galaxies, are expected to show similar patterns, provided that the disk material can cool rapidly enough. Movies of these simulations can be found at \\url{http://www.strw.leidenuniv.nl/AstroHydro3D/}. \\subparagraph{Acknowledgements} V.I. expresses his gratitude to Raghvendra Sahai and Hugo Schwarz for lively discussions that were the primary cause for taking up this subject. The research of G.M. has been made possible by a fellowship of the Royal Netherlands Academy of Arts and Sciences. The software used in this work was in part developed by the DOE-supported ASCI/Alliance Center for Astrophysical Thermonuclear Flashes at the University of Chicago. Our work was sponsored by the National Computing Foundation (NCF) for the use of supercomputer facilities, with financial support from the Netherlands Organization for Scientific Research (NWO), under grant number $614.021.016$." }, "0311/hep-ph0311171_arXiv.txt": { "abstract": "{ Within the Quantum Field Theory context the idea of a ``cosmological constant'' (CC) evolving with time looks quite natural as it just reflects the change of the vacuum energy with the typical energy of the universe. In the particular frame of Ref. \\cite{Letter}, a ``running CC'' at low energies may arise from generic quantum effects near the Planck scale, $M_P$, provided there is a smooth decoupling of all massive particles below $M_P$. In this work we further develop the cosmological consequences of a ``running CC'' by addressing the accelerated evolution of the universe within that model. The rate of change of the CC stays slow, without fine-tuning, and is comparable to $H^2\\,M_P^2$. It can be described by a single parameter, $\\nu$, that can be determined from already planned experiments using SNe Ia at high z. The range of allowed values for $\\nu$ follows mainly from nucleosynthesis restrictions. Present samples of SNe Ia can not yet distinguish between a ``constant'' CC or a ``running'' one. The numerical simulations presented in this work show that SNAP can probe the predicted variation of the CC either ruling out this idea or confirming the evolution hereafter expected.} ", "introduction": "The Standard Cosmological Model fits our universe, in the large, into an homogeneous and isotropic Friedmann-Lema\\^{\\i}tre-Robertson-Walker (FLRW) cosmological type\\, \\cite{Peebles}. Its 4-curvature is determined from the various contributions to its total energy density, namely in the form of matter, radiation and cosmological constant. Evidence for a dominant content of energy in the form of cosmological constant was found by tracing the rate of expansion of the universe along z with high--z Type Ia supernovae \\cite{p99, Riess98}. This measurement combined with the measurements of the total energy density $\\OT^0$ from the CMB anisotropies \\cite{CMBR,WMAP}, indicates that $\\OL^0\\sim 70\\%$ of the critical energy density of the universe is cosmological constant (CC) or a dark energy candidate with a similar dynamical impact on the evolution of the expansion of the universe. The matter content, on the other hand is dominated by the dark matter, whose existence is detected by dynamical means \\cite{Peebles}, and amounts to $\\OM^0\\sim 30\\%$ of the critical density. The CC value found from Type Ia supernovae at high z \\cite{p99,Riess98} is: \\begin{equation}\\label{CCvalue} \\CC_0=\\OL^0\\,\\rc^0\\simeq 6\\,h_0^2\\times 10^{-47}\\,GeV^{4}\\,. \\end{equation} {Here $\\rc^0 \\simeq \\left( 3.0\\,\\sqrt{h_{0}}\\times 10^{-12}\\,GeV\\right) ^{4}$ is the present value of the critical density, and $h_{0}\\sim 0.7\\pm 0.1$\\ sets the typical range for today's value of Hubble's constant $H_{0}\\equiv 100\\;(Km/sec\\, Mpc)\\;h_{0}$}. In the context of the Standard Model (SM) of electroweak interactions, this measured CC should be the sum of the original vacuum CC in Einstein's equations, $\\CC _{vac}$, and the induced contribution from the vacuum expectation value of the Higgs effective potential, $\\CC_{ind}=\\langle V_{\\rm eff} \\rangle$: \\begin{equation} \\CC=\\CC _{vac}+\\CC _{ind}\\,. \\label{lambdaphys} \\end{equation} It is only this combined parameter that makes physical sense, whereas both $\\CC _{vac}$ and $\\CC _{ind}$ remain individually unobservable\\footnote{{In general the induced term may also get contributions from strong interactions, the so-called quark and gluon vacuum condensates. These are also huge as compared to (\\ref{CCvalue}), but are much smaller than the electroweak contribution $V_{\\rm eff}$.}}. From the current LEP 200 numerical bound on the Higgs boson mass, $\\,M_{H}> 114.1\\,GeV$ \\cite{LEPHWG}, one finds $\\,\\,\\,\\left| \\CC _{ind}\\right|=(1/8)M_H^2\\,v^2> 1.0\\times 10^{8}\\,GeV^{4}$, where $v\\simeq 246\\,GeV$ is the vacuum expectation value of the Higgs field. Clearly, $\\left| \\CC _{ind}\\right|$ is $55$ orders of magnitude larger than the observed CC value (\\ref{CCvalue}). Such discrepancy, the so-called ``old'' cosmological constant problem \\cite{weinRMP,CCRev}, manifests itself in the necessity of enforcing an unnaturally exact fine tuning of the original cosmological term $\\CC _{vac}$ in the vacuum action that has to cancel the induced counterpart $\\,\\CC _{ind}$ within a precision (in the SM) of one part in $\\,10^{55}$. The measured CC remains very small as compared to the huge CC value predicted in the SM of Particle Physics. {Actually, if the physical value of the CC would conform with that one predicted in the SM, the curvature of our universe would be so high that the Special Theory of Relativity could not be a solution to Einstein equations to any reasonable degree of approximation. Therefore, the SM prediction of the CC violently contradicts our experience, whereas the small measured value (\\ref{CCvalue}) is perfectly compatible with it. The Cosmological Constant Problem (CCP) is a fundamental problem. It is most likely related to the delicate interplay between Gravity and Particle Physics, and it has become one of the main poles of attention \\cite{weinRMP,CCRev}. All attempts to deduce the small value of the cosmological constant from a sound theoretical idea ended up with the necessity of introducing severe fine-tuning. This concerns also, unfortunately, the use of supersymmetry and string theory (see e.g. \\cite{wittenDM,Carroll2}). In this respect we recall that, for a realistic implementation of the existing versions of M-Theory, one would like to have a negative (or at least vanishing) cosmological constant in the remote future, such that it does not prevent the construction of the asymptotic S-matrix states in accelerated universes\\,\\cite{MTheory}. {Since the presently observed value of the CC is positive, there is the hope that a variable cosmological term may solve this problem.} {There is a permanently growing flux of proposals concerning the CCP.} On the first place there is the longstanding idea of identifying the dark energy component with a dynamical scalar field \\cite{Dolgov,PSW}. More recently this approach took the popular form of a ``quintessence'' field slow--rolling down its potential\\,\\cite{Caldwell}. This proposal has, on its own, given rise to a wide variety of models \\cite{PeebRat,Moreq}. Extended models of this kind (``k--essence'') are also based on scalar fields but with a non-canonical kinetic energy \\cite{kessence}. {The main advantage of the quintessence models is that they could explain the possibility of a variable vacuum energy.} This may become important in case such variation will be someday detected in the observations. Recently other approaches have appeared in which the dark energy is mimicked by new gravitational physics \\cite{Carroll}. From the point of view of the CCP, all these approaches lead {to the introduction of either a very small parameter} or a very high degree of fine-tuning. In another, very different, vein the possibility to accept the observed value of the CC within the context of a many world pool is offered by the anthropic proposal \\cite{antrop}. Let us finally mention the intriguing proposal of non-point-like gravitons at sub-millimeter distances suggested in \\cite{Sundrum2}, or the possibility of having multiply degenerate vacua\\,\\cite{Yokoyama}. When assessing the possibility to have variable dark energy, other no less respectable possibilities should be taken into account. In a series of recent papers \\cite{JHEPCC1,cosm}, the idea has been put forward that already in standard Quantum Field Theory (QFT) one would not expect the CC to be constant, because the Renormalization Group (RG) effects may shift away the prescribed value, in particular if the latter is assumed to be zero. Thus, in the RG approach one takes a point of view very different from e.g. the quintessence proposal, as we deal all the time with a ``true'' cosmological term. It is however a variable one, and therefore a time-evolving, or redshift dependent: $\\CC=\\CC (z)$. Although we do not have a QFT of gravity where the running of the gravitational and cosmological constants could ultimately be substantiated, a semiclassical description within the well established formalism of QFT in curved space-time (see e.g. \\cite{birdav,book}) should be a good starting point. From the RG point of view, the CC becomes a scaling parameter whose value should be sensitive to the entire energy history of the universe -- in a manner not essentially different to, say, the electromagnetic coupling constant. One of the main distinctions between our approach and all kinds of quintessence models is that these models imply the introduction of a phenomenological equation of state $p_{\\chi}=w\\,\\rho_{\\chi}$ for the scalar field $\\chi$ mimicking the CC, where $w$ is a negative index (smaller than $-1/3$). Whether constant or variable, a ``true'' cosmological parameter has, instead, no other equation of state associated to it apart from the exact $w=-1$ one. Attempts to apply the RG for solving the CC problem have been made earlier \\cite{Polyakov,lam}. The {canonical form} of renormalization group equation (RGE) for the $\\CC$ term at high energy is well known -- see e.g. \\cite{book,Nelson}. {However, at low energy decoupling effects of the massive particles may change significantly the structure of this RGE, with important phenomenological consequences.} This idea has been retaken recently by several authors from various interesting points of view \\cite{JHEPCC1,cosm,Babic,Reuter03a}. However, it is not easy to achieve a RG model where the CC runs smoothly without fine tuning at the present epoch. {In Ref.\\,\\cite{Letter,IRGA03} a successful attempt in this direction has been made, which is based on possible quantum effects near the Planck scale.} At the same time, the approximate coincidence of the observed $\\CC$ and the matter density, $\\ \\OL^0\\sim \\OM^0\\,$, i.e. the ``new'' CC problem, or ``time coincidence problem'' \\cite{weinRMP,CCRev} {can be alleviated in this framework if we assume the standard (viz. Appelquist-Carazzone\\,\\cite{AC}) form of the low-energy decoupling for the massive fields.} In the present paper we elaborate on this idea further. We develop a semiclassical FLRW model whose running CC is driven smoothly, without fine tuning, due to generic quantum effects near the Planck scale. We show that, due to the decoupling phenomenon, the low-energy effects (in particular the physics from the SM scale) are irrelevant for the CC running, and so the approximate coincidence between $\\OM^0$ and $\\OL^0$ is not tied to any particular epoch in the history of the universe. Furthermore, the new effects imply deviations from the standard cosmological equations due to quantum effects. Our ``renormalized'' FLRW model provides a testable framework that can be thoroughly checked from SNAP data on Type Ia supernovae at very high z -- see \\cite{SNAP, goods} and references in \\cite{rev}. If these experiments detect a $z$-dependence of the CC similar to that predicted in our work, we may suspect that some relevant physics is going on just below the Planck scale. Alternatively, if they find a static CC, this might imply the existence of a desert in the particle spectrum near the Planck scale. The structure of the paper is as follows. In the next section we compare constant versus variable CC models. In Section \\ref{sect:RCC} we present our variable CC model based on the Renormalization Group. In Section \\ref{sect:solving} we solve the FLRW cosmologies with running CC. In Section \\ref{sect:numanalysis} we study the numerical behaviour of these cosmologies, and the predicted deviations from the standard FLRW expectations. In Section \\ref{sect:newmodel}, we introduce the magnitude-redshift relation for the analysis of the SNe Ia. In Section \\ref{sect:snap} we perform the simulations on the SNAP data in order to test the sensitivity with which the features of the new model can be determined. In the last section we draw our conclusions. {Two appendices are included at the end: one to discuss some technical issues inherent to our QFT framework, the other providing some background related to the statistical analysis.} ", "conclusions": "\\label{sect:conclusions} To summarize, we have constructed a semiclassical FLRW model with variable CC at the present cosmic scale. The variation of the vacuum energy is provided without introducing special scalar (quintessence-like) fields and is completely caused by quantum effects of vacuum. The evolution of the CC is due to the Renormalization Group running triggered by the smooth decoupling of the massive fields at low energies, while the RG scale $\\mu$ being associated to the Hubble parameter, $H$, at the corresponding epoch. Although the $\\bCC$-function itself is proportional to the fourth power of the masses, the decoupling does still introduce an inverse power suppression by the heavy masses, and thus one is left with a residual quadratic law $\\bCC\\sim H^2 M^2$. The effective scale $M$ summarizes the presence of the heaviest degrees of freedom available. This peculiar form of decoupling stems directly from: \\,i) the Appelquist-Carazzone (AC)-decoupling theorem, \\,ii) general covariance, and also from\\, iii) the \\textit{non}-fine-tuning hypothesis between the $n=1$ (``soft decoupling'') terms of $\\bCC$ (Cf. Eq. (\\ref{newRG1})) whereby the overall coefficient of the quadratic contribution $H^2 M^2$ does not vanish. This particular form of decoupling is a specific characteristic of the CC because it is of dimension four. There is no other parameter either in the SM or in the GUT models with such property. In constructing the cosmological model we have explored the possibility that the heaviest d.o.f. may be associated to particles having the masses just below the Planck scale. This assumption is essential to implement the soft decoupling hypothesis within the $\\mu=H$ setting. Indeed, the present value of $\\,H_0^2 M^2\\,$ is just of the order of the CC, and therefore it insures a smooth running of the cosmological term around the present time. In this setting the $\\bCC$-function has only one arbitrary parameter $\\,\\nu\\,$ (\\ref{nu}), and as a result the model has an essential predictive power. In general we expect $\\,|\\nu|\\ll 1\\,$ from phenomenological considerations, mainly based on the most conservative hypotheses on the nucleosynthesis framework. Furthermore, for $\\,|\\nu|\\ll 1$, we insure the absence of the trans-Planckian energies. It is not completely clear how much the cosmological index $\\,\\nu\\,$ can approach the value $1$ from below. Despite the values $\\nu\\lesssim 1$ are not completely ruled out, it is not clear that the conditions for the nucleosynthesis could be safe. However, if accepted, it opens up the possibility that the CC could be in transit from a $\\,\\CC>0\\,$ regime into a future $\\,\\CC\\leq 0\\,$ one, which would be a welcome feature for string/M-theory. Our RG model offers the possibility to explore the existence of the sub-Planck physics in direct cosmological experiments, such as SNAP (and the very high--z SNe Ia data to be obtained with HST). For example, for the flat FLRW case and a moderate and positive value $\\nu \\sim 10^{-2}$, we predict an increase of $10-20\\%$ in the value of $\\Omega_{\\Lambda}$ at redshifts $z\\gtrsim 1-1.5$ perfectly reachable by SNAP. For similar, but negative, values of $\\nu$ we predict that the CC should become negative beyond redshifts $z\\gtrsim 2$. For $\\nu\\sim 0.1$ corrections to some FLRW cosmological parameters become as large as $50\\%$ or more. In general, this model has a wide spectrum of implications that could be tested by SNAP, even for fairly moderate values of $\\nu$ compatible with the most conservative bounds from nucleosynthesis. The simulations of the SNAP data in Section 7 show that the single parameter $\\nu$ of our model could be pinned down, for $\\nu\\lesssim 0.1$, with a precision of around $20-60\\%$. Although this accuracy is not very high, the sign of the parameter could be determined and the effects would be manifest. Undoubtedly, this would be a good starting point to identify the presence of quantum corrections to the FLRW classical cosmology. The semiclassical FLRW model that we have proposed here explains the variation of the CC due to the ``relic'' quantum effects associated to the decoupling of the heaviest degrees of freedom below the Planck scale, and suggest that a time dependence of the CC may be achieved without resorting to scalar fields mimicking the cosmological term or to modifications of the structure of the SM of the strong and electroweak interactions and/or of the gravitational interactions. At the same time, our model provides a phenomenological parametrization for possible correlated deviations of the cosmological equations for matter density and dark energy. From a more fundamental point of view, we have shown that the cosmological constant and the matter density may evolve in a correlated way due to quantum effects without resorting to exceedingly exotic frameworks. Using the RG as a basic QFT tool and extrapolating the standard AC law of decoupling to the CC case, we have straightforwardly predicted a cubic redshift evolution law for the CC and the matter density. The larger the redshift that we can eventually explore the larger the effects that we predict. Most important, we have shown that this cubic law can be thoroughly tested by the next generation of cosmological measurements, which will be able to reach depths up to $z\\sim 2$." }, "0311/astro-ph0311629_arXiv.txt": { "abstract": "When a main sequence star evolves into a red giant and its Kuiper Belt Object's (KBO's) reach a temperature of ${\\sim}$170 K, the dust released during the rapid ice-sublimation of these cometary bodies may lead to a detectable infrared excess at 25 ${\\mu}$m, depending upon the mass of the KBO's. Analysis of IRAS data for 66 first ascent red giants with 200 L$_{\\odot}$ $<$ $L$ $<$ 300 L$_{\\odot}$ within 150 pc of the Sun provides an upper limit to the mass in KBO's at 45 AU orbital radius that is usually less than ${\\sim}$0.1 M$_{\\oplus}$. With improved infrared data, we may detect systems of KBO's around first ascent red giants that are analogs to our Solar System's KBO's. ", "introduction": "The formation and evolution of large solids such as comets, asteroids and planets in astrophysical environments is of great interest. Here, we investigate whether other stars possess Kuiper Belt Objects (KBO's) similar to those found in the Solar System. We assume that the hypothetical KBO's around other stars resemble those in our Solar System and have a composition of ice and dust similar to that of comets (see Luu \\& Jewitt 2002). During the main sequence phase of the star's evolution, the KBO's quiescently remain in stable orbits. However, when the star becomes a red giant, the KBO's become sufficiently hot that ice sublimates and previously embedded dust particles are ejected into the surroundings. In $\\S$2, we sketch our argument, which is presented in full detail in $\\S$3, that sufficient dust may be released during this red giant phase that the star can display a detectable infrared excess. Stern, Shull \\& Brandt (1990) and Ford \\& Neufeld (2001) computed the fate of ice sublimated from comets that have orbital radii greater than 100 AU when the host star evolves onto the Asymptotic Giant Branch and attains a luminosity near 10$^{4}$ L$_{\\odot}$. Now that the Kuiper Belt in the Solar System with most objects lying near 45 AU from the Sun is becoming better understood (Luu \\& Jewitt 2002), it is possible to imagine the response of such a system to a star's first ascent up the red giant branch when its luminosity exceeds 100 L$_{\\odot}$. However, although progress is being made, there are still significant uncertainties about the Kuiper Belt in the Solar System. Recent estimates of its mass range from 0.01 M$_{\\oplus}$ (Bernstein et al. 2003) to 0.1 M$_{\\oplus}$ (Luu \\& Jewitt 2002). Therefore, although we describe a procedure to investigate Kuiper Belt-like systems around other stars, our own outer Solar System is so poorly understood that an exact comparison is not yet possible. Two current descriptions of the KBO's differ mainly but not exclusively in estimates of the numbers of objects with radii smaller than ${\\sim}$200 km. With the expectation that short period comets originate in the Kuiper Belt, Luu \\& Jewitt (2002) propose a rapid rise in the numbers of such smaller KBO's. In contrast to this expectation, Bernstein et al. (2003) used the Hubble Space Telescope to search for KBO's as faint as 28.3 mag, corresponding to radii of ${\\sim}$15 km, and found many fewer objects than expected from the model given by Luu \\& Jewitt (2002). However, with this observational result of Bernstein et al. (2002), the origin of the short period comets becomes a mystery. Perhaps there is a bimodal distribution of KBO's. In this paper, given the uncertainties in current knowledge of the outer Solar System, we consider models both with and without large numbers of KBO's smaller than 200 km in radius. Most first ascent red giants do not show any evidence for circumstellar dust. The minority of first ascent red giants with infrared excesses (Judge, Jordan \\& Rowan-Robinson 1987, Jura 1990, 1999, Kim, Zuckerman \\& Silverstone 2001, Smith 1998, Zuckerman, Kim \\& Liu 1995, Plets et al. 1997) do not seem to have sublimating KBO's, since the characteristic inferred dust temperature is less than 70 K instead of 170 K, as we predict in $\\S$3. Here, we focus on the possibility of detecting relatively warm dust from disintegrating KBO's. ", "conclusions": "We have computed the evolution of a system of KBO's when a star becomes a red giant. For KBO's at 45 AU from the host star, we find that the dust released during the sublimation of ice from these objects might produce a detectable infrared excess around red giants of luminosities in the range 200 L$_{\\odot}$ to 300 L$_{\\odot}$ if the mass of the KBO's is at least 0.1 M$_{\\oplus}$. To date, there is no strong evidence for such sublimating KBO's, but with improved data, it may be possible to detect systems which are analogs to our Solar System's KBO's around first ascent red giants. This work has been partly supported by NASA." }, "0311/astro-ph0311303_arXiv.txt": { "abstract": "Six low-luminosity active galactic nuclei have been imaged at multiple frequencies from 1.7--43~GHz (2.3--15~GHz for three of the galaxies) using the Very Long Baseline Array. In spite of dynamic ranges of about 100 in several frequency bands, all six galaxies remain unresolved, with size limits at 8.4~GHz of $10^3$--$10^4$ times the Schwarzschild radii of the black holes inferred at their galactic centers. The galaxy spectra are roughly flat from 1.7 to 43~GHz, rather than steepening to classical optically thin synchrotron spectra at high frequencies. Although the spectral slopes somewhat resemble predictions for advection-dominated accretion flows, the luminosities are too high for the black hole masses of the galaxies and the slight spectral steepening at high frequencies cannot be explained by standard simple models of such accretion flows. In contrast, compact jets can accommodate the average spectral index, the relatively high radio luminosity, and the unresolved appearance, but only if the jets in all six galaxies are fairly close to our line of sight. This constraint is in agreement with inclination angle predictions for five of the six AGNs based on the dusty torus unification model. ", "introduction": "} The physical processes that control the accretion and outflow in low-luminosity active galactic nuclei (LLAGNs) are still not well understood. In their study of the Palomar Seyfert sample, \\citet*{Ho_FS1997.0} found that almost 40\\% of bright ($B_T < 12.5$~mag) nearby galaxies have optical spectra indicating the presence of an AGN. Although their line emission resembles scaled down versions of more luminous Seyfert nuclei, which are thought to be driven by accretion in thin disks, LLAGN properties at other wavelengths suggest that different processes are at work. Whereas ``classical'' Seyfert galaxies have radio spectral indices $\\alpha \\approx -0.7$, where $S_\\nu \\propto \\nu^{+\\alpha}$, almost half of the low-luminosity Seyfert galaxies in the Palomar Seyfert sample have flat or positive spectral indices \\citetext{\\citealp{Ulvestad_W1989}; \\citealp{Ho_U2001}, hereafter \\citetalias{Ho_U2001}; \\citealp{Ulvestad_H2001.0}, hereafter \\citetalias{Ulvestad_H2001.0}}. Similar results have been found for low-ionization nuclear emission regions (LINERs) and other AGNs in the Palomar sample (\\citealp{Nagar_ea2000,Nagar_ea2002}; \\citealp*{Filho_BH2000,Filho_BH2002}). These flat-spectrum objects typically appear compact on arcsecond scales, and high resolution studies with the Very Long Baseline Array\\footnote{The VLBA is operated by the National Radio Astronomy Observatory, a facility of the National Science Foundation, operated under cooperative agreement by Associated Universities, Inc.} \\citep[VLBA;][]{Napier_ea1993} show that they remain compact even on milliarcsecond scales \\citep[hereafter \\citetalias{Ulvestad_H2001.1}]{Falcke_ea2000,Ulvestad_H2001.1}. Following the work of \\citet{Ferrarese_M2000}, \\citet{Gebhardt_ea2000}, and others, it has become clear that both normal and active galaxies with bulges have supermassive black holes at their nuclear centers. Accretion onto these black holes is thought to fuel the AGN process for both ``normal'' and low-luminosity AGNs. Since their black hole masses cover roughly the same range \\citep[see, for example, ][]{Ho_2002}, the mass of the central object cannot account for the difference in emission levels. The explanation for why LLAGNs emit so little radiation may be either that they have far less mass accreting onto them, or that the dominant mode of the accretion mechanism itself must be different, or both. The center of our own Milky Way Galaxy has been used as the testbed for many models that attempt to explain the physics of LLAGNs. Acceleration measurements of stellar orbits near the source Sagittarius~A$^\\ast$ (Sgr~A$^\\ast$) leave little doubt as to the existence of a supermassive black hole of $\\sim 3.3\\times 10^6~\\mathrm{M_\\Sun}$ at the Galactic Center \\citep{Ghez_ea2003,Schoedel_ea2003}. Invisible in the optical and near-infrared, Sgr~A$^\\ast$ has a mildly inverted spectrum from 1 to 100~GHz \\citep{Falcke_ea1998}. The luminosity of Sgr~A$^\\ast$ is extremely low, even compared with other LLAGNs. Advection dominated accretion flows (ADAFs) have been used to explain the spectral energy distributions (SEDs) of Sgr~A$^\\ast$ and other LLAGNs during the past decade (\\citealp{Narayan_Y1994,Narayan_Y1995.1}; \\citealp*{Narayan_YM1995,Narayan_MQ1998}). In the ADAF model, a hot plasma develops in which the protons decouple from the electrons, allowing the protons to carry most of the energy released by the accretion process into the black hole, limiting the amount that is radiated away. Hot thermal electrons generate synchrotron radiation that is self-absorbed, resulting in a radio spectral index of $\\sim 0.4$ \\citep{Mahadevan_1997}. Bremsstrahlung and inverse-Compton processes generate additional emission, producing a relatively well-defined spectrum from radio through X-ray frequencies. An alternative hypothesis for the emission from LLAGNs is the compact-jet model \\citep{Falcke_B1995,Falcke1996,Falcke_B1999,Falcke_M2000}. Plasma is ejected in a collimated beam away from the black hole, and expands sideways to form a cone-shaped emission region. Synchrotron emission is initially self-absorbed, becoming progressively more transparent to lower frequencies as the material travels away from the black hole, producing an overall radio spectrum that can be quite flat. With both ADAF (and other related low-radiation accretion flows) and jet models failing to account for the overall SEDs in LLAGNs (see, for example, \\citealp{Narayan_MQ1998}; \\citealp*{Yuan_MF2002}), several attempts have been made recently to combine the two models. \\citet{Yuan_MF2002} have reasonable success in explaining the spectrum of Sgr~A$^\\ast$ with this combined model. Although this work has been extended to \\NGC{4258} \\citep{Yuan_ea2002}, rigorous testing of a large sample of LLAGNs in the nearby universe has yet to be performed. These LLAGNs are many orders of magnitude \\emph{more} luminous than Sgr~A$^\\ast$, but are also several orders of magnitude \\emph{less} luminous than ``classical'' AGNs. They therefore occupy a luminosity range that has thus far had few detailed studies. We have formed an LLAGN sample from the \\citetalias{Ho_U2001} study, selecting objects which have 5~GHz Very Large Array (VLA) peak flux densities between 5 and 30~mJy and spectral indices at milliarcsecond resolution of $\\alpha_{1.4}^{5} > -0.35$. In this paper we discuss simultaneous, multifrequency radio imaging of six LLAGNs. The purpose of these observations, and the point of the present paper, is to measure the sizes and radio spectra of LLAGNs within $10^4$~Schwarzschild radii of their black holes, and to determine what constraints these parameters can place on possible ADAF and jet models. ", "conclusions": "} We have performed multifrequency, milliarcsecond-scale radio imaging of six LLAGNs with flat/inverted spectra. All six galaxies have slightly rising spectra throughout the gigahertz range, and are unresolved on scales of $10^3$--$10^4$ Schwarzschild radii. The radio powers are higher than predicted by simple ADAF models, but the radio powers and spectra are consistent with dominance by compact radio jets. However, the unresolved nature of the radio emission implies that modified ADAF models including nonthermal electrons are still viable models. Otherwise, jet models require the jets to be seen closer to end-on than to side-on in all six galaxies. This result is consistent with the fact that at least five of the size galaxies are Type~1 galaxies, where unified schemes predict that our viewing angle is within $\\sim 45\\degr$ of the jet axis." }, "0311/astro-ph0311135_arXiv.txt": { "abstract": "We report the detection of the first 2 microlensing candidates from the Wendelstein Calar Alto Pixellensing Project (WeCAPP). Both are detected with a high signal-to-noise-ratio and were filtered out from 4.5 mill. pixel light curves using a variety of selection criteria. Here we only consider well-sampled events with timescales of $1\\E{d} < \\tfwhm < 20\\E{d}$, high amplitude, and low $\\chi^2$ of the microlensing fit. The two-color photometry $(R,I)$ shows that the events are achromatic and that giant stars with colors of $(R-I)\\approx 1.1\\E{mag}$ in the bulge of M31 have been lensed. The magnification factors are 64 and 10 which are obtained for typical giant luminosities of $M_I=-2.5\\E{mag}$. Both lensing events lasted for only a few days ($t_\\M{fwhm}^\\M{GL1}=1.7\\E{d}$ and $t_\\M{fwhm}^\\M{GL2}= 5.4\\E{d}$). The event GL1 is likely identical with PA-00-S3 reported by the POINT-AGAPE project. Our calculations favor in both cases the possibility that MACHOs in the halo of M31 caused the lensing events. The most probable masses, $0.08\\,M_\\odot$ for GL1 and $0.02\\,M_\\odot$ for GL2, are in the range of the brown dwarf limit of hydrogen burning. Solar mass objects are a factor of two less likely. ", "introduction": "Microlensing experiments are an ideal method to search for dark objects within and between galaxies. A large number of microlensing events have been detected towards the Galactic bulge constraining the number density of faint stars in this direction \\citep{alard99, afonso99,alcock00a,udalski00}. Towards the LMC only 13-17 microlensing events have been reported so far \\citep{alcock00b}. If all this events are attributed to $0.5\\,M_\\odot$ MACHOs, the associated population of dark objects would contribute up to the 20\\% level to the dark matter content of the Milky Way \\citep{alcock00b}. However, both the relatively large size of the LMC relative to its distance and the nature of the lenses has cast doubt on this interpretation. It is indeed likely that a large fraction of the microlensing events towards the LMC are due to self-lensing of stars within the LMC (\\citet{lasserre00,evans00} and references therein). Studying microlensing events towards M31 allows to separate self-lensing and halo-lensing in a statistical way, since the optical depth for halo lensing is larger on the far side of M31. In M31 individual stars can not be resolved and one therefore has to use the pixellensing technique \\citep{crotts92,baillon} to follow the variability of sources blended with thousands of other sources within the same pixel. First detections of possible microlensing events were reported by several pixellensing experiments \\citep{crotts96,ansari99,auriere01, paulinhenriksson02,paulinhenriksson03,calchinovati03}. But since the candidate nature of only 5 of these events is convincing, no conclusions concerning the near-far asymmetry or the most likely dark matter lensing masses could be drawn yet. The Wendelstein Calar Alto Pixellensing Project (WeCAPP, \\citealt{arno01}) started in 1997 with test observations. Since 1999 the bulge of M31 was monitored continuously during the time of visibility of M31. The analysis of our 4 year data will allow not only the identification of very short duration events (eg., in the 4$^\\M{th}$ year data of the combined field have been taken on 83 \\% of possible nights) but also the separation of long duration ($<20\\E{d}$) microlensing events from long periodic variables as Mira stars. For this letter we analyzed the short duration events ($\\tfwhm<20\\E{d}$) within one season of Calar Alto data and restricted the detection to high-signal-to noise, high-magnification events. We report our first 2 microlensing candidates of that type. ", "conclusions": "We presented the first two high ($S/N$), short timescale microlensing events from WeCAPP. GL1 is likely identical to PA-00-S3 found by POINT-AGAPE. Combining the data from AGAPE with ours shows that the error bar of the derived Einstein time scale becomes smaller by a factor of 5 compared to the individual error bar. This demonstrates the importance of a good time sampling of the events. We derived the colors of the lensed stars, the amplification factors and likely lens masses for both bulge/disk self-lensing and MACHO lensing. We showed that red giants are the likely source objects, while main sequence stars are highly implausible. Self-lensing in the bulge can only be separated from MACHO lensing statistically. Halo-lensing events show a spatial asymmetry because the optical depth for lensing events is higher for stars on the far side of M31 than on its near side \\citep{crotts92,theory}. In contrast, bulge self-lensing is symmetric. The bulge self-lensing hypothesis yields lensing stars at or below the the main sequence turn-off of the M31 bulge. On the other hand, if the lensing events are caused by MACHOs, their masses are typically very low, most probable below $0.1\\,M_\\odot$. Masses in the range of $0.5\\,M_\\odot$ to $1\\,M_\\odot$ are more unlikely. So far, we have analyzed one observing season and restricted the lensing search to short-time, high-amplification events in order to avoid confusion with variable stars. The whole WeCAPP dataset will allow us to identify all variables and thus will enable a search for lower amplitude and longer duration microlensing events. Decreasing the amplitude threshold will increase the detected rate of events in two ways. As the event rate is proportional to the inverse of the minimum required magnification $A_{0,\\M{min}}$ in the pixellensing regime we expect to detect more lensed giants. On the other hand lowering the amplification threshold could make it possible to detect also highly amplified main sequence stars \\citep{hangould} which exceed the evolved stars in the bulge of M31 by a factor of more than a hundred. How many more lensing events will be detected depends on the mass function of the lenses but we can expect at least a factor of a few \\citep{theory}. Finally, the effects of time sampling and noise properties of our sample on the detectability of lensing events have to be taken into account. Results of the modeling of these effects for events of different durations and amplitudes using Monte-Carlo simulations will be presented in a future publication. With the full dataset we expect therefore to increase the number of lensing events to detect the predicted asymmetry of MACHO lensing or to rule out a significant MACHO population in the halo of M31." }, "0311/astro-ph0311245_arXiv.txt": { "abstract": "We develop a formalism wherein the solution of the equations of equilibrium for a self-gravitating, magnetized interstellar cloud may be accomplished analytically, under the only hypotheses that ({\\it a}\\/) the cloud is scale-free, and ({\\it b}\\/) is electrically neutral. All variables are represented as series of scalar and vector spherical harmonics, and the equilibrium equations are reduced to a set of coupled algebraic e\\-qua\\-ti\\-ons for the expansion coefficients of the density and the magnetic vector potential. Previously known axisymmetric solutions are recovered and new non-axisymmetric solutions are found and discussed as models for pre-protostellar molecular cloud cores. ", "introduction": "{\\it ``Do you think that non-symmetric, magnetized, self-gravitating equilibria can exist?''} This question is a typical example of the classic, fundamental kind of problems that Frank likes to address. This contribution presents a summary of the results obtained by the writer in his effort to answer Frank's question. ", "conclusions": "Whether or not real cloud cores are triaxial ellipsoids that pulsate because of their intrinsic topological non-equilibrium, as suggested by the theoretical arguments discussed in the previous sections, the material presented in this contribution illustrates well the richness of possible implications concealed in a typical question asked by Frank." }, "0311/astro-ph0311559_arXiv.txt": { "abstract": "{It is well known that the surface brightness of elliptical galaxies and of bulges of spiral galaxies is best fitted by the Sersic $r^{1/n}$ profile. It is thus interesting to explore the lensing properties of the Sersic model because of its wide range of applicability. To this aim, we evaluate the lensing potential, the deflection angle, the time delay between the images and the amplification for a circularly symmetric lens whose surface density is described by the Sersic profile. We estimate the same quantities also adding a shear term to the deflection potential in order to study (at the lowest order) the impact of deviations from radial symmetry or the contribution from other nearby lenses. Moreover, we investigate the systematic errors due to the use of a de\\,Vaucouleurs profile instead of the correct Sersic model also taking into account the presence of a dark halo. ", "introduction": "Elliptical galaxies present a striking regularity in their global luminosity distributions in the sense that, within a wide range of sizes, their surface brightness profiles can be described by simple functions. The de Vaucouleurs or $r^{1/4}$\\,law (\\cite{deV48}) was first proposed as a quite general function to fit the light profile of the elliptical galaxies and of the bulges of lenticular and spiral galaxies. However, it was soon realized that the Sersic $r^{1/n}$ profile, which generalizes the de Vaucouleurs law, is best suited to describe the surface brightness distribution of these systems (\\cite{CCD93,GC97,PS97}). The regular properties of ellipticals have been the subject of different approaches, concerning both photometric and spectroscopic parameters, resulting in interesting scaling relations. Some well known examples are the fundamental plane relating the effective radius $r_e$, the luminosity intensity $I_e$ at $r_e$ and the central velocity dispersion $\\sigma_0$ (\\cite{DD87,7S87}); the photometric plane linking the three parameters of the Sersic profile (\\cite{KWKM00,G02}) and the entropic plane (\\cite{Metal01}). The regularity of elliptical galaxies is an attractive feature for cosmological applications of gravitational lensing since it makes it possible to reduce the degeneracy among lens parameters and hence the systematic uncertainty in the estimate of cosmological quantities. The time delay among the images in multiply imaged quasars systems may be used as a tool to determine the Hubble constant $H_0$ (\\cite{Ref64}) once a model for the lens has been fitted to the observed images configuration, flux and time delay ratios. The use of scaling relations makes it possible to narrow the parameter space expressing some of them as function of the remaining ones. On the other hand, it has been estimated that at least 80$\\%$ of gravitational lenses are elliptical galaxies (\\cite{TOG84,FT91}). Hence, it is important that statistical lensing studies, such as the number counts of lens systems, the distribution of image angular separations and of time delays, take into account the most correct model of elliptical galaxies in the analysis of the available data. Up to now, this has not been done since these studies have been concentrated on the dark halos rather than on the luminous components of the galaxies. Therefore, this latter component is usually neglected or simply modeled with the spherical Hernquist profile (\\cite{H90}) which approximately reproduces the de\\,Vaucouleurs law when projected to the lens plane. The only remarkable exception is the work by Maoz \\& Rix (1993) where the lens is modeled using the $r^{1/4}$\\,law directly, together with an isothermal dark halo. It should be interesting to repeat their analysis using the more realistic Sersic profile and the scaling relations we have quoted above. As a first step, one has to study the lensing properties of this model. This is the aim of the current work. The paper is organized as follows. In Sect.\\,2, we determine the lensing potential and the deflection angle for a circularly symmetric lens described by the Sersic model. We are thus able to write the lens equations whose solutions are found in Sect.\\,3. Section\\,4 is devoted to the evaluation of the amplification of the images, while the effect of a shear term, due to deviations from circular symmetry or to nearby lenses, is investigated in Sect.\\,5. A study of the systematic errors due to the use of a de\\,Vaucouleurs profile mimicking the best\\,-\\,fit Sersic model to the lens surface brightness is presented in Sect.\\,6. We summarize and conclude in Sect.\\,7. ", "conclusions": "Simple analytical calculations show that elliptical galaxies form 80$\\%$ of the gravitational lenses. Hence, it is very interesting to investigate their lensing properties taking into account models which are as realistic as possible. The Sersic $r^{1/n}$ profile is well known to fit the surface brightness of both elliptical galaxies and bulges of lenticular and spiral galaxies. In this paper we have studied its lensing properties assuming circular symmetry of the surface mass distribution. Under this hypothesis, we have derived analytical expressions for the lensing potential, the deflection angle and the image amplification. Solving the lens equations, we have found that a circularly symmetric Sersic lens produces two images of the source that are merged in the Einstein ring if $r_s = 0$, i.e. the source and the lens centre are perfectly aligned. One of the two images may be deamplified and hence, for some choice of $n$, a Sersic lens does not produce two observable images, an effect that must be taken into account in statistical studies such as lens number counts. No radial critical curve is formed by a circularly symmetric Sersic lens, while the tangential critical curve is the Einstein ring. Adding a shear term to the lensing potential allows us a first order investigation of the impact of deviations from circular symmetry and to take into account the contribution of nearby lenses. We have solved the lens equations in this case assuming, without loss of generality, that the shear is on\\,-\\,axis. Due to the breaking of radial symmetry, the number and positions of the images formed depend now both on $r_s$ and $\\theta_s$ and not only on $r_s$ as in the symmetric case. We have found that an on axis source gives rise to two images, while two or four images may appear as the source moves away from the lens axis. The presence of the shear also changes the amplification of the images and the shape of the critical curves. The tangential critical curve is deformed into an ellipse, while an inner radial critical curve appears. The size of both curves depends on the Sersic model parameters and decreases with the exponent $n$. \\begin{figure} \\centering \\resizebox{8.5cm}{!}{\\includegraphics{masserrsers+plum.eps}} \\hfill \\caption{Same as Fig.\\,\\ref{fig: masserriso}, but with $\\beta = -2$ for the halo model.} \\label{fig: masserrplum} \\end{figure} Our results suggest that a Sersic lens can produce up to four images of a distant source if a shear term is added to the potential. This is in agreement with the recent result by Kochanek (2003) who was able to fit constant $M/L$ models to two quadruple QSOs, namely RXJ0911+0551 (\\cite{Ba97,Bu98}) and PG1115+080 (\\cite{W80,Ietal96}), using the de\\,Vaucouleurs profile, that is to say a Sersic lens with $n = 4$. However, a direct comparison is not possible. To fit RXJ0911+0551, one has to use two lensing galaxies modeled with the de Vaucouleurs profile, taking also into account the presence of the cluster which the lenses belong to. A similar situation also holds for PG1115+080 where a single de Vaucouleurs lens is unable to fit the images configuration and the contribution of a nearby cluster is strongly required to reproduce the observable quantities. The Sersic profile has been shown to be a better fit to the surface brightness distribution of elliptical lensing galaxies than the de\\,Vaucouleurs law. Nonetheless, the de\\,Vaucouleurs model is usually used in lensing studies. We have thus investigated the systematic errors induced by this procedure. This analysis has shown that the mass\\,-\\,to\\,-\\,light ratio $\\Upsilon$ may be seriously under\\,- or overestimated (depending on $n$ being lower or larger than 4) when considering constant $M/L$ models. On the other hand, if a dark halo is added to the visible component of the lens galaxy, the use of a de\\,Vaucouleurs model instead of the correct Sersic one leads to a wrong estimate of the dark halo projected mass of the dark halo inside the Einstein ring. These considerations could suggest that the use of de\\,Vaucouleurs profile to model galaxies that are best fitted by the Sersic law must be avoided. Actually, these results should be treated with great caution since they have been obtained using simulated galaxies whose model parameters have been fixed with a procedure that is by no means unique. In particular, as described in detail in the Appendix, we have used the photometric plane (\\cite{KWKM00,G02}) to relate the Sersic parameters $(n, r_e, I_e)$. Actually, this relation has been determined from galaxies belonging to the Virgo and Fornax clusters, while we have extended it to higher redshift galaxies the validity of which is still to be proven. Henceforth, a different prescription to fix the model parameters could be implemented and a detailed study of how this affects the results is needed. However, the procedure we have adopted leads to Sersic parameters describing galaxies which look quite reasonable so that we are confident that our main results are trustworthy. Actually, the best strategy is to resort to a careful study of real lens systems in order to check whether the effects we have reported are indeed true, or a consequence of how we have chosen the model parameters. The analytical expressions of the lensing potential and of the deflection angle here obtained for a circularly symmetric Sersic lens may be easily generalized to the flattened case following the procedure described, e.g., in Keeton (2001). The resulting potential, deflection angle, and amplification may then be used to try to fit the observed lens systems (both double and quadruple) in order to investigate if constant $M/L$ models are viable solutions or whether a dark matter component has to be invoked to reproduce the observable lensing quantities (images positions and flux ratios). Since the Sersic model parameters $(n, r_e, I_e)$ may be directly measured from detailed photometry, the only unknown lens quantity is the mass\\,-\\,to\\,-\\,light ratio $\\Upsilon$ and hence we can study eventual correlations among $\\Upsilon$ and the photometric parameters. On the other hand, we can assume that lensing galaxies obey the same scaling relations (such as the photometric plane and the fundamental plane) as the low redshift systems. If we model the lens as the sum of a luminous Sersic component and a dark halo, we can reduce the degeneracy among model parameters excluding those solutions which lead to Sersic parameters not allowed by the quoted scaling relations. Moreover, if the time delays among the images were measured, this approach should reduce the systematic error in the determination of the Hubble constant through the time delay method (\\cite{Ref64}). On the other hand, statistics of gravitational lensing, such as lenses number counts and distribution of images angular separation, are known to be powerful tools to discriminate among different cosmological models. The modeling of the lenses plays a key role in this kind of analysis and it is thus very important to use realistic lens models. The Sersic lens we have studied may thus be a first step towards a reanalysis of the available data in order to better investigate if it is possible to reconcile them with constant $M/L$ models and if the constraints on the cosmological parameters are changed by the use of more realistic models of elliptical galaxies. These questions will be addressed in a series of forthcoming papers." }, "0311/astro-ph0311073_arXiv.txt": { "abstract": "We searched for spectral signatures from the exosphere of HD209458 in the UV and the near-IR with UVES and ISAAC at the VLT. We looked in particular for the helium absorption feature at 10 830 \\AA\\, predicted to be among the strongest ones. The upper limit of the He I line derived is 0.5\\% at 3 $\\sigma$ for a 3 \\AA\\ bandwith. Planet-induced chromospheric activity search on HD209458 was also performed. ", "introduction": "The He I feature at 10 830 \\AA\\ is predicted to be one of the deepest absorption lines of the diffuse and extended exosphere in the optical and near-IR transmission spectrum The goal of the present study is to search for exospheric signatures in the spectrum of HD209458, during transits of the planet. Shkolnik et al. (2003) suggested that a chomospheric activity can be induced by the planet orbit. We searched for such a signature in stellar chromospheric lines. ", "conclusions": "" }, "0311/astro-ph0311590_arXiv.txt": { "abstract": "Optical spectroscopy of the Type Ia supernova SN 2002ic obtained on 2003 June 27.6~UT, i.e., $\\sim$ 222 rest-frame days after explosion, is presented. Strong H emission indicates an interaction between the expanding SN ejecta and an H-rich circumstellar medium (CSM). The spectrum of SN~2002ic resembles those of SNe 1997cy and 1999E. The three SNe also have similar luminosities, suggesting that they are the same phenomenon and that the CSM is also similar. We propose a new classification, Type IIa SNe, for these events. The observed line profiles and line ratios are measured and discussed within the ejecta-CSM interaction scenario. The emission in H Balmer, [\\ion{O}{3}], and \\ion{He}{1} lines, and in permitted \\ion{Fe}{2} blends, resembles the spectra of the Type IIn SN 1987F and of Seyfert 1 galaxies. A high-density, clumpy CSM is inferred. Strong, very broad [\\ion{Ca}{2}]/\\ion{Ca}{2} and [\\ion{O}{1}]/\\ion{O}{1} emissions imply that not all the outer SN ejecta were decelerated in the interaction, suggesting that the CSM is aspherical. ", "introduction": "\\citet{ham03} reported strong \\ion{Fe}{3}, \\ion{Si}{2}, and \\ion{S}{2} features in the early-time spectra of SN~2002ic and classified it as a Type Ia supernova (SN Ia). However, strong H$\\alpha$ emission was also observed. The detection of H$\\alpha$ is unprecedented in an SN Ia. (For reviews on SN spectra, see \\citealt{fil97}). The emission was broad (FWHM $> 1000$ km~s$^{-1}$), suggesting that it was intrinsic to the SN. \\citet{ham03} suggested that it arose from the interaction between the SN ejecta and a dense, H-rich circumstellar medium (CSM), as in SNe IIn (e.g., \\citealt{chu91,che94}). If this interpretation is correct, SN~2002ic may be the first SN Ia to show direct evidence of the circumstellar (CS) gas ejected by the progenitor system, presenting a unique opportunity to explore the CSM around an SN Ia and the nature of the progenitor system. In this Letter, we present optical spectroscopy of SN~2002ic obtained more than 200 days after explosion, and discuss it within the context of the ejecta-CSM interaction. ", "conclusions": "\\label{sec-disc} The currently preferred model for SNe Ia is the thermonuclear explosion of a C+O white dwarf (WD) in a binary system, reaching the Chandrasekhar limit via either accretion from a normal companion (the SD scenario, which is generally favored) or merging with another WD (the DD scenario). (For recent reviews, see \\citealt{nom00,liv00}.) The SD scenario predicts the presence of an H/He-rich CSM. The discovery of strong CSM interaction in SN~2002ic may prove that this scenario does exist in nature. SNe~1997cy and 1999E may also be CS-interacting SNe Ia. However, such events are rare. Solid observational evidence of CSM has not yet been found in other SNe Ia \\citep{lun03}. Our spectral analysis suggests a high-density \\ion{H}{1}-emitting CSM ($n \\sim 10^8-10^9$~cm$^{-3}$); so this is probably clumpy. Assuming the CSM was formed in a progenitor wind, we relate the mass-loss rate $\\dot{M}$ and the wind velocity $u$ to the H$\\alpha$ luminosity, through $L({\\rm H_\\alpha})\\sim 1-5\\times 10^{39}$ $(\\dot{M}/10^{-2}{\\rm~M_\\sun~yr^{-1}})^2$ $(u/100 {\\rm~km~s^{-1}})^{-2}$ $r_{16}^{-1} f^{-1}$ ergs~s$^{-1}$, where $r_{16}$ is the radius of the CSM shock in units of $10^{16}$ cm and $f$ is the CSM filling factor. For $u\\sim 100$ km s$^{-1}$ and $f\\sim 0.01$, $\\dot{M}$ can be as high as $\\sim 10^{-2}$ M$_\\sun$~yr$^{-1}$. The case B emissivity used here may underestimate $L({\\rm H_\\alpha})$ by $\\sim 50\\%$, considering that most H$\\beta$ photons may have cascaded to H$\\alpha$. We have identified high-velocity lines (FWHM $\\gtrsim 10^4$ km$^{-1}$) emitted in the ejecta of SN~2002ic. Based on the hydrodynamical model of T. Suzuki et~al. (2004, in preparation), we suggest an equatorially concentrated structure for the CSM to explain the coexistence of these lines with the strong CS interaction. Such a geometry is not unexpected for the mass loss from a binary system or for stars approaching the end of the asymptotic giant branch (AGB). A preexisting clumpy disk was also suggested by \\citet{wan04}, based on spectropolarimetry. Is $\\dot{M}\\sim 10^{-2}$ M$_\\sun$~yr$^{-1}$ too high for the SD scenario? The highest $\\dot{M}$ observed in AGB stars is between $10^{-3}$ and $10^{-4}$ M$_\\sun$~yr$^{-1}$ \\citep{ibe95}; $\\dot{M}$ in our estimate is scaled as $u$. The bulk of the CSM may be near the equator and at low velocity ($\\sim 10$ km~s$^{-1}$), which cannot be resolved. The observed H$\\alpha$ P Cygni, dominated by the emission component, could be produced by some high-velocity CSM ($\\gtrsim 100$ km~s$^{-1}$) along the pole, which should be close to our line of sight. Similar wind patterns have been observed in some symbiotic stars (e.g., \\citealt{sol85}), which have been suggested as candidates for SN Ia progenitors. Further observations and modeling are required to understand the nature of the CSM giving rise to this type of event, which may be classified as ``Type IIa.'' Suggestions include a common envelope \\citep{liv03}, WD accretion wind \\citep{hac99}, and the superwind from an AGB star exploding as a ``type 1.5'' event \\citep{ham03}." }, "0311/astro-ph0311523_arXiv.txt": { "abstract": "We apply the tidal truncation model proposed by \\citet{neg01} to arbitrary Be/compact star binaries to study the truncation efficiency dependance on the binary parameters. We find that the viscous decretion disks around the Be stars could be truncated very effectively in narrow systems. Combining this with the population synthesis results of Podsiadlowski, Rappaport and Han (2003) that binary black holes are most likely to be born in systems with orbital periods less than about 30 days, we suggest that most of the Be/black-hole binaries may be transient systems with very long quiescent states. This could explain the lack of observed Be/black-hole X-ray binaries. We also discuss the evolution of the Be/black-hole binaries and their possible observational features. ", "introduction": "X-ray binaries consist of a neutron star (NS) or a black hole (BH) accreting from strong stellar winds or via Roche lobe overflow of the companion star. Depending on the masses of the optical companions, they are conventionally divided into low- and high-mass X-ray binaries. By measuring the radial velocity curve of the non-degenerate donor, the value of the mass function can be determined, which provides a lower limit on the mass of the accreting object. When it is combined with the information of the spectra of the donor and the orbital light curves, actual measurements of the mass can be obtained. In this way, currently the masses of 18 compact objects in X-ray binaries have been found to exceed the the maximum mass possible for a neutron star. These binaries are thought to be black-hole X-ray binaries (BHXBs) (see \\citet{mcc03} and references therein). Among the 18 BHXBs, 3 are persistently bright X-ray systems and 15 are X-ray novae in which 6 show recurrent outbursts. All of the X-ray novae are low-mass X-ray binaries (LMXBs). The three persistent sources, i.e. Cyg X$-1$, LMC X$-1$ and LMC X$-3$ are high-mass X-ray binaries (HMXBs) with massive O/B companion stars. In the most recent catalogue of high mass X-ray binaries (HMXBs) edited by \\citet{liu00}, only 20 out of 130 HMXBs are O/B supergiant systems including the black hole binaries. Though roughly two thirds of HMXBs are Be/X-ray binaries, and X-ray pulsations have been found in about sixty Be/X-ray binary systems \\citep{zio02}, there are no acknowledged Be/BH binaries. The existence of Be/BH binaries becomes an open problem. The solutions may be related to two questions: Do they exist? if yes, how can they be observed? The first question is connected with the birth-rate estimation of BH/massive X-ray binaries (BMXBs). Using Monte-Carlo simulation, \\citet{rag99} calculated the number and distribution of binary BHs with Be stars. They obtained the expected number of Be/BH binaries to be of order of unity per $20-30$ Be/NS systems. The Be/BH binaries were found to be highly eccentric with orbital periods lying in a range of 10 days to several years. The actual number of Be/BH binaries could be even higher since in their work, the lower limit of $50 \\, M_{\\sun}$ for the progenitor of the BH was adopted, based on the mass of the supergiant Wray 977 \\citep{kap95}. But current stellar evolution models suggest that stars with mass higher than about $20-25\\, M_{\\sun}$ will result in BHs\\citep{fry99, fry01}. More recently \\citet{pod03} did the binary population synthesis calculation with the mass of the BH progenitors in the range of $25\\, M_{\\sun}4$ \\AA, and a quiescent population with no significant star formation at the present day. We have investigated how this distribution depends on environment, as characterised by: \\begin{itemize} \\item The projected surface number density of galaxies $\\Sigma_5$, determined from the distance to the fifth-nearest neighbour. \\item The three-dimensional number density of galaxies measured on 1.1 Mpc and 5.5 Mpc scales, $\\rho_{1.1}$ and $\\rho_{5.5}$. \\item The velocity dispersion of the embedding structure, as determined from the group catalogues of \\citet{Eke-groups} and Nichol, Miller et al. (in preparation). \\end{itemize} We use these different measures of environment to establish the scales and structures on which the present-day galaxy population depends. Our findings are summarised as follows: \\begin{itemize} \\item[1.] The distribution of H$\\alpha$ line strength for the star-forming population, selected on \\ewha\\ or $(g-r)$ colour, does not itself depend strongly on environment. Thus, it is unlikely that SFRs are gradually decreasing in a substantial number of star-forming galaxies in or near dense environments today. \\item[2.] The fraction of galaxies with \\ewha$>4$\\AA\\ decreases steadily with increasing local density. There is evidence that it decreases more strongly at densities exceeding $\\Sigma_5\\sim1$ Mpc$^{-2}$, or $\\rho_{1.1}\\gtrsim 0.05$ Mpc$^{-3}$. The persistence of this correlation at low densities means that ram--pressure stripping, at any redshift, cannot be the only physical mechanism at work. \\item[3.] The fraction of galaxies brighter than $M^\\ast+1$ with \\ewha$>4$\\AA\\ is never more than $\\sim 70$ per cent, even in the least dense environments explored here. We have shown that this means the recent decline in globally-averaged star formation rate cannot be wholly due to the growth of large scale structure. \\item[4.] The emission-line fraction of a galaxy population appears to depend both on the local environment (on $\\sim 1$ Mpc scales) {\\it and} on the large-scale structure as parameterised by the density $\\sim 5.5$ Mpc scales. There is little further dependence on the velocity dispersion of the group or cluster in which the galaxy is embedded. This result suggests that the composition of the galaxy population today is likely related {\\it indirectly} to its present environment \\citep[see also][]{ZM98,ZM00}. \\end{itemize} The most likely physical explanations for the correlation between \\ewha\\ distribution and environment at $z=0$ are those which are effective over a large range of environment, affect the SFR on short ($<1$ Gyr) timescales, and were much more effective in the past. As suggested by previous authors \\citep[e.g.][]{Z+96,MZ98,H+98,HO00}, one good candidate is starbursts induced by galaxy interactions, since close pairs of galaxies are the only environment known to directly provoke a physical transformation. Such interactions will likely be more common at high redshift, and will have had more time to influence galaxies that end up in high density environments." }, "0311/astro-ph0311465_arXiv.txt": { "abstract": "A number of very small isolated \\HII\\ regions have been discovered at projected distances up to 30 kpc from their nearest galaxy. These \\HII\\ regions appear as tiny emission line objects in narrow band images obtained by the NOAO Survey for Ionization in Neutral Gas Galaxies (SINGG). We present spectroscopic confirmation of four isolated \\HII\\ regions in two systems, both systems have tidal \\HI\\ features. The results are consistent with stars forming in interactive debris due to cloud-cloud collisions. The \\Ha\\ luminosities of the isolated \\HII\\ regions are equivalent to the ionizing flux of only a few O stars each. They are most likely ionized by stars formed in situ, and represent atypical star formation in the low density environment of the outer parts of galaxies. A small but finite intergalactic star formation rate will enrich and ionize the surrounding medium. In one system, NGC 1533, we calculate a star formation rate of $1.5\\times10^{-3}$ \\msun yr$^{-1}$, resulting in a metal enrichment of $\\sim1\\times10^{-3}$ solar for the continuous formation of stars. Such systems may have been more common in the past and a similar enrichment level is measured for the `metallicity floor' in damped Lyman-$\\alpha$ absorption systems. ", "introduction": "\\label{sect:intro} \\HII\\ regions signifying the presence of highly ionizing OB stars are usually found in the luminous inner regions of galaxies \\citep[e.g.][]{Martin01}. \\HII\\ regions are also located in the faint outer arms of spirals \\citep[e.g.][]{Ferguson98b}, and as single or multiple star forming knots in narrow emission line dwarfs (\\HII\\ galaxies). In each case, new stars are formed in the vicinity of an existing stellar population. However, recent observations by \\cite{Gerhard02} have spectroscopically confirmed an isolated compact \\HII\\ region on the extreme outskirts of a galaxy (NGC 4388) in the Virgo cluster. Several luminous \\Ha-emitting knots have also been discovered in a compact group in the A1367 cluster \\citep{Sakai02}. In these cases it appears that the \\HII\\ regions are due to newly formed stars where no stars existed previously, albeit in a galaxy cluster environment. As they evolve, OB stars increase the metal abundance in their local environment. Absorption line studies show that the intergalactic medium (IGM) and galaxy halos, including our own, are enriched \\citep[e.g.][]{Chen01,Tripp02,Collins03}. Isolated \\HII\\ regions provide a potential source for this enrichment. In situ star formation in the IGM offers an alternative to galactic wind models to explain metal enrichment hundreds of kilo-parsecs from the nearest galaxy. Here we present a number of very small isolated \\HII\\ regions that have been discovered by their \\Ha\\ emission in the narrow band images obtained by the NOAO Survey for Ionization in Neutral Gas Galaxies (SINGG). SINGG is an \\Ha\\ survey of an \\HI-selected sample of nearby galaxies. The survey is composed of nearly 500 galaxies from the \\HI\\ Parkes All-Sky Survey (HIPASS, Barnes et al. \\citeyear{Barnes01}; Meyer et al. \\citeyear{Meyer03}), of these about 300 have been observed in \\Ha. Since a gaseous reservoir is a prerequisite for star formation, SINGG measures a broad census of star formation in the local Universe. The \\HII\\ regions appear as tiny emission line objects at projected distances up to 30 kpc from the apparent host galaxy. \\HII\\ regions are defined as isolated when they are projected at least twice the $\\mu_R =$ 25 ${\\rm mag\\, arcsec^{-2}}$ isophotal radius from the apparent host galaxy. This is typically much further than outer disk \\HII\\ regions in spiral galaxies \\citep{Ferguson98b}. In fact, for the systems discussed in detail here it is not totally clear whether the isolated \\HII\\ regions are even bound to the apparent host, hence we refer to them as ``intergalactic''. Their high equivalent widths suggest they are due to newly formed stars where no stars existed previously. In Section~2 spectra is presented for five isolated \\HII\\ regions candidates in three systems. These five candiates are referred to as the spectroscopically {\\it{detected}} emission line objects or \\HII\\ region candidates. All but one source has \\Ha\\ detected at a comparable recessional velocity to the nearest galaxy, two sources are also detected in \\oiii\\ as further confirmation. These four objects are referred to as the spectroscopically {\\it{confirmed}} isolated \\HII\\ regions. Optical spectra and \\HI\\ distributions for all three systems are described in Sections~3$-$5. In Section~6 models of the underlying stellar population, scenarios for star formation, enrichment of the IGM, implications of the intergalactic star formation rate, and the possibility that isolated \\HII\\ regions are progenitors of tidal dwarf galaxies are discussed. H$_0$=75 \\kms Mpc$^{-1}$ is used throughout. ", "conclusions": "\\label{sect:conclusion} The discovery of intergalactic \\HII\\ regions presented here and in other recent publications provides a small but finite source of enrichment and ionization of the IGM. In two cases the fact that these emission line objects are detected in both \\Ha\\ and \\oiii\\ rules out the possibility that they are background emitters. The \\Ha\\ luminosities imply that each isolated \\HII\\ region is ionized by 4$-$7 O stars. If these stars have formed in situ they represent atypical star formation in a low density environment. The low level of continuum emission from three of four confirmed isolated \\HII\\ regions suggests the stellar populations are very young and have formed where no stars existed previously. If part of a normal IMF, the corresponding total cluster mass would be $\\sim10^{3}$ \\msun. In two out of three systems, isolated \\HII\\ regions are associated with tidal \\HI\\ features, providing a reservoir of neutral gas. In one particular system, NGC~1533, the mass, distribution and velocity dispersion of the \\HI\\ suggests the rate of star formation ($1.5\\times10^{-3}$ \\msun yr$^{-1}$) could be sustained by the collision of clouds. This would result an increase in the metal abundance by $\\sim1\\times10^{-3}$ solar. This is the same abundance level as seen in the DLA `metallicity floor' \\citep{Prochaska03}. The amount of intergalactic high column density \\HI\\ and rate of collision-triggered intergalactic star formation may have been higher in the past. On-going investigations into the metallicities and underlying stellar population of these and other isolated \\HII\\ regions in the SINGG images will shed more light on their nature and origin." }, "0311/astro-ph0311186_arXiv.txt": { "abstract": "Subarcsecond infrared and radio observations yield important information about the formation of super star clusters from their surrounding gas. We discuss the general properties of ionized and molecular gas near young, forming SSCs, as illustrated by the prototypical young, forming super star cluster nebula in the dwarf galaxy NGC~5253. This super star cluster appears to have a gravitationally bound nebula, and the lack of molecular gas suggests a very high star formation efficiency, consistent with the formation of a large, bound cluster. ", "introduction": "While the first suggestions that young analogues to globular clusters exist in the local universe came from ground-based imaging (Arp \\& Sandage 1985), the age of the super star cluster really dawned with the subarcsecond imaging capabilities of HST, which allows super star clusters (SSCs) to be resolved in nearby galaxies (e.g., O'Connell et al. 1995, Whitmore \\& Schweizer 1995) The Antennae galaxies alone have thousands of young (Myr) SSCs (Whitmore et al. 1999.) Some of these clusters are revealed only by infrared imaging (Vigroux et al. 1996); presumably these are the youngest regions. Clues to the mystery of how large and potentially bound clusters form lie in the environments of the youngest, embedded regions, which are visible only in the radio and infrared. That large extragalactic compact HII regions exist and can be detected was first inferred for the starburst in M83 by Turner, Ho, \\& Beck (1987) on the basis of a high Brackett line/radio continuum ratio, although their 8\\arcsec\\ (160 pc at M83) infrared beam did not allow them to define the nature of the unexpected Brackett excess (cf. Thompson 1987). With arcsecond mid-infrared imaging, Telesco and Gezari (1992) isolated a compact 12 $\\mu$m source that they suggested was a forming globular cluster in M82. Pina et al. (1992) and Keto et al. (1993) found a similar strong and compact mid-infrared source in the starburst galaxy, NGC~253. High spatial resolution IR imaging and spectroscopy with HST and large ground-based telescopes and the expanded high frequency imaging capability at the VLA now allow the study of nebulae in the infrared and radio at subarcsecond resolution, corresponding to scales of $\\sim$1 pc in local galaxies. With this resolution one can now study the star formation processes in individual regions. Due to time and space limitations we will focus on the impact of these observations on issues in the formation of large bound clusters as exemplified by the well-studied young embedded SSC in the dwarf galaxy NGC 5253. ", "conclusions": "High resolution radio and infrared observations in nearby galaxies have found luminous and compact nebulae excited by obscured super star clusters. The IR emission of these nebulae can be a significant, or even dominant, part of the total galactic flux. In NGC 5253, there are so many young stars in such a small volume that the gravitational attraction appears to be slowing or stopping the rapid expansion of the nebula. The star formation efficiency also appears to be quite high in this galaxy, $\\sim 60-75$\\%, which is consistent with the formation of bound clusters. Star formation in SSC-starburst galaxies does not much resemble that closer to home. \\begin{figure} \\plotone{turner_j_fig3.eps} \\caption{CO in NGC 5253, from Meier et al. 2002. The white cross marks the supernebula. There are 5 clouds of gas and dust in this figure, marked A through E. Only cloud D is resident within the galaxy, while the other clouds are falling in. Cloud D has a mass of $\\sim 4 \\times 10^6~ \\rm M_\\odot.$} \\end{figure}" }, "0311/astro-ph0311192_arXiv.txt": { "abstract": "We present high resolution ($\\leq 20$~mas) observations of the radio continuum emission at 1.4 GHz from three high-redshift quasars: J1053$-$0016 ($z=4.29$), 1235$-$0003 ($z=4.69$), and J0913$+$5919 ($z=5.11$), thereby doubling the number of $z>4$ radio-loud quasars that have been imaged at mas resolution. The observations were carried out with the Very Long Baseline Array (VLBA) of the NRAO. All three sources are unresolved in these observations, with source size limits of a few mas. In all cases the flux densities measured by the VLBA are within 10$\\%$ of those measured with the VLA, implying that the sources are not highly variable on yearly timescales. We find no indication for multiple images that might be produced by strong gravitational lensing on scales from 20 mas (VLBA) to a few arcseconds (VLA), to dynamic range limits of $\\sim 100$. ", "introduction": "Surveys such as the Sloan Digital Sky Survey (SDSS) \\citep{YOR00} and the Digitized Palomar Sky Survey \\citep{DJO99} have revealed large samples of quasi-stellar objects (QSOs) out to $z \\sim 6$. Studies by \\citet{FAN02} have shown that at such a high redshift we are approaching the epoch of reionization, the edge of the ``dark ages'', when the first stars and massive black holes were formed. Eddington-limit arguments suggest that the supermassive black holes at the center of these QSOs are on the order of $10^9~M{_\\odot}$. If the correlation between bulge and black-hole masses \\citep{GEB00,FEME00} also holds at these high redshifts, then these sources have associated spheroids with masses on the order of $\\sim 10^{12}~M_\\odot$. It is challenging to explain the formation of such massive structures on relatively short timescales {($\\sim$~1~Gyr)}. However, \\citet{WL02} estimate that almost one third of known quasars at $z\\sim6$ ought to be lensed by galaxies along the line of sight. If these quasars are indeed gravitationally lensed, the estimated masses of their associated spheroids could be smaller by up to an order of magnitude; this would allow a less efficient assembly process. High resolution radio observations of high-redshift radio-loud quasars can be used to test for strong gravitational lensing by looking for multiple imaging on scales from tens of milliarcseconds (mas) to arcseconds. Also, Very Long Baseline Interferometry (VLBI) observations of core-jet radio sources in quasars over a large range in redshift have been used to constrain the cosmic geometry, under the assumption that such sources are (roughly) 'standard rulers' \\citep{GKF99}. In general, the high resolution of the VLBI observations permits a more detailed look at the physical structures in the most distant cosmic sources. To date, only three radio loud quasars at $z > 4$ have been imaged at mas resolution \\citep{FRE97,FRE03}. In this paper we present Very Long Baseline Array (VLBA) observations of three more quasars at $z> 4$: SDSSp J105320.42$-$001649.7 at $z=4.29$, SDSSp J123503.04$-$000331.7 at $z=4.69$, and SDSSp J091316.56$+$591921.5 at $z=5.11$ (hereafter J1053$-$0016, J1235$-$0003, and J0913$+$5919). J1053$-$0016 was first identified by \\citet{IMH91} and \\citet{STD94} as BRI 1050$-$0000, while the other two sources were discovered in the five-color imaging data from the SDSS \\citep{FAN00,AND01}. All three quasars are known to be radio-loud \\citep{CAL01,PET03} Throughout this paper, we assume a flat cosmological model with $\\Omega_{m}=0.3$, $\\Omega_\\Lambda=0.7$, and ${H_{0}=65}$~km~s$^{-1}$~Mpc$^{-1}$. In this model, 1~mas corresponds to about 7~pc at $z=4.5$. ", "conclusions": "The VLBA imaging of three $z> 4$ radio loud quasars (J1053$-$0016, J1235$-$0003, and J0913$+$5919) shows that these sources are smaller than a few mas in size, corresponding to physical scales of $\\le 20$ pc. On the other hand, all three sources show falling spectra at high frequency, and none of the sources are variable on yearly timescales. These results suggests that the sources are likely core-jet radio sources, and hence may be used to extend to very high redshifts the studies of cosmic geometry using core-jet radio sources (Gurvits et al. 1999). Higher frequency VLBA observations at higher spatial resolution are planned to test this possibility. We find no indication of multiple radio components in the fields of these sources on scales of 20 mas to a few arcseconds, to dynamic range limits of $\\sim 100$. A similar conclusion was reached for three other $z>4$ radio loud quasars \\citep{FRE97,FRE03}. These results imply that at least these six high $z$ quasars are not strongly gravitationally lensed. The compact nature of these quasars make them excellent candidates for future H~{\\footnotesize I} 21~cm absorption experiments to detect the neutral IGM in their host galaxies \\citep{FL02}, in particular the steep spectrum source J0913$+$5919. Such a search is currently under-way using the Giant Meter Wave Radio Telescope. Knowledge of source structure, as presented herein, is critical for both identifying potential candidates for H~{\\footnotesize I} 21~cm absorption searches, and for subsequent interpretation of the results." }, "0311/astro-ph0311471_arXiv.txt": { "abstract": "It is demonstrated that not only gravity, but also neutrostriction forces due to optical potential created by coherent elastic neutron-neutron scattering can hold a neutron star together. The effect of these forces on mass, radius and structure of the neutron star is estimated. ", "introduction": "Interaction of neutrons with matter at low energies is characterized by optical potential \\begin{equation}\\label{1} V_o(\\rr)=\\frac{\\hbar^2}{2m}4\\pi n(\\rr)b, \\end{equation} where $m$ is neutron mass, $n(\\rr)$ is atomic density at a point $\\rr$, and $b$ is coherent s-wave neutron-nucleus scattering amplitude (see, for example~\\cite{ig0,ign}). It is important that though the amplitude $b$ is the result of short range strong interactions, the potential $V_o$ is the long range one. Neutron-nucleus scattering amplitude $b$ is of the order of several fm, the density of matter in the earth conditions is of the order $10^{23}$ cm$^{-3}$, so the optical potential of matter is of the order $10^{-7}$ eV. Interaction of neutrons with neutron matter is also described by eq. (\\ref{1}), and since the amplitude $b$ is negative, the potential (\\ref{1}) of the neutron matter is attractive. The attractive force $F_o=-\\na V_o(\\rr)$ is called neutrostriction. Neutron-neutron s-wave scattering can take place only in singlet state, and the singlet amplitude $b_s$ at low energies is $b_s\\approx -18$ fm~\\cite{huhn}. The coherent amplitude obtained by averaging over all possible spin states of two neutrons is 4 times lower, therefore $b=-4.5$ fm. For such values of $b$ at star densities larger than $10^{36}$ neutrons/cm$^{3}$ the potential $V_o$ is larger than 1 MeV. Let's compare the total gravitational and optical energies for a star of radius $R$, mass $M$ and uniform density \\begin{equation}\\label{1a} n=\\frac{M}{(4\\pi/3)R^3m}. \\end{equation} The total gravitational energy is \\begin{equation}\\label{1b} U_g=\\frac35G\\frac{M^2}{R}, \\end{equation} where $G$ is gravitational constant, while the optical energy is \\begin{equation}\\label{1c} U_{opt}=\\frac{\\hbar^2}{2m}4\\pi nbN=3\\frac{\\hbar^2}{2m^3}bM^2/R^3, \\end{equation} where $N$ is the total number of neutrons in the star. For amplitude $|b|=4.5$ fm we have $U_{opt}>U_g$, when~\\cite{conf} \\begin{equation}\\label{1d} R1$. We prove that the currently used expression for $\\Gamma$ is wrong. ", "introduction": "\\subsection{Motivations} The recent discovery of the acceleration of the Universe \\cite{1,2} has produced an overwhelming number of papers on its possible explanations, alternative to a plain cosmological constant (see Refs. \\cite{3,4,4bis} for reviews). Historically, the most favourite one is the introduction of a scalar field, usually minimally coupled with gravity (but see Refs. \\cite{4bis1,4bis2,5,6,7,8,9,9bis} for other possible models). This poses the problem of the choice of a suitable potential. Moreover, as the equations are not exactly solvable in general, there is also a problem of qualitative analysis of the solutions and of the choice of initial conditions. Usually, very crude approximations are made, first of all the socalled \\textquotedblleft slow roll\\textquotedblright\\ condition, i.e., $\\dot{\\varphi}^{2}\\ll V(\\varphi )$. The aim of this paper is to illustrate some subtleties involved by this situation, by means of a complete treatment of the general exact solutions of the Einstein equations in presence of both nonrelativistic matter and scalar field, with a suitable physical potential, and with a new treatment of the well known \\textquotedblleft tracking condition\\textquotedblright, $% \\Gamma>1$ . One result is that the slow roll condition is indeed not necessary and that insisting on it can bring to lose very interesting models; closely connected to this, is a discussion on the meaning of the statement that the state equation $w\\equiv p_{\\varphi }/\\rho _{\\varphi }$ for the scalar field should be ``almost constant''. By means of an analytic example, we shall see that this expression is highly misleading, indeed. Our choice for the potential is a particular class of exponential functions \\begin{equation} V(\\varphi)=B^{2}e^{-\\sigma\\varphi} \\,, \\label{eq1} \\end{equation} with $B$ arbitrary and $\\sigma\\equiv\\sqrt{3/2}$ (in units $8\\pi G=1$). The reason for this special choice will be clear below. First of all, we have to say that, although exponential potential is natural in higher dimension theories, supergravity and superstring models \\cite{9ter,9quater,9quinque}, it is not generally considered as feasible, even though it is often used at least as a first tool of analysis \\cite{10,12,11}. The usual objection to this potential is in fact that it cannot really be the right one, because \\textquotedblleft it scales approximately\\textquotedblright\\ like the matter component, so that, if the dark energy is dominant now, it should have been dominating also in the past, which is in contrast with BBN observations \\cite{13}. This potential was nonetheless proposed by some of us in Ref. \\cite{14}, where also some explanations of the reason why this is not true in our case are given. Anyway, in Sec. III the problem is fully explored, and we show why the above objection is wrong. A second common objection is that it describes a very peculiar situation, due to a fine tuning of initial conditions. We insist on the point that the solution found in Ref. \\cite{14} is general and exact. Here, we start indeed from a particular choice, but then we go further and release it, showing that the same \\textquotedblleft final situation\\textquotedblright\\ is reached, starting from a wide set of \\textquotedblleft initial conditions\\textquotedblright. Again, we use quotation marks because the expressions are ambiguous and misleading, so that a preliminary discussion is needed to clarify what we mean. A third possible objection is that this model could be in contrast with the observations. As a matter of fact, some papers have already appeared, successfully comparing this model with present day observations \\cite{15,16,17}. In Sec. II.B a first comparison with observational data is made, with very encouraging results. However, this paper is mainly concentrated on theoretical discussion. Our purpose is to show that the model we consider can emulate very well a standard model with dust plus a cosmological constant, well beyond any reasonable improvement of the observational techniques. In order to do this, a preliminary discussion on what is really measurable is again needed. A second important result of our considerations consists in showing that many of the current statements about tracking solutions are based on an incorrect result and should therefore be revisited. The paper is organized as follows. First, we examine some issues usually investigated when building a cosmological model with dark energy; key features like cosmic components, choice of the scalar field potential, measurability, and tracking behavior are therefore shortly sketched. Sec. II is then devoted to the exponential potential case and the general, exact solution that can be given to the cosmological equations. A comparison with a constant $\\Lambda $ model and observational data is made; also, there can be found an analysis of the scalar field equation of state and a more detailed one of the tracking behavior. In Sec. III, the features usually assigned to the tracking notion are critically examined, giving new insights into such a concept. Finally, concluding remarks are drawn in Sec. IV. \\subsection{Preliminary discussion} First of all, before going through the analysis of our potential, we think it is useful to discuss some general questions which should be of help in understanding better what follows. \\subsubsection{How can we set the choice for the potential?} At the moment there is no idea of which form the potential of the scalar field should take. The widely discussed inverse power law \\cite{4bis,20}, for instance, is certainly useful for a qualitative treatment of the problem, but only under the assumption of the slow rolling condition. It seems that the answer to this problem is simple in principle: the right form is the one that gives the best fit to the observed luminosity distance curve $d_{L}(z)$, and the fact that we cannot establish it firmly is only due to our limited observational precision. Indeed, it is possible to show that, whichever is the expression of $d_{L}(z)$, we can find, in principle, a suitable form for $V(\\varphi)$, which gives it back. Unfortunately, the value of $\\Omega_{0m}$ turns out to be \\textit{arbitrary}! The reason of this is due to the fact that the introduction of an arbitrary function in the model is equivalent to taking an infinite number of free parameters into account. Worst of all, this result is possible only if one has the \\textit{exact} functional form of $d_{L}(z)$. This is clearly impossible without already having a model in mind, so that one is led again at the starting blocks \\cite{20bis1,20bis2,20bis3,21,22}. In our opinion, the only possible way to escape from this \\textit{impasse} is to ask for potentials which have some possible explanation in terms of fundamental physics or which exhibit some nice mathematical feature, which at least leaves the hope that it is due to some still unknown fundamental physical property. As for the first requirement, to our knowledge, the best candidates are the exponential potential, like that in Eq. (\\ref{eq1}) (but with $\\sigma$ undetermined), and polynomials of the form \\begin{equation} V(\\varphi)=\\frac{1}{2}m^{2}\\varphi^{2}+\\lambda\\varphi^{4} +\\text{\\textit{% hig. ord.}} \\label{eq2} \\end{equation} It is important to observe, however, that this is a generic expansion of an even function $f(\\varphi)$, with a suitable minimum, so that any function of this type will be good as well. As for the second requirement, we shall see below that the potential proposed in Eq. (\\ref{eq1}), due to the special value of $\\sigma$, exhibits a Noether symmetry in the equations, which is just the reason why exact integration is possible. Also, the potential of Eq. (\\ref{eq2}) can be seen as the expansion of a generalized hyperbolic sine, with the same type of symmetry \\cite{23} (see below). On the other hand, it is also possible to show that a sum of exponentials of the class in Eq. (\\ref{eq1}) plus a cosmological constant is essentially the most general type which can lead to exact integration in elementary way \\cite{24}. \\subsubsection{What is really measurable?} In the context chosen by us, we have a two-dimensional configuration space, with variables $\\left\\{ a,\\varphi \\right\\} $ plus the \\textit{velocities} $% \\left\\{ \\dot{a},\\dot{\\varphi}\\right\\} $, so that the total number of variables is four. Consequently, this is the maximum number of independent quantities which can be measured. As a matter of fact, none of these quantities is in fact practically measurable. In flat cosmologies, $a_{0}$ is not measurable \\textit{in principle}. This is connected with the choice $% k=0$, i.e., in terms of measurable quantities, $\\Omega _{tot.}=1$ exactly, which is clearly artificial. Instead of these rather abstract objects, we can also use derived quantities. One of them is, of course, $\\Omega _{0m}$. This parameter is very important, not only for its physical meaning, but also because its value can, in principle, be measured independently from the SNIa observations \\cite{25}. Much has also been said about the value of the acceleration parameter $q_{0}$; but, again, it is not directly measurable in practice. However, $q_{0}$ is also present in the second term of the series \\begin{equation} d_{L}(z)=H_{0}^{-1}\\left( z-\\frac{(1-q_{0})}{2}z^{2}+...\\right) \\text{ }\\,, \\label{eq3} \\end{equation}% so that it seems possible to measure it, after all. Unfortunately, truncation of Eq. (\\ref{eq3}) after the second term is valid only at rather low redshifts (say $z\\leq 0.1$). At these values, it is practically impossible to obtain sufficient precision, while at higher values the other terms cannot be neglected. What is thus left? It seems reasonable to stick to what is for the moment firmly established, i.e., $h\\approx0.7$ and $\\Omega_{0m}\\approx0.3$, with $3\\sigma$ level $\\sim\\!10\\%$ and $\\sim50\\%$ errors, respectively\\footnote{These values are based on estimates not strictly dependent on the cosmological model.}. If we are very optimistic, we can expect an improvement of precision in the future of a factor 5, say. We shall see that even this could be not enough to discriminate among models. Also, a third useful variable could be the age of the universe $t_{0}$, for which we have only a rather poor estimate $t_{0}>12$ Gyr. \\subsubsection{Which is the precise meaning of the expression \\textquotedblleft tracking behavior\\textquotedblright ?} One of the greatest advantages of the introduction of a scalar field consists in the possibility to avoid the fine tuning of initial conditions, obtaining the same final behavior \\textquotedblleft for a wide range of initial possibilities\\textquotedblright. The last sentence is intentionally rather ambiguous. In fact, the situation is often depicted as \\textquotedblleft similar but not exactly equal\\textquotedblright\\ to that of an attractor in the theory of dynamical systems. Let us thus try to be more precise. The equations which we have to consider are the Einstein equations plus the Klein-Gordon equation for the scalar field (for sake of simplicity we write them directly in the flat case, but the following argument holds in the general case as well) \\begin{equation} 3H^{2}=\\rho _{\\varphi }+\\rho _{m}\\text{ } , \\label{eq4} \\end{equation} \\begin{equation} 2\\frac{\\ddot{a}}{a}+\\left( \\frac{\\dot{a}}{a}\\right) ^{2}+\\frac{1}{2}\\dot{ \\varphi}^{2}-V(\\varphi )=0\\text{ } \\,, \\label{eq5} \\end{equation} \\begin{equation} \\ddot{\\varphi}+3H\\dot{\\varphi}+V^{\\prime }(\\varphi )=0\\text{ } \\,. \\label{eq6} \\end{equation} Eq. (\\ref{eq4}) is a first integral of the other two. The integration constant is hidden into $\\rho _{m}$. For instance, in the case of dust, we have $\\rho _{m}=Da^{-3}$; again, the value of $D$ is not directly measurable, but we shall not in fact make use of this equation. Instead, we shall use Eqs. (\\ref{eq5}) and (\\ref{eq6}). The reason is that it can be shown that, when $\\rho _{m}$ is related to a pure dust component, they are derived by the following Lagrangian \\cite{26,27} \\begin{equation} L=3a\\dot{a}^{2}-\\frac{1}{2}a^{3} \\dot{\\varphi}^{2} +a^{3}V(\\varphi )\\text{} \\,. \\label{eq7} \\end{equation} Eq. (\\ref{eq4}) is then obtained as conservation of the energy function $% E_{L}\\equiv \\left( \\partial L/\\partial \\dot{a}\\right) \\dot{a}+\\left( \\partial L/\\partial \\dot{\\varphi}\\right) \\dot{\\varphi}-L$. A first consequence of this fact is that Liouville's theorem holds for this system. Thus, the phase space volume of possible initial conditions is conserved and has no attractors. Nevertheless, the tracking behavior can be recovered by the following argument. Consider a set of independent variables for this system (not necessarily $\\{a,\\varphi ,\\dot{a},\\dot{\\varphi}\\}$) and a volume of possible initial conditions. During the evolution, the volume is deformed, and it is possible that its projection on a two(three)-dimensional subspace converges to a point or a line. It is clear now that the result is strongly dependent on the choices of variables and projection. It is also clear that the gain in information on the variables of the subspace is compensated by a loss of information on the variables of the complement. In our specific case we can take advantage of the fact that, as said above, only a part of the possible variables is practically measurable. Therefore we can make the usual choice for the subspace, i.e., consider $\\{\\log a,\\log \\rho _{\\varphi }\\}$, and show that, for a large set of initial conditions of these two variables\\footnote{% Apparently, $a$ is constrained to be zero at the initial point, but in fact the models cannot be pushed up to the initial singularity, so that the initial time cannot be really zero. This gives some arbitrariness in the choice of initial $a$. (See next Section for details.)}, the orbits converge to a line. Another possibility is to consider $\\{H,\\Omega _{m}\\}$, and check that the orbits converge to a point. It is also clear that only \\textit{observable} quantities should be taken into account. For any other parameter the concept of fine tuning is, in our opinion, simply meaningless. The situation is in fact rather involved and will be discussed in Sec. II.D, by means of an example. We think that the concept illustrated above is the only one which is relevant for the solution of the coincidence problem. However, in the literature the term \\textquotedblleft tracking\\textquotedblright\\ is often used in a different meaning. The idea is that a large set of initial conditions should converge towards a \\textit{nearly constant} $w$, well \\textit{before} the present epoch, and maintain this condition forever. We shall show in Sec. II.C that this situation, although of course possible, is not at all necessary. ", "conclusions": "We have presented above a class of physically meaningful exponential potentials, which allow general exact solution of the Friedman and Klein-Gordon equations. We have also shown that this solution, in the limits of validity of the model, meets all the requests which are generally made for a quintessence model, except one: the possibility of switching off the acceleration in the future, which is needed for the asymptotic freedom of the model \\cite{28}. This can indeed be obtained by a modification of the potential. One possibility, studied in Refs. \\cite{31,32,33}, is to bring the value of $ \\sigma$ to the range $\\sqrt{2}<\\sigma<\\sqrt{3}$, so that the limit value for $w$ is raised to a value greater than $-1/3$. Another possibility is to pass to a combination of exponentials, which we have treated in another paper \\cite{23}. This last example is particularly interesting, and it is worthwhile to summarize some results. Using the potential \\begin{equation} V(\\varphi )=(Ae^{\\sigma \\varphi /2}-Be^{-\\sigma \\varphi /2})^{2}\\,, \\label{doubpot} \\end{equation}% it is possible to obtain essentially the same situation as shown here (with a simple exponential potential) for the past, until present epoch. Towards the future, instead, we get that $w$ oscillates forever between $-1$ and $1$, but the oscillations of $\\rho _{\\varphi }$ turn out to be damped, so that it effectively scales like matter, and dark energy never reaches complete domination of dynamics. (See Ref. \\cite{23} for details, and Ref. \\cite{33bis} for a closely related situation.) Coming back to the potential in Eq. (\\ref{eq1}), it is possible to guess a slightly positive spatial curvature. As said in the introduction, this is not excluded by the present data. It is then possible to add a curvature term into Eqs. (\\ref{eq4}), (\\ref{eq5}) and (\\ref{eq6}), and try to integrate them numerically. An appropriate choice for $\\Omega_{k}$ gives the plot of Fig. 12, where we see that the acceleration starts from negative values, reaches a regime of positive values, and then goes back to negative in the future. We have now to shortly discuss which are the shortcomings of the model we have considered so far in this paper. First of all, although exponential potentials are widely considered as possible candidates to some kind of exotic dark energy (for physical reasons which are \\textit{a priori} with respect to the framework here presented), we have found no explanation which justifies the particular class of exponential potentials we have chosen, except the nice feature of the Noether symmetry of the Lagrangian (\\ref{eq7}) (but see Refs. \\cite{26,27,29,30} for some discussion on this point). Secondly, as we have repeatedly stressed, this model loses its validity when radiation is not negligible, and we have seen that a lot of interesting features occur just during that epoch. An extension is thus needed, but this must probably be done with numerical methods. Due to the very small values of some parameters, one must in fact be very careful to avoid numerical errors. In order to generate properly the plots of Sec. II, for instance, we had to work with $40$ digits floating numbers, and there was not cumulated error typical of numerical integration. A study along these lines is presented in Ref. \\cite{33}, where they find a behavior very similar to the one depicted by us for a reasonable range of values of $\\sigma$ and initial conditions. They include radiation in numerical computation, but the procedure is somewhat different by ours. Most of all, our discussion shows not only that this model can perfectly emulate a cosmological constant, but also gives no answers about the real possibility to discriminate, at least between these two models. It seems that we have shown that this discrimination is however impossible on the pure basis of luminosity distance observations. We have obtained this already known result \\cite{34} by showing that there is an important degeneracy in the $\\Omega_{m}$ parameter. In principle, this parameter can be measured independently, but it appears very difficult to reduce its present error of a factor, say, $10$ in the near future. We have also shown elsewhere that the two models are equally compatible with the observations on the peculiar velocities of galaxies \\cite{35}. This kind of measure is also affected by large errors, is time consuming and can say something only at rather low redshift ($z\\approx 0.15)$. More promising seems to be the analysis of CMBR, as said above, and gravitational lensing, particularly of very distant QSO (see Ref. \\cite{36}, for instance, and references therein). Here the problem appears to be the presence of a lot of other parameters, leading to huge degeneracy. In our opinion, however, the main result which we obtain in this paper is probably pedagogical: we have learned that some statements, taken for granted by the majority of people, can be misleading. Let us summarize some of them here again. We have seen that the most interesting epoch in our model is when $w$ is mostly variable, but only with respect to $\\log a$. This shows that to say `` $w$ is necessarily almost constant'' is badly stated, and risks to be merely a prejudice. The study of the tracker behavior is subtle and strongly depends on the representation chosen. This is so even in the case when a general exact solution is given. This does not mean that qualitative analysis cannot be done in less favourable situations, but that it should be very carefully and unambiguously presented. The requirement $\\Gamma >1$ seems to be not necessary and, in any case, many widely accepted statements on exponential potentials and tracking solutions are indeed based on a result which has revealed to be incorrect. In conclusion, we think that the final answer to the problem of the acceptable form for the potential is far to be reached, and that more insight on the physical motivations for the dark energy potential should be developed." }, "0311/astro-ph0311584_arXiv.txt": { "abstract": "We have developed a three dimensional Monte Carlo photoionization code tailored for the study of Galactic H~{\\sc ii} regions and the percolation of ionizing photons in diffuse ionized gas. We describe the code, our calculation of photoionization, heating \\& cooling, and the approximations we have employed for the low density H~{\\sc ii} regions we wish to study. Our code gives results in agreement with the Lexington H~{\\sc ii} region benchmarks. We show an example of a 2D shadowed region and point out the very significant effect that diffuse radiation produced by recombinations of helium has on the temperature within the shadow. ", "introduction": "A major characteristic of the interstellar medium (ISM) is that it is clumpy, with filamentary structure over scales ranging from hundreds of parsecs through sub-AU cloudlets. The spatial density of electrons is approximately a power law over an astonishing 12 orders of magnitude \\citep{ars}, from $\\sim$300 pc to $<$0.1 AU in the ISM. The magnetic field of the Galaxy varies in a similar fashion on scales from 0.01 to 100 pc. Molecular clouds show a fractal (scale-free) spatial variation of the intensity of CO (e.g., \\citealt{fal}) and neutral H \\citep{sl}. Sub-parsec variations are shown by the variation of ionic column densities along different ISM sightlines towards individual stars in close binaries \\citep{lmb}. Large area velocity resolved surveys of the molecular (e.g., \\citealt{sjc}), neutral (e.g., \\citealt{hb}), and ionized \\citep{r02} gas now allow us to probe the geometry, kinematics, ionization, and temperature structure of the ISM. Our development of Monte Carlo radiation transfer techniques to study photon scattering, ionization, and radiative equilibrium dust temperatures in complex geometries is motivated by these and the many other observations that clearly require three dimensional modeling analyses. In this paper we describe improvements and extensions to our Monte Carlo photoionization code \\citep{wl} that will enable us to model the 3D temperature, ionization structure, emission line images, and spectra of low density H~{\\sc ii} regions and the Diffuse Ionized Gas (DIG), often called ``the warm ionized medium\" or ``Reynolds layer''. This diffuse ionized gas has a complicated structure and surely requires three-dimensional modeling, both as regards the sources of its ionization and the transfer of the ionizing radiation. How can Lyman continuum photons from O and B stars percolate from their midplane origins to ionize the high latitude gas? A non-uniform ISM is required that provides low density paths for ionizing photons to reach the halo. Using a ``Str\\\"omgren volume'' ionization technique and a statistical approach to clumping, \\citet{mc} showed that O and B stars can ionize the DIG (see also \\citealt{ds94}; \\citealt{dsf}). Using Monte Carlo simulations for hydrogen ionization, \\citet{c02} have investigated ionization structures in a fractal ISM as proposed by \\citet{elm}. In Elmegreen's picture, the observed H$\\alpha$ emission may arise from the ionized surfaces of fractal clouds. This suggestion clearly needs to be investigated with 3D radiation transfer techniques. Other features of the DIG that may require 3D analyses are the temperature structure and relative ionization structure of He and H. The temperature of the DIG, as determined from line intensity ratios (e.g., \\citealt{rht}), appears to require additional heating over and above that provided by photoionization heating \\citep{m00}. Among the possible heating mechanisms, dissipation of energy in a turbulent medium has been proposed \\citep{ms}, definitely 3D! Another question is, can clumping/shadowing explain the apparent lack of He$^0$ in H$^+$ regions \\citep{rt}? In this scenario, direct stellar photons are prevented from ionizing He in shadow regions behind dense clumps, but diffuse ionizing photons from H and He recombinations can percolate into the shadow regions and ionize H, but not He. Our work on the analysis of scattered light in fractal reflection nebulae \\citep{mww} highlighted the crucial role of geometry when trying to determine dust properties from scattered light observations. Much of what we know about abundances of elements comes from analysis of spectra utilizing one-dimensional photoionization techniques (see e.g. \\citealt{s02}). Over the last decade several three dimensional photoionization codes have been developed. The code presented by \\cite{bdg} uses the ``on the spot'' approximation for the diffuse radiation field, while that of \\cite{gruen} employs an adaptive mesh grid and an iterative procedure to determine the contribution of the diffuse radiation field. Monte Carlo techniques naturally include diffuse radiation within three dimensional systems and Monte Carlo photoionization codes have been described by \\citet{erc} for the three-dimensional case and \\citet{olr} for the one-dimensional case. We have extended our basic hydrogen-only Monte Carlo photoionization code \\citep{wl} to treat the photoionization, heating, \\& cooling for the complex geometry expected in low density H~{\\sc ii} regions and the diffuse ISM. Our code is similar to those described by \\citet{olr} and \\citet{erc}, but calculates the rates of photoionization and heating ``on the fly'', rather than calculating the mean intensity of the radiation field (see also Lucy 1999). It also treats reprocessing of radiation in a different manner to that described by \\citet{olr}, using ``photon packets'' rather than ``energy packets.'' In \\S2 we describe our assumptions and approximations to model the DIG. \\S3 describes the Monte Carlo technique, our method of discretizing the radiation field into ``photon packets'' rather than ``energy packets,'' and how we determine the diffuse radiation field using ratios of recombination rates. \\S4, \\S5, \\& \\S6 deal with our photoionization and heating/cooling algorithms, \\S7 presents the atomic data used in our code, and \\S8 compares our code with standard H~II region benchmarks and gives an example of a 2D calculation. ", "conclusions": "" }, "0311/astro-ph0311067_arXiv.txt": { "abstract": "The Wide-field X-ray Monitor (WXM) is one of the scientific instruments carried on the High Energy Transient Explorer 2 (HETE-2) satellite launched on 2000 October 9. HETE-2 is an international mission consisting of a small satellite dedicated to provide broad-band observations and accurate localizations of gamma-ray bursts (GRBs). A unique feature of this mission is its capability to determine and transmit GRB coordinates in almost real-time through the burst alert network. The WXM consists of three elements: four identical Xe-filled one-dimensional position-sensitive proportional counters, two sets of one-dimensional coded apertures, and the main electronics. The WXM counters are sensitive to X-rays between 2 keV and 25 keV within a field-of-view of about 1.5 sr, with a total detector area of about 350 cm$^2$. The in-flight triggering and localization capability can produce a real-time GRB location of several to 30 arcmin accuracy, with a limiting sensitivity of $10^{-7}$ erg cm$^{-2}$. In this report, the details of the mechanical structure, electronics, on-board software, ground and in-flight calibration, and in-flight performance of the WXM are discussed. ", "introduction": "\\label{sect:intro} The origin and nature of Gamma-ray bursts (GRBs) has been one of the outstanding mysteries of astrophysics since the phenomenon was discovered in the 1970s. The discovery of GRB afterglows by the BeppoSAX X-ray satellite~\\citep{1997Natur.387..783C} produced a breakthrough in the study of GRB. Since the discovery, follow-up observations of many GRB afterglows have been successfully performed. The observations have furnished strong evidence that at least the long class of GRB originates from explosions of super-massive stars at cosmological distances. The HETE-2~\\citep{wh.2003.grr} is the first satellite dedicated to GRB observations and is intended to provide the astronomical community with prompt, high-accuracy GRB positions --- a few tens of arcminutes to a few tens of arcseconds, with delay times of 10 s to a few hours. This capability not only increases the number of detected GRB counterparts, but also provides an opportunity for multi-wavelength observations of GRBs in their very early phase. Such observations were nearly impossible prior to the HETE-2 era. The HETE-2 instrument complement consists of three scientific instruments: the French Gamma Telescope (FREGATE), the Wide-field X-ray Monitor (WXM), and the Soft X-ray Camera (SXC). FREGATE~\\citep{wh.2003.atteia} consists of four identical scintillation detectors. It is sensitive to photons with energies from 6 to 400 keV. It is responsible for triggering and spectroscopy in the hard X-ray and Gamma-ray energy ranges. Its field of view is 70$^{\\circ}$ (half width at zero maximum) and its total detection area is 160 cm$^2$. The SXC~\\citep{wh.2003.joel} is a 1-D coded mask system using MIT-LL CCID-20's as the detecting elements, and is sensitive to photons with energies from 500 eV to 14 keV, with 2\\% resolution at 6 keV and with a position resolution of 10$''$ for bright bursts (1 Crab; 10 s). SXC's total detection area is 75 cm$^2$ and its field of view is $65^{\\circ}$ (full width at zero maximum). The WXM, on the other hand, has a detection area of 352 cm$^2$ and a field of view of $80^{\\circ}$ (FWZM). Thus, the WXM is more sensitive to weak bursts than the SXC, and the two X-ray instruments play a complementary role to each other, balancing detection sensitivity against location resolution. ", "conclusions": "We have calibrated the alignment between the WXM and the HETE-2 spacecraft aspect system. We estimate the systematic uncertainty of WXM locations of GRB to be 2$'$ for each X and Y direction at the 68\\% level. The corresponding two-dimensional systematic error is 4$'$ (90\\% confidence level). We calibrated the gain scale of the proportional counters using observations of the Crab Nebula and using the signal from the calibration radio-isotopes. At present, the systematic uncertainties in fitting a power-law spectrum model to the Crab spectrum are: $\\pm 0.1$, $\\pm 0.2\\times 10^{22}$ cm$^{-2}$ and 10\\% for power-law index, absorption column density, and absolute flux, respectively. Improvements in the spectral calibration are in progress. The in-orbit flight localization performance of the WXM was examined based on XRB observations and Monte Carlo simulations. The sensitivity of the flight localization is estimated to be $> 10^{-7}$ erg~cm$^{-2}$, and the estimated false alert rate of real-time localizations is 20\\% for $1.5\\times 10^{-7}$--$2 \\times 10^{-6}$ erg~cm$^{-2}$. Based on the recent GRB localizations, the rate of GRB localization is 28 per year, and the real-time localization rate is approximately one per month." }, "0311/astro-ph0311251_arXiv.txt": { "abstract": "{ In this paper we present Very Large Array (VLA) 1.4 GHz (21 cm) observations of the region between the centres of A3558 and A3562, in the major cluster merger complex of the Shapley Concentration. Our final catalogue includes a total of 174 radio sources above the flux density limit of 0.25 mJy b$^{-1}$. By cross-correlation with optical and spectroscopic catalogues we found 33 optical counterparts belonging to the Shapley Concentration. \\\\ We investigated the effects of cluster merger on the radio emission properties of the galaxy population by means of the radio source counts and the radio luminosity functions (RLF). We found that the radio source counts are consistent with the field source counts. The RLF of elliptical and S0 galaxies in the region surveyed here, is consistent with the ``universal'' RLF for early--type galaxies. This result suggests that the deficit in radio galaxies found in our previous works over the whole A3558 chain, is entirely due to the cluster A3558. A population of faint radio galaxies (logP$_{1.4~GHz}$(W Hz$^{-1}$) \\ltsim 22) is also found. Half of these objects are also blue, suggesting that starburst is the main mechanism driving the radio emission. Finally, we detected 14 spiral radio galaxies, whose ratio between radio and optical emission is similar to those found in galaxies located in rich and dynamically evolved clusters. \\\\ Our results are briefly discussed in the light of the age and stage of the merger in the A3558 cluster complex. ", "introduction": "Evidence is accumulating that the interaction processes between clusters of galaxies, known as cluster mergers, may significantly affect the radio emission characteristics of the cluster galaxy population. In particular, the evolution of galaxies and the properties of their nuclear and/or star forming activity are undoubtely influenced by the interaction with the environment. In this frame, cluster merging and group accrection seem to play an important role, but it is not yet completely understood how the merging environment affects the nuclear and starburst emission in galaxies. Owen et al. (\\cite{owen99}) suggested that merging may trigger the radio emission, both in the form of nuclear activity and starburst phenomena. Burns et al. (\\cite{burns94}) interpreted the presence of post-starburst galaxies in the X--ray and radio bridge connecting the Coma cluster and the NGC 4839 group as consequence of a recent merging process between these two structures. Using numerical simulations, Bekki (\\cite{bekki99}) concluded that the tidal gravitational field of a cluster merger may drive a considerable transfer of gas to the central regions of galaxies, leading to enhanced star formation activity or feeding the central engine of active galactic nuclei. Vollmer et al. (\\cite{vollmer01b}) suggested that a local burst of star formation could be due to re-accretion of gas after a ram pressure stripping event, which is thought to be responsible for the HI deficit of spiral galaxies in the central regions of nearby clusters with respect to the field galaxies of the same morphological type and optical size (Gunn \\& Gott, \\cite{gunn72}; Bothun \\& Dressler, \\cite{bothun86}). On the other hand, Fujita et al. (\\cite{fujita99}) and Balogh et al. (\\cite{balogh98}) claimed that gas stripping resulting from ram pressure is important in preventing gas supply to the central regions of the galaxies and may suppress star formation. Finally Venturi et al. (\\cite{venturi00}, hereinafter V2000) found evidence that merging may anticorrelate with the radio emission, possibly switching off previously existing radio galaxies, or temporarily inhibiting the nuclear radio activity. \\\\ A major problem in our current knowledge of the role of cluster mergers on the radio emission and in the interpretation of the observations is the lack of statistics. Only few clusters of galaxies were deeply imaged over their whole extent (i.e. A2125, Dwarakanath \\& Owen, 1999; A2255, Miller \\& Owen, 2003 (hereinafter MO03); A2256, Miller et al., 2003; A2645, Owen et al., 1999; the Shapley supercluster chains A3528, A3558 and A3571, Venturi et al. 2001, 2000 and 2002 respectively), and the complexity of the phenomenon clearly needs a much larger observational support. Cluster mergers evolve over a timescale of $\\sim$ Gyr, while typical ages of radio sources are of the order of few times 10$^7$ -- 10$^8$ years, therefore it is of crucial importance to identify the effects of the various stages of a merger event. For this reason it is essential to carry out deep radio observations of a large number of galaxy clusters, at different merger stages. \\\\ In order to study the effects of a major cluster merger event after the first core--core encounter on the nuclear and starburst radio emission, in this paper we present 1.4 GHz Very Large Array (VLA) observations of the cluster A3562 and the poor groups SC 1329$-$313 and SC 1327$-$312, at the eastern side of the A3558 cluster complex. In the following we will refer to the area surveyed as A3558--C. This work is part of a larger project which aims to study the influence of the ongoing merger in the core of the Shapley Concentration on the radio/optical properties of the cluster galaxies, both from a statistical point of view and through a detailed analysis of the physical properties of the extended radio galaxies. In section \\ref{sec:intro} we briefly overview the optical and X--ray properties of A3558--C, with emphasis on the cluster A3562 and on the poor groups SC 1329$-$313 and SC 1327$-$312; the observations and data reduction are presented in Section \\ref{sec:obs}; in Section \\ref{sec:sample} we present the 1.4 GHz radio sample and the source counts; in Section \\ref{sec:ident} we deal with the optical identifications and in Section \\ref{sec:prop} we present the general properties of the radio galaxies in this region; the results of the analysis of the radio luminosity function (RLF) for the early-type and late-type galaxy populations are given in Section \\ref{sec:RLF}; discussion and conclusions are given in Section \\ref{sec:disc}. We assume $H_0 = 100$ km s$^{-1}$ Mpc$^{-1}$ and $q_o$=0.5. If we define h=H$_o$/100, at the average redshift of the Shapley Concentration ($z=0.05$) this leads to a linear scale of 1 arcsec = 0.67 h$^{-1}$ kpc. We will assume $S \\propto \\nu^{-\\alpha}$. \\vskip 1 truecm \\noindent ", "conclusions": "\\label{sec:disc} In this paper we presented deep 1.4 GHz VLA observations of the major cluster merger in the A3558 complex. The area under study, A3558--C, defined as the region between the centres of the two Abell clusters A3558 and A3562, including also the two groups SC 1329--213 and SC 1327--312, is thought to have recently experienced a major merger between two massive clusters. What we see now is expected to be the result of the first core--core encounter. The observational properties in this region, together with numerical simulations, suggest that A3558 is the main cluster, while the whole chain, beyond A3558 itself, is the remains of the colliding cluster (Bardelli et al. 2002). \\\\ A picture is emerging, in which radio AGN and starburst activity, radio relics and halos, or the lack thereof, are signature of cluster mergers at different stages. In particular, Venturi et al. (2002) proposed an evolutionary merger sequence to account for the diversity in the radio properties of the three main merging cluster systems in the core of the Shapley Concentration; Kempner \\& Sarazin (2001) postulated that radio halos and relics may form at different times during mergers; MO03 explained the different fraction of radio emitting galaxies (starburst and AGNs) in a cluster sample as due to different merger stages. With this study we have a unique opportunity to connect the observed properties in the radio band with a well defined cluster merger stage. This is of crucial importance for a better understanding of the complex effects of cluster mergers, with particular emphasis to the stage of the merger. \\\\ The most relevant results of our analysis can be summarised as follows: \\begin{itemize} \\item[(i)] a faint radio halo is found at the centre of A3562, whose properties are consistent with the idea that it is a young source at the beginning of the reacceleration phase (V2003) induced by a recent merger event; \\item[(ii)] the origin of the extended emission in the radio galaxy J1332--3146a, associated with the dominant galaxy in the group SC 1329--313, is unclear. No radio jets are present in the nuclear region of J1332--3146a, and the nuclear radio component has steep spectrum. It is possible that this extended emission has actually cluster origins (see Section 6.2), being either (a) a ``bright'' area of a very low brightness bridge of radio emission, connecting A3562 and SC 1329--313, or (b) a ``revived'' radio emission region, where pre--existing old electrons were reaccelerated; \\item[(iii)] a large number of radio sources associated with A3558--C galaxies was found, i.e. 33 objects, most of them with radio emission at low power levels. Our analysis (see Section 6.3) suggests that 26/33 radio galaxies are candidate starbursts. Among them, 11 also show a blue excess; \\item[(iv)] the total number of radio AGNs detected in this region is consistent with the expectations from the RLF of LO96, suggesting that the cluster merger has not affected the probability of an early--type galaxy to develop a nuclear radio source; \\item[(v)] the distribution of the radio/optical ratios for the spirals in A3558--C is similar to what is found in rich and evolved environments. \\end{itemize} The main question is if and how this wealth of observables is connected to the cluster merger in this region. Beyond the noticeable finding of the radio halo in the centre of A3562 (V2003), the radio properties of the Shapley galaxies in this region may contain important pieces of information. V2000 showed that the RLF for early type galaxies in the whole A3558--C shows a deficit of radio galaxies compared to the ``universal'' RLF presented in LO96 (V2000), over the whole power range. On the other hand, the RLF for AGNs presented here matches the expectations of LO96; furthermore the outskirts of the chain (A3556 and A3562) contain the largest fraction of Shapley radio galaxies (see also Section 2). This suggests that A3558 itself, the most massive cluster in the chain, is the main responsible for the lack of radio sources in the RLF found in V2000. A possible explanation is the key role of A3558 in the merger, i.e. it is experiencing the most dramatic effects of the merger, being the result of the interaction of core regions of the two colliding clusters. If our interpretation is correct, then the role of cluster merger on the radio emission from AGNs may be many--fold, depending on the age and strength of the merger. \\\\ Optical photometric information on the faint population of radio galaxies in A3558--C confirms that at least 50\\% of these objects are most likely starburst candidates. However, a radio/optical analysis carried out following MO03 provides only weak evidence that the fraction of radio emitting galaxies in A3558--C is higher than in non--merging environments. \\\\ The statistical results on the late--type galaxies show that the radio emission in the A3558--C spirals is similar to those in rich and dynamically evolved clusters. This suggests that the radio emission in spiral galaxies may be one of the first ``parameters'' to react to a cluster merger event. In conclusion, V2003 showed that the radio halo at the centre of A3562 is consistent with a reacceleration phase which started $\\sim$ a few $10^8$ years ago. This ongoing merger is therefore {\\it advanced}, in the sense that the core--core encounter has already taken place, but it is still young if compared to the total duration expected for a cluster merger, i.e. 10$^9$ yr. Our study suggests that on this timescale, the effect of cluster merger on the radio emission from cluster galaxies is many--fold. In particular, we found only marginal evidence of enhanced radio emission of starburst origin, but we found significant enhanced radio emission from spirals. We argue that the role of cluster merger on the nuclear activity in early--type galaxies is a complex phenomenon, since observational evidence in the whole A3558 cluster chain clearly shows that the deficit in radio galaxies found in V2000 is entirely due to the cluster A3558, which has experienced the most violent consequences of the merger." }, "0311/astro-ph0311366.txt": { "abstract": "\\noindent We use a principal components analysis of radio-selected (3CRR, 6CE and 7CRS) AGN datasets to define two parameters related to low-frequency ($151$ MHz) radio luminosity $L_{151}$ and [OIII] luminosity $L_{{\\rm [OIII]}}$: a parameter $\\alpha$ encoding the $L_{151} - L_{{\\rm [OIII]}}$ correlation and a parameter $\\beta$ encoding scatter about this correlation. We describe methods for constructing generalized luminosity functions (GLFs) based on $\\alpha, \\beta$, redshift and schemes for unifying quasars and radio galaxies. These GLFs can be used to generate radio luminosity functions (RLFs) which improve on those of \\citet{wrlf}, mostly because they incorporate scatter and are therefore much smoother. Luminosity-dependent unified schemes (e.g. a receding-torus scheme) have been invoked to explain the low quasar to radio galaxy fraction at low $\\alpha$ and the differences in emission-line luminosities of radio quasars and radio galaxies. With the constraints of the 3CRR, 6CE and 7CRS datasets and radio source counts, our GLF approach was used to determine whether a receding-torus-like scheme is required if there are two populations of radio sources: one at low $\\alpha$, consisting of `starved AGN'; the other at high $\\alpha$ consisting of `Eddington-tuned AGN'. Because of the overlap between these two populations and the effects of the $\\beta$ parameter, schemes with or without a receding torus can produce a low quasar fraction at low $\\alpha$ and differences in [OIII] luminosity between radio galaxies and quasars. The receding torus may be a physical process important in one or more populations of radio sources, but this is not yet proved either by the quasar fraction or the emission-line properties of radio-selected samples. ", "introduction": "\\label{sec:intro} It has been recognised for some time that there is a strong positive correlation between the extended-radio luminosities and narrow-emission-line luminosities of 3C radio sources \\citep{bh, retal, mcarthy}, and by combining the 3C sample with the 7C Redshift Survey (7CRS), \\citet{wem} found that these correlations were not primarily due to redshift. This suggests that the sources of narrow-line emission and radio jets are linked, with explanations for this link ranging from the effects of environment \\citep{dp93}, to jet-disk symbiosis \\citep{rs91, fb}. Unified schemes for radio galaxies and radio quasars propose that they are the same objects viewed at different angles between their radio axis and their line of sight. An obscuring torus is invoked to hide the nucleus at large angles to the jet axis i.e. in radio galaxies. Emission from narrow-line regions is believed to be broadly independent of the jet-axis orientation, as it is emitted from a region larger than the extent of the torus. The opening angle of the torus $\\Theta_{\\rm trans}$ marks the transition from the object being viewed as a quasar to a radio galaxy. We therefore expect for simple, arguably naive, unified schemes, where $\\Theta_{\\rm trans}$ is constant, that the distribution of emission-line luminosities should be similar for radio quasars and radio galaxies. \\citet{jb} found that [OIII] luminosities of radio quasars are $\\sim 5-10$ times more luminous than radio galaxies. In similar comparisons of radio quasars and radio galaxies, \\citet{hbf} found no difference using [OII], and \\citet{jr} found no difference in [OIII] at high redshift. In summary, some studies comparing the narrow emission line strengths of radio quasars and radio galaxies are seemingly in agreement with the predictions of the simplest unified schemes, whilst others are not. These seemingly contradictory, results can perhaps be understood in the context of a `receding torus' model for AGN. This model \\citep{law, hgd} proposes that the inner radius $r$ of the obscuring torus is determined by the radius at which dust sublimates, scaling as $r \\propto L_{\\rm phot}^{0.5}$. Assuming that the half-height $h$ of the torus is independent of $L_{\\rm phot}$, then the half-opening-angle of the torus, $\\theta = \\tan^{-1} r / h$, will be larger in the higher-$L_{\\rm phot}$ objects. This means that more luminous objects, with higher $L_{\\rm phot}$, are more likely to be viewed as quasars so that quasar fraction rises systematically with $L_{\\rm phot}$. It also means that orientation-independent quantities which scale with $L_{\\rm phot}$ will be higher on average for quasars than for radio galaxies. \\citet{law} found that the fraction of narrow-line objects in the low-frequency-selected 3CR sample decreased with increasing radio luminosity, and that narrow-line objects have weaker [OIII] at a fixed radio luminosity. He argued that this was inconsistent with the simplest unified schemes but that it could be explained by a cone-angle dependence on luminosity, e.g a receding-torus model. \\citet{simp} argued that the [OIII] emission line is a much better indicator of $L_{\\rm phot}$ than [OII], explaining why differences between quasars and radio galaxies are more obvious in [OIII] than [OII]. \\citet{sconf} revisited the arguments of \\citet{simp} and corrected a small error (compare Fig.~2 of \\citealp{sconf} with Fig.~4 of \\citealp{simp}). He concluded that, allowing for a spread in $h$, the receding torus model predicts that radio quasars should be a factor of a few brighter in $L_{\\rm phot}$ than radio galaxies in samples exhibiting a wide range of quasar fractions. \\citet{wsubmm} found that radio quasars have higher submillimetre luminosities by a factor of $\\sim 4$ than radio galaxies of the same radio luminosity and redshift, a factor which cannot be reduced below $\\sim 2$ by accounting for possible synchrotron contamination. This result supports the idea that the simplest unified schemes, where $\\Theta_{{\\rm trans}}$ is constant, are not an adequate description of the relationship between radio galaxies and quasars: submillimetre emission comes from cool dust grains in optically thin regions, and therefore radiates isotropically; moreover, such emission could not be obscured by a torus even if it were emitted from regions close to the nucleus. Submillimetre luminosity is therefore an orientation-independent quantity which might scale closely with $L_{\\rm phot}$ either because the cool dust is heated directly by the quasar or because it is heated by a starburst whose luminosity scales with $L_{\\rm phot}$. We conclude that the submillimetre study of \\citet{wsubmm} is in quantitative agreement with the receding torus model of \\citet{sconf}. Other arguments for a receding-torus-like model include the evidence that the fraction of lightly reddened $z \\sim 1$ 3C quasars decreases with increasing radio luminosity in agreement with the higher fraction of lines of sight expected to graze the torus at lower $L_{\\rm phot}$ (at fixed $h$) in the receding torus model (\\citealp*{hgd}; \\citealp{srl}). Also the near-infrared spectral indices of quasars from the 3CR sample are correlated with luminosity, whereas the optical spectral indices are uncorrelated with the quasar luminosity or orientation, so that the strength of the `big red bump', relative to the ionizing continuum, appears to be less in the more luminous objects \\citep{sr,sconf}. However, there is significant scatter in the relationship between radio and emission-line luminosities and a receding torus is not the only way of explaining the differences in the emission-line properties of radio quasars and radio galaxies or the luminosity-dependence of the quasar fraction (the fraction of objects that show broad emission lines). Two-population radio luminosity functions (RLFs) have been used to provide a best-fit to the 3CRR, 6CE and 7CRS radio source redshift surveys and radio source counts \\citep{wrlf}. It is possible that a two-population model with a simple unified scheme in one population, combined with the effects of scatter, could mimic the effects of a receding torus in producing both emission-line differences between radio galaxies and radio quasars and the gradual rise in quasar fraction with emission-line luminosity. \\citet{wqf} found a drop in the quasar fraction of AGN at low luminosity, postulating that their results are consistent with either the emergence of a second population of low-luminosity radio sources, which lack a well-fed nucleus, or a gradual decrease in $\\Theta_{{\\rm trans}}$ with decreasing radio luminosity. Traditionally, radio galaxies have been divided into two populations, FRI and FRII, \\citep{fr}, based on radio structure but correlated with radio luminosity such that above $\\log_{10}(L_{151} / {\\rm W\\,Hz^{-1}\\,sr^{-1}}) \\sim 25$, objects are FRII, whereas below that critical $L_{151}$, they are FRI. \\citet{ccc02} found that the FRII population is inhomogeneous and that not all of them can be unified with quasars. They find a low-radiative-efficiency accretion, weak or absent broad-line emission and a lack of a significant nuclear absorbing structure for weak-jet, low-ionisation narrow-line galaxies. For broad-line objects and obscured high-ionization narrow-line galaxies they see or infer intense ionizing emission, powerful jets and a torus-like absorber. In low-radio-frequency-selected samples, the fraction of objects with observed broad lines changes rapidly from $\\sim 0.4$ above $\\log_{10}(L_{151} /{\\rm W\\,Hz^{-1}\\,sr^{-1}}) \\sim 26.5$ to $\\sim 0.1$ below this. The less luminous population from \\citet{wrlf} is composed of FRIs and FRIIs with weak/absent emission lines and their more luminous population of strong-emission-line FRII radio galaxies and quasars. A two-population model like this is motivated by the fact that the presence or absence of a quasar nucleus (as shown by emission-line strength) seems likely to be connected to the properties of a central engine of the radio source. \\citet{ka97}, and others before them, have argued that radio structure is influenced by the environment. It seems likely that a high jet power is necessary for a highly collimated non-dissipative jet, thus all high luminosity sources are FRII, but as lower jet powers are reached, a jet is more likely to disrupt resulting in a FRI source. The exact environmental density and homogeneity would determine the radio luminosity threshold of FRI/FRII within the lower luminosity population. \\citet{lo96} found that the FRI/FRII division is proportional to the square of the optical luminosity of the host galaxy, which is plausibly related to such environmental effects, although other causes have been proposed. This simple picture is shown to be incomplete by the discovery by \\citet{br} of an optically powerful quasar with FRI radio structure. It also seems that there is evidence that some FRI sources in 3C have hidden quasar nuclei \\citep{cao}, and one clear example of a 3CR FRI with broad lines (3C386, \\citealp{s96}) is known. In this paper we follow \\citet{hl}, \\citet{laing} and \\citet{jw} in assuming that the FRI/FRII division is strongly influenced by environmental effects and is much less fundamental than a division based on accretion properties of the central engines. It is well known (e.g. \\citealt{dp90}) that the evolution in comoving space density of radio galaxies depends on both $L_{151}$ and $z$. In two-population models this can be expressed as differential density evolution of these populations \\citep{wrlf}. \\citet{sb} have criticized this approach on the basis of the artefacts it produces when the populations join. The primary motivation of this work was to investigate in a quantitative manner the differences between the emission-line properties of radio-loud quasars and radio galaxies alongside the luminosity dependence of the quasar fraction. In this paper we will quantify the three-dimensional distribution in [OIII] emission-line luminosity $L_{{\\rm [OIII]}}$, $151$ MHz luminosity $L_{151}$ and redshift $z$ using a new method applied to complete samples. A principal components analysis (PCA) is used to find the axis which causes the most differentiation between the objects, and it is possible that it can identify parameters which are more physically meaningful than the attributes of each data point (radio and emission-line luminosity) used in the PCA. We carry out a PCA and define a generalized luminosity function (GLF) in this new parameter space. The GLF is then used to make proper comparisons between receding-torus and two-population models, allowing for cosmic evolution of the populations. The paper is organized as follows. In Sec.~\\ref{sec:data} the data are described. In Sec.~\\ref{sec:pca}, the principal components analysis is described. In Sec.~\\ref{sec:model}, simple one-population GLFs are described to enable a comparison between unified schemes with and without a receding torus. The effects of using two-population GLFs are presented in Sec.~\\ref{sec:2pop}, and in Sec.~\\ref{sec:discussion} we compare our results with previous studies and discuss possible meanings for the new parameters found by the PCA. We assume throughout that ${\\rm H_0} = 70\\,{\\rm km\\,s^{-1}\\,Mpc^{-1}}, \\Omega_{\\rm M} = 0.3$ and $\\Omega_{\\Lambda} = 0.7$. ", "conclusions": "A new approach to investigating unified schemes has been presented, based on a principal components analysis of the 3CRR, 6CE and 7CRS complete samples. Generalized luminosity functions have been derived based on these principal components: $\\alpha$ which encodes the $L_{151} - L_{\\rm [OIII]}$ correlation and $\\beta$ which encodes the scatter about this correlation. The main advantage of this new approach has been that the unified scheme parameters have been found by taking into account the intrinsic scatter in the $L_{151} - L_{\\rm [OIII]}$ correlation. The main conclusions to be drawn from this analysis are as follows. \\begin{enumerate} \\item A receding-torus model is strongly favoured over a simple non-luminosity-dependent unified scheme for GLFs with one population of radio sources. \\item With the extra constraint of 6C and 7C radio source counts and the normalization of the local RLF, two-population GLFs were derived. The GLFs give rise to RLFs which are very similar but smoother than those of \\citet{wrlf}. There is very little difference in likelihoods between a GLF with a receding torus in both populations, a GLF with a torus opening angle that does not vary with ionizing luminosity in both populations, and a GLF with a receding-torus in the high-$\\alpha$ population and a constant-opening-angle torus in the low-$\\alpha$ population. \\item Two-population models reproduce the radio survey data well, and can provide a natural explanation for the rise in quasar fraction with emission-line luminosity and the emission-line differences between radio quasars and radio galaxies. The receding-torus may be a feature in both populations but this is not yet proved. \\end{enumerate}" }, "0311/astro-ph0311414.txt": { "abstract": "This article gives an analysis of the behavior of polarizing grids and reflecting polarizers by solving Maxwell's equations, for arbitrary angles of incidence and grid rotation, for cases where the excitation is provided by an incident plane wave or a beam of radiation. The scattering and impedance matrix representations are derived and used to solve more complicated configurations of grid assemblies. The results are also compared with data obtained in the calibration of reflecting polarizers at the Owens Valley Radio Observatory (OVRO). From these analysis, we propose a method for choosing the optimum grid parameters (wire radius and spacing). We also provide a study of the effects of two types of errors (in wire separation and radius size) that can be introduced in the fabrication of a grid. ", "introduction": "The literature on wire grids is abundant and they have been studied with different techniques and for numerous applications. Most of the analysis were however restricted to special cases of incident field and grid orientations. The more general and arbitrary situation seems to have been first studied by Wait (see \\citet{Wait 1955a} and \\citet{Larsen}). This problem is addressed again in this paper and follows a line of analysis fairly similar to the one used by Wait. Our treatment is, however, more general in that we do not assume that the wires of the grid are induced with only a longitudinal current; we will indeed show that an azimuthal component is also present. We also solve for the induced current by considering the tangential components of both the electric and magnetic fields at the surface of the wires. This analysis is carried out in the next two sections and will serve as our basis for the treatment of the reflecting polarizer (section \\ref{sec:RP}) and the introduction of the scattering and impedance matrix representations for a grid (section \\ref{sec:imp}) which will in turn enable us to briefly discuss more complicated systems. These matrices will be particularly useful in allowing us to define what will be called the principal axes of a grid. These are two orthogonal and independent directions of polarization in the plane of the incident radiation along which an arbitrary electric field can be decomposed and shown to scatter without cross-polarization. With this representation at hand, it will then be possible to derive a set of optimal parameters (wire radius and spacing) to be used in the selection of a grid. We will also present an analysis of the effects of random errors that can be introduced in the fabrication of grids, the results obtained will then be compared to experimental results previously published by \\citet{Shapiro}. The last section will be dedicated to the study of the more subtle impacts that the nature of the incoming radiation can have on the response of a grid assembly such as a reflecting polarizer (section \\ref{sec:RP}). Although limited to this particular case, our discussion could possibly apply to other types of instruments. We have also included at the end (Appendix B) a list of the symbols used in the different equations. ", "conclusions": "In this paper, a general solution for the analysis of polarizing grids was presented; it is valid for arbitrary angles of incidence and of grid rotation. With it and the scattering matrix representation that derives from it, basically any configuration or system of grids can be analyzed as long as some assumptions concerning the wire radius and spacing are respected ($\\lambda>40\\, a$ and $d>4\\, a$). This is not a severe restriction as most grid currently available satisfy those conditions, we refer the reader to \\citet{Chambers 1986, Chambers 1988} for cases where a larger size of wire is needed. Our analysis also allowed us to define a set of optimum values for both the wire radius and spacing as specified by the following equations: \\begin{eqnarray*} a & \\simeq & \\left[\\frac{\\lambda^{5}}{\\left(1-\\alpha^{2}\\right)^{4}\\pi^{7}\\sigma Z_{o}}\\right]^{\\frac{1}{6}}\\\\ d & \\simeq & 2\\pi a\\,.\\end{eqnarray*} We provided an analysis of the effects that two types of random errors can have on the performance of a grid. It was shown that errors in the wire spacing were the most important and could have some impact on the amount of unwanted polarization transmitted through a grid. In that respect, our model showed to be in good agreement with the experimental results of \\citet{Shapiro}. Comparisons with experimental data obtained in the calibration of a reflecting polarizer used at the OVRO were also presented and predictions from our model are in good agreement with it. The only discrepancies appeared in the nature of a resonance, more precisely its width. But we have shown that it could be accounted for by including in the analysis a proper treatment of the effects of the nature of the incident radiation on the response of the polarizer. We are grateful to J. B. Shapiro and E. E. Bloemhof for their permission to use their previously published experimental results. We wish to thank the staff of the Owens Valley Radio Observatory, and O. P. Lay for numerous discussions and suggestions. The Owens Valley Radio Observatory is funded by the National Science Foundation under Contract No. AST 96-13717 and the polarimetry project at OVRO through NASA grant NAG5-4462. M. H. work was supported in part by a grants from FCAR and the d\u00e9partement de physique de l'Universit\u00e9 de Montr\u00e9al. \\pagebreak \\appendix" }, "0311/astro-ph0311317_arXiv.txt": { "abstract": "We have used high resolution {\\it WFPC2-HST} and wide field ground-based observations to construct a catalog of blue straggler stars (BSS) in the globular cluster 47 Tuc spanning the entire radial extent of the cluster. The BSS distribution is highly peaked in the cluster center, rapidly decreases at intermediate radii, and finally rises again at larger radii. The observed distribution closely resembles that discovered in M3 by Ferraro et al (1993,1997). To date, complete BSS surveys covering the full radial extent (\\hst\\ in the center and wide field CCD ground based observations of the exterior) have been performed for only these two clusters. Both show a bimodal radial distribution, despite their different dynamical properties. BSS surveys covering the full spatial extent of more globular clusters are clearly required to determine how common bimodality is and what its consequence is for theories of BSS formation and cluster dynamics. ", "introduction": "\\label{sec:intro} Blue straggler stars (BSS) are found primarily in star clusters, where they appear as a sparsely populated extension of the main sequence above the turnoff point. Superficially they appear to be main sequence stars with masses larger than expected for the cluster at its turnoff age. There are two proposed mechanisms thought to produce BSS: the first is mass exchange in a binary system and the second is the merger of two stars induced by stellar interactions (either single or in binaries) in a dense stellar environment. Globular cluster cores are obvious targets for BSS not only because of their high stellar density, but because primordial binaries in the clusters may well have sunk to the cores. With the advent of the Hubble Space Telescope (\\hst) it became possible for the first time to search dense cluster cores for BSS. Searches in the ultraviolet (UV) can be particularly effective, not because BSS are especially bright in the UV, but because the red giants are faint in the UV. The photometric blends which mimic BSS in visible CMDs are far less problematic in the best UV CMD planes, and it is possible to obtain complete BSS samples in even the densest cores. With \\hst, the densest cluster cores are {\\it the obvious place} to search for BSS, and there have been few systematic searches for BSS in the outer parts of clusters in recent years. Indeed only one cluster, M3, has been adequately surveyed for BSS over its entire radial extent \\citep{f93,f97}. Since it is a massive cluster with high central density, 47~Tuc is an obvious target for a BSS search, and its core has been the subject of several earlier investigations \\citep{gua92,demarchi47bss,ferraro47bss,knigge47}. These studies have covered only a fraction of the core either because they used the Faint Object Camera or just the PC of WFPC2. Albrow et al. (2001) found BSS in their WFPC2 search for eclipsing biniaries in the core of 47~Tuc. Blue stragglers have also been studied outside the core of 47 Tucanae by Sills et al. (2000) who also modeled the BSS formation rate. Kaluzny et al. (1998) note several BSS candidates in their CMD of 47~Tuc, and there are clear BSS candidates in the Str\\\"omgren photometry of Grundahl et al. (2002). Only a small part of the cluster exterior was covered by these studies. With new wide field imagers on ground based telescopes at sites with excellent seeing it is now possible to perform BSS surveys which yield complete samples over the entire extent of the cluster. Here we present results covering both the full cluster core using \\hst\\ WFPC2, and the exterior of the cluster using the wide field imager at the ESO 2.2\\,m telescope. ", "conclusions": "F97 argued that the bimodal distribution of BSS in M3 was evidence that two formation scenarios were at action in the same cluster: the Exterior BSS (EBSS) arising from mass transfer in primordial binaries and the Central BSS (CBSS) arising from stellar interactions which lead to mergers. As had been earlier noted by Bailyn \\& Pinsonneault (1995), the EBSS and CBSS luminosity functions differed in the sense theoretically expected for two mechanisms. One difficulty with primordial binaries as a source for the EBSS is that one ordinarily would expect that they would have settled to the cluster centers since relaxation times are typically less than 1~Gyr \\citep{djor-physparm}. Simulations to address the question of how many primordial binaries can form in the external regions of the cluster and how long primordial binaries with mass approaching twice the turnoff mass could remain in the outer parts of a cluster are now in progress (M. Mapelli et al 2003, in preparation). Sigurdsson \\etal\\ (1994) offered another explanation for the bimodal BSS distribution in M3. They suggested that the EBSS were formed in the core and then ejected into the outer regions by the recoil from the interactions. Those binaries which get kicked out to a few $r_c$ rapidly drift back to the center of the cluster due to mass segregation, leading to a concentration of BSS near the center and a paucity of BSS in the outer parts of this region. More energetic kicks will take the BSS to larger distances; these stars require much more time to drift back toward the core and may account for the overabundance of BSS at large distance. How does the discovery of a bimodal radial distribution of BSS in 47~Tuc affect these arguments? Dynamically 47~Tuc is quite different from M3: its central density is 40 times larger \\citep{pm93}, and its core is rich in X-ray binaries, millisecond pulsars, and other interaction products \\citep{ferraro47bss,grindlay01, camilo00,e03}. We have already shown \\citep{bss6} that CBSS populations cannot be characterized by a simple parameter like collision rate, so the fact that CBSS in 47~Tuc and M3 have comparable specific frequencies should not come as a surprise. It should be noted that an BSS specific frequency upturn similar to that observed in 47 Tuc and M3 has been detected also in M55 (Zaggia et al. 1997). This is of particular significance because M55 has a low central density---0.1 that of M3. Note that although the data set presented by Zaggia et al. for M55 are ground based observations and covered only a quadrant of the cluster, there is at least preliminary evidence that a upturn is present also in a low density cluster like M55. We note that the effect found by Zaggia et al in M55 could be even stronger since ground-based observations tend to hide BSS in the central region of the cluster. For example ground based observations of the center of M3 led Bolte et al. (1993) to claim depletion of BSS in the center whereas \\hst\\ observations eventually found a peak. M55 is surely less concentrated than M3, but still intermediate crowding conditions and the presence of highly saturated giants in the center can produce some level of incompleteness at the BSS level. The implications of the bimodality and relatively large populations of EBSS are not yet clear beyond the fact that EBSS may be fairly common and form in diverse clusters. If the BSS bimodality in M55 is confirmed, we will be faced with the result that clusters with central densities ranging over a factor of 400 can produce a bimodal BSS distributions. Any dynamical model like that of Sigurdsson et al. would seem more applicable to 47~Tuc than M3. It would be a real stretch to reach the dynamical conditions in M55. Given these factors it seems even more likely that primordial binaries play a fundamental role in the BSS formation mechanism. Better information on the abundance of primordial binaries in clusters and their survival rate at a given distance from the cluster center is essential. In 47 Tuc a recent extensive search for binary stars in the core (Albrow et al 2001) has led to an overall binary frequency of $14\\%\\pm 4\\%$. Their Fig.~21 shows that the binaries appear to be significantly more centrally concentrated with respect to the normal cluster stars but less concentrated than the BSS. Hence, if the cluster is in dynamical equilibrium, the CBSS are among the heaviest visible stellar components in the core of 47 Tuc. Albrow et al. also note that the W~UMa stars form a sequence in the CMD which like that which would be expected for equal mass binaries. The brightest lie in the BSS region of the CMD. Since W~UMa systems are contact binaries they may well merge to become BSS. Unfortunately, binary populations have been measured in only a few additional clusters (most notably NGC~6752, NGC~288), and they probably have diverse origins: {\\it (i)} in the core of NGC~6752 \\citep{ruba1} the binaries are probably not primordial, since NGC~6752 is such a dynamically evolved cluster \\citep[and references there-in]{ferraro6752bh}; {\\it (ii)} in the low density cluster NGC~288 \\citep{bellazzini288bin}, the binaries are probably primordial. Despite the much longer relaxation time of NGC~288, the binaries have all settled to within the inner half light radius; The technique used by Rubenstein \\& Bailyn (1997) and Bellazzini et al. (2002) requires only modest \\HST\\ observation time and should be applicable to any cluster. Even though such studies of the binary populations are painful, we need more. Clearly, it is desirable to have full BSS surveys for more clusters. It may well be a fluke that the first two clusters to be fully surveyed have bimodal distributions. We have pointed out \\citep{bss6} that all properly studied clusters have CBSS. The existence of EBSS in M3 \\citep{sandage53} marked the discovery of BSS. Similarly there is clear evidence for EBSS in ground based CMDs of 47~Tuc \\citep{sills00,grundahl47,kaluzny47}. Quite possibly all clusters with significant EBSS populations are bimodal. On the other hand, high quality CMDs of other clusters like M5 \\citep{sandquistm5} and M80 \\citep{brocatom80} show evidence for few if any EBSS. If more careful studies of such clusters do not reveal a yet undetected EBSS population, we will have to search for a mechanism which produces bimodal radial BSS distributions in some clusters but not in others. The population of BSS discovered here opens a new window for the investigation of the origin and the formation mechanism of BSS in GGCs. In fact, this discovery suggests that the {\\it peculiar} radial distribution firstly found in M3 is much more {\\it common} that was thought. Possibly in a few years we will refer to Figure 5 as the typical radial BSS distribution in stellar aggregates. In the meanwhile let us still refer to it as {\\it peculiar} until its generality can be finally addressed. We are now exploiting the wealth of information in the large dataset presented here by performing an extensive spectroscopic survey of the BSS population in 47 Tuc using the new multi-fiber Spectrograph FLAMES at the ESO-Very Large Telescope. These data should be highly informative. For instance, radial velocity determinations for a significant number of EBSS would help to clarify their origin: if the EBSS were ejected from the cluster center their velocity dispersion should be low because they are expected to be observed near the apocenter of an highly elliptical orbit; if EBSS arise from primordial binaries the velocity dispersion might be close to the Keplerian velocity (circular orbits which avoid the cluster center might allow the binaries to remain in the exterior) and show small variations due to low mass companions. The same spectra could lead to many mass estimates: so far the only BSS with a measured mass ($1.7 \\pm 0.4\\,\\msun$) is in 47~Tuc and was obtained with \\hst\\ \\citep{shara-bssmass}." }, "0311/astro-ph0311207_arXiv.txt": { "abstract": "The Major Atmospheric Gamma ray Imaging Cherenkov Telescope (MAGIC) is in commissioning phase and will start to become fully operative by the end of 2003. Located at {\\it El Roque de los Muchachos} in La Palma (Canary Islands, Spain), it has the largest reflector area (17 m diameter) of all the existing Cherenkov telescopes. New technologies have been used to reduce the energy threshold for gamma-ray detection to about 30 GeV. Due to its characteristics, the catalog of very high energy sources will considerably increase with the MAGIC observations, anticipating exciting results for the near future. An overview of the telescope, its current status and first results, together with a highlight of the scientific research is presented. ", "introduction": "Gamma ray astronomy has provided in the las few years spectacular results, led by the success of the CGRO ({\\it Compton Gamma Ray Observatory}) satellite mission which revealed that the high energy Universe was more exciting than expected. During its lifetime, the EGRET telescope onboard the CGRO provided an all-sky survey above 100 MeV consisting of 271 sources (Hartman et al. 1999). Apart of the remarkable detection of 66 AGN Blazars, the vast majority of these sources are still unidentified. Almost during the same period, the ground-based gamma ray astronomy has also been developed. The technique consists of the detection of the atmospheric Cherenkov light emitted by the particle showers initiated by gamma radiation on entering into the atmosphere. This Cherenkov flash lasts for a few {\\it ns}. The Whipple Cherenkov telescope opened this ground-based era by the detection of the Crab Nebulae in 1989. Then, extragalactic Blazar sources not seen by satellites were also discovered by using this technique. Other instruments based on the same principle have confirmed the Whipple results (HEGRA, CANGAROO, CAT) and nowadays about a dozen gamma ray sources have been detected to emit in the TeV energy range. The observations done by satellites measured well below around 10 GeV, while the existing Cherenkov telescopes have detected sources above 300 GeV. This energy gap, which is still virtually unexplored, is really important to understand which is the origin of the cut-off on the spectra which will explain the lack of sources measured by the Cherenkov telescopes in comparison to EGRET. The MAGIC (Major Atmospheric Gamma Imaging Cherenkov) telescope was designed in 1998 (Barrio et al. 1998) with the main goal of being the Imaging Atmospheric Cherenkov Telescope (IACT) with the lowest gamma energy threshold possible with the technological improvements affordable and based on the experience acquired with the first generation of Cherenkov telescopes. By using this detection technique, which provides much large effective areas (and much superior flux sensitivity) than satellite detectors, good angular resolution, acceptable energy resolution and a well tested capability to separate gammas from backgrounds, eventually a plethora of new sources will be discovered since for most of known sources the energy spectrum is of power-law nature and therefore they should exhibit a much higher flux in that energy region than at higher energies. \\begin{figure}[h] \\plotfiddle{DSCN0011.eps}{5,5cm}{0}{27}{27}{-115.}{0.} \\caption{View of the MAGIC Telescope in August 2003. Almost all the mirrors are already on place.} \\end{figure} The 17 m diameter f/D=1 MAGIC telescope is the largest of the new generation of IACTs. MAGIC is located in the Canarian island of La Palma (28.8 N, 17.9 W) at the {\\it Roque de los Muchachos} observatory (ORM), 2200 m above sea level. Its 241 m$^2$ parabolic dish is composed of 964 $49.5\\times 49.5$ cm$^2$ all-aluminum spherical mirror tiles mounted on a lightweight ($< 10$ ton) carbon fiber frame. The parabolic shape was chosen to minimize the time spread of the Cherenkov light flashes on the camera plane, which allows to reduce the rate of fake events induced by night-sky background light. Mirrors are grouped in panels of three or four, which can be oriented during the telescope operation through a novel active mirror control system to correct for the possible deformations of the telescope structure. The camera is made of 577 good quantum efficiency, fast photomultipliers with hemispherical photocatode that allows for light double-crossing. Each photomultiplier is coupled to a small light collecting cone to maximize the active surface of the camera. An special wavelength-shifting coating provides red extended sensitivity and allows for light-trapping, which increases the photomultiplier effective quantum efficiency. The total field of view of the camera is of about 4$\\deg$. The photomultiplier signals are transmitted to a distant Control House ($\\sim$150 m) by using analog optical fiber signals. Signals are processed by a multilevel trigger system and 300 MHz FADC are used for pulse digitalization. While all other new generation Cherenkov telescopes aim at the improvement of sensitivity and energy resolution in the 100 GeV regime by using stereoscopic systems of relatively small (10 m) telescopes, MAGIC, through the choice of a single, larger reflector, will achieve the lowest energy threshold among IACTs, of about 30 GeV. Its altazimuth mount can point to anywhere in the sky in about 20 seconds, a unique feature which is essential for the study of transient events like GRBs. ", "conclusions": "The MAGIC telescope is in its final commissioning phase and it is expected to start regular observations by the end of this year. If the telescope behaves as expected, it will soon be able to provide exciting results on a wide variety of astrophysical phenomena." }, "0311/astro-ph0311177_arXiv.txt": { "abstract": "Using {\\it Far Ultraviolet Spectroscopic Explorer} (FUSE) and {\\it Hubble Space Telescope} (HST) observations of the QSO PG\\,1259+593, we detect \\ion{D}{1} Lyman-series absorption in high velocity cloud Complex~C, a low-metallicity gas cloud falling onto the Milky Way. This is the first detection of atomic deuterium in the local universe in a location other than the nearby regions of the Galactic disk. We construct a velocity model for the sight line based on the numerous \\ion{O}{1} absorption lines detected in the ultraviolet spectra. We identify 8 absorption-line components, two of which are associated with the high velocity gas in Complex~C at $\\approx -128$ and $\\approx -112$ \\kms. A new Westerbork Synthesis Radio Telescope (WSRT) interferometer map of the \\ion{H}{1} 21\\,cm emission toward PG\\,1259+593 indicates that the sight line passes through a compact concentration of neutral gas in Complex~C. We use the WSRT data together with single-dish data from the Effelsberg 100-meter radio telescope to estimate the \\ion{H}{1} column density of the high velocity gas and to constrain the velocity extents of the \\ion{H}{1} Lyman-series absorption components observed by FUSE. We find N(\\ion{H}{1}) = $(9.0\\pm1.0)\\times10^{19}$ cm$^{-2}$, N(\\ion{D}{1}) = $(2.0\\pm0.6)\\times10^{15}$ cm$^{-2}$, and N(\\ion{O}{1}) = $(7.2\\pm2.1)\\times10^{15}$ cm$^{-2}$ for the Complex~C gas (68\\% confidence intervals). The corresponding light-element abundance ratios are D/H = $(2.2\\pm0.7)\\times10^{-5}$, O/H = $(8.0\\pm2.5)\\times10^{-5}$, and D/O = $0.28\\pm0.12$. The metallicity of Complex~C gas toward PG\\,1259+593 is approximately 1/6 solar, as inferred from the oxygen abundance [O/H] = $-0.79\\pm^{0.12}_{0.16}$. While we cannot rule out a value of D/H similar to that found for the local ISM (i.e., D/H $\\sim 1.5\\times10^{-5}$), we can confidently exclude values as low as those determined recently for extended sight lines in the Galactic disk (D/H~$< 1\\times10^{-5}$). Combined with the sub-solar metallicity estimate and the low nitrogen abundance, this conclusion lends support to the hypothesis that Complex~C is located outside the Milky Way, rather than inside in material recirculated between the Galactic disk and halo. The value of D/H for Complex~C is consistent with the primordial abundance of deuterium inferred from recent {\\it Wilkinson Microwave Anisotropy Probe} (WMAP) observations of the cosmic microwave background and simple chemical evolution models that predict the amount of deuterium astration as a function of metallicity. ", "introduction": "Observations of the abundance of deuterium relative to hydrogen (D/H) in different environments provide insight into the evolution of the light elements in the universe. With the excellent concordance in the estimates of the cosmic baryon density from measurements of D/H in low-metallicity quasar absorption-line systems and measurements of the cosmic microwave background, the cosmic baryon density and the primordial value of D/H are now tightly constrained (Burles, Nolett, \\& Turner 2001; O'Meara et al.\\ 2001; Spergel et al.\\ 2003). Thus, it should be possible to test chemical evolution models by examining the progression of D/H with time (see Lemoine et al.\\ 1999 and Olive, Steigman, \\& Walker 2000 for recent discussions). The abundance of deuterium is expected to decrease with time since there are no known sources of deuterium capable of increasing the cosmic abundance significantly (Epstein, Lattimer, \\& Schramm 1976; see also Prodanovi\\'c \\& Fields 2003). Many chemical evolution models predict moderate levels of deuterium destruction by stellar nucleosynthesis, typically less than a factor of 3--5 (Clayton 1985; Edmunds 1994; Steigman \\& Tosi 1995; Tosi et al.\\ 1998). More recent models suggest that slightly lower levels of astration are also possible (e.g., Chiappini, Renda, \\& Matteucci 2002). Unfortunately, measurements of D/H are particularly difficult, and there have been relatively few opportunities to directly measure the detailed changes in the abundance of deuterium as a function of metallicity or time. A key piece of information missing in discussions of the evolution of the light element abundances with time is the behavior of D/H in environments with metallicities between those of the high-redshift systems (typically $Z\\lesssim0.01Z_\\odot$) and those of gas in the local neighborhood of the Sun (typically $Z \\sim Z_\\odot$). There are several reasons why it it is important to determine the D/H ratio in a wide variety of galactic and extragalactic environments. First, there have been few high-precision estimates of D/H at moderate to high redshifts ($z \\sim 2-4$), where the amount of stellar processing of deuterium is presumably low, as evidenced by low metallicity (e.g., O'Meara et al.\\ 2001; Pettini \\& Bowen 2001 and references therein). The measurements that have been made appear to yield conflicting values for the primordial abundance of deuterium, with the observed values depending on the type of system observed (e.g., Lyman-limit or damped Ly$\\alpha$ systems -- see Pettini \\& Bowen 2001). Second, estimates of D/H in many locations yield a more global perspective of the chemical history of gas at different epochs than is possible from a few isolated measurements. Chemical evolution models seeking to describe the general evolution of the light-element abundances need a large sample of measurements to avoid systematic problems encountered by relying upon data for only a few types of environments. Third, although measurements of deuterium in nearby gas clouds imply a relatively constant value of D/H within the local interstellar medium (ISM; Moos et al.\\ 2002 and references therein), substantial variations may exist in D/H and D/O over distances of only a few hundred parsecs (Jenkins et al.\\ 1999; Sonneborn et al.\\ 2000; Hoopes et al.\\ 2003; H\\'ebrard \\& Moos 2003). If a sufficiently large number of high-precision D/H and D/O measurements can be made in a diverse set of nearby environments, it may be possible to understand the exact causes of this variability and the degree to which galactic chemical evolution and accretion of intragroup gas clouds influence the scatter in the observed ratios, both locally and at high redshift. This goal is a major science driver for the {\\it Far Ultraviolet Spectroscopic Explorer} (FUSE) mission (Moos et al.\\ 2000). To bridge the gap in D/H between low and high metallicity environments, we have obtained an extensive set of FUSE, {\\it Hubble Space Telescope} (HST), and interferometric \\ion{H}{1} 21\\,cm observations of the quasar PG\\,1259+593 behind high velocity cloud (HVC) Complex~C. The HVC is located at least 3.5 kpc from the Galactic plane (Wakker 2001), well beyond all Milky Way clouds with current D/H determinations. Unlike previous investigations of D/H and D/O in either high-redshift clouds or the local ISM, we know which gaseous system is responsible for the high-velocity \\ion{D}{1} Lyman-series absorption observed toward PG\\,1259+593. A global description of the gas in Complex~C is available from both emission and absorption-line measurements. The neutral gas in Complex~C has been mapped extensively in \\ion{H}{1} 21\\,cm emission (see Wakker 2001 and references therein) and low-ionization absorption (e.g., Wakker et al.\\ 1999; Richter et al.\\ 2001b; Collins, Shull, \\& Giroux 2003). The ionized gas in Complex~C has been investigated in absorption by Sembach et al.\\ (2003) and Fox et al.\\ (2003), and in emission by Tufte, Reynolds, \\& Haffner (1998) and Wakker et al.\\ (1999). Complex~C is an excellent site to determine D/H for comparisons with the high-redshift values because it is chemically young (Richter et al.\\ 2001b; Collins et al. 2003; Tripp et al.\\ 2003) and has a metallicity (10--25\\% solar) lower than that of the general ISM of the Milky Way and higher than that of intergalactic clouds at high redshifts. In this paper we describe these new measurements and the resulting D/H and D/O ratios in Complex~C. In \\S2 we describe the FUSE and HST Space Telescope Imaging Spectrograph (HST/STIS) absorption-line observations and the \\ion{H}{1} 21\\,cm interferometer observations. Section~3 contains a short summary of the properties of the PG\\,1259+593 sight line. In \\S4 we outline the methods and general assumptions used to determine the column densities of \\ion{H}{1}, \\ion{D}{1}, and \\ion{O}{1} in Complex~C. We determine the \\ion{H}{1} column density from interferometric \\ion{H}{1} 21\\,cm emission data and use the FUSE and HST/STIS ultraviolet absorption-line data to determine the \\ion{O}{1} and \\ion{D}{1} column densities. Sections 5, 6, and 7 contain descriptions of these determinations for oxygen, hydrogen, and deuterium, respectively, and provide estimates of the various errors. The column densities and error ranges are summarized and discussed in \\S8, and comments on future progress appear in \\S9. We summarize the results of the study in \\S10. ", "conclusions": "We summarize the Complex~C column densities of \\ion{H}{1}, \\ion{D}{1}, and \\ion{O}{1} in Table~9, together with the $1\\sigma$ (68\\% confidence interval) and $2\\sigma$ (95\\% confidence interval) error estimates based on the descriptions of errors given above. Systematic errors dominate the uncertainties for all three species. We also list the D/H, D/O, and O/H column density ratios, along with the propagated errors from the column density determinations. A summary of the light element abundance ratios for Complex~C and other environments is given in Table~10, along with references for the ratios. The environments considered in this table include the local ISM, the ISM of the Galactic disk at distances of 200--1000 pc from the Sun, and high-redshift absorption line systems (both Lyman-limit systems and damped Ly$\\alpha$ systems). We list both D/H and D/O when estimates of [O/H] are available. These comparisons are not meant to be exhaustive summaries of all the available data in the literature; rather they are synopses of several studies given to provide insight into how the Complex~C values compare to typical values found elsewhere. The data in Table~10 are shown graphically in Figure~17, where we plot D/H as a function of metallicity determined through measurements of [O/H], or [Si/H] if [O/H] is not available. For the Lyman-limit systems where [Si/H] is used, the metallicity may be somewhat uncertain because large ionization corrections are necessary to convert estimates of N(\\ion{Si}{2})/N(\\ion{H}{1}) into estimates of N(Si)/N(H). Various other metallicity indicators are available, but [O/H] tends to be the best for the reasons described in \\S5 (see also the detailed discussion presented by Timmes et al.\\ 1997). Our estimate of D/H in Complex~C rules out values of D/H greater than $3.3\\times10^{-5}$ or less than $1.1\\times10^{-5}$ with a reasonable degree of confidence ($\\sim95$\\%). Higher values over-produce the amount of \\ion{D}{1} absorption in the higher order Lyman series lines, and lower values under-produce the amount of absorption expected in the Ly$\\delta$ line. Both limits have interesting implications. Most models of Galactic chemical evolution predict that astration of deuterium should result in a factor of $\\sim1.5-3$ decline in the cosmic abundance of deuterium from the Big Bang to the present time (Edmunds 1994; Tosi et al.\\ 1998; Chiappini et al. 2002), though higher levels of astration may be possible if special conditions, such as a Galactic wind, are invoked to counter over-production of elements heavier than helium (see, e.g., Scully et al.\\ 1997). For a simple closed-box model with the assumption of instantaneous recycling, a deuterium astration efficiency of 60\\%, and standard oxygen yields, Pagel (1997) finds that the there should be only a few percent reduction of deuterium relative to its primordial abundance if [O/H] $\\lesssim -1$, as found for Complex~C. In Figure 17, we show the results for a simple chemical evolution model from Fields et al.\\ (2001) having (D/H)$_p$ and little destruction of deuterium at low metallicities from Population~III stars. More complicated chemical evolution models requiring bimodal episodes of star formation to explain the scatter in D/H values measured at high redshift can be found in that work. Our $2\\sigma$ upper limit of D/H $< 3.3\\times10^{-5}$ suggests that the value of D/H produced by Big Bang nucleosynthesis cannot be too much larger than this upper limit unless the assumptions about the destruction of deuterium and production of oxygen are invalid. Similar arguments have been made to infer a primordial value of (D/H)$_p \\lesssim 3\\times10^{-5}$ from determinations of D/H in high-redshift, low-metallicity systems (Tytler et al.\\ 2000; O'Meara et al.\\ 2001; Kirkman et al.\\ 2003). Values of (D/H)$_p \\lesssim 3\\times10^{-5}$ imply values of the baryon density, $\\Omega_bh^2$, and the baryon-to-photon ratio, $\\eta$, in concordance with estimates derived from cosmic microwave background (CMB) measurements. Using the standard Big Bang nucleosynthesis predictions for the abundance of deuterium (Burles et al. 2001), a value of (D/H)$_p \\lesssim 3\\times10^{-5}$ implies $\\Omega_bh^2 \\gtrsim 0.02$ and $\\eta \\gtrsim 5.6\\times10^{-10}$. The CMB estimates cluster near values of $\\Omega_bh^2 \\approx 0.022-0.024$ (Netterfield et al.\\ 2002; Pryke et al.\\ 2002; Spergel et al.\\ 2003), and imply (D/H)$_p = (2.62\\pm^{0.19}_{0.20})\\times10^{-5}$ (Spergel et al. 2003; see also Steigman 2003). The region between the dashed horizontal lines in Figure~17 is the $\\pm2\\sigma$ range of (D/H)$_p$ allowed by the recent {\\it Wilkinson Microwave Anisotropy Probe} (WMAP) observations. With the PG\\,1259+593 data, we can confidently rule out very high D/H values for Complex~C, like those claimed for some intergalactic systems [e.g., D/H $= (2.0\\pm0.5)\\times10^{-4}$ at $z\\approx0.7$ toward PG\\,1718+4807; Webb et al.\\ 1997]. This is important because the Complex~C measurement is the only other measurement of D/H outside the Milky Way at $z < 1.0$. Such high values are also inconsistent with the CMB measurements, and recently the value for the $z=0.7$ absorber toward PG\\,1718+4807 has been challenged on other grounds (e.g., Kirkman et al.\\ 2001 -- but see Crighton et al.\\ 2003). On the low end of the D/H range, we cannot exclude the possibility that D/H in Complex~C is similar to D/H in the local ISM (D/H $\\sim1.5\\times10^{-5}$, Moos et al. 2002; Linsky 2003). However, we can rule out D/H values as low as those found for some Galactic disk sight lines extending beyond the local ISM (i.e., at $d \\gtrsim 100$ pc). Values of D/H less than $1.0\\times10^{-5}$ have been determined for two extended sight lines in the Galactic disk using FUSE data. Hoopes et al.\\ (2003) find D/H = $(0.85\\pm^{0.34}_{0.34})\\times10^{-5}$ ($2\\sigma$) toward HD\\,195965 ($d \\sim 800$ pc), and D/H = $(0.78\\pm^{0.52}_{0.25})\\times10^{-5}$ ($2\\sigma$) toward HD\\,191877 ($d \\sim 2200$ pc). Using IMAPS data, Jenkins et al.\\ (1999) find D/H = $(0.74\\pm^{0.19}_{0.13})\\times10^{-5}$ (90\\% confidence) toward $\\delta$\\,Ori\\,A ($d \\sim 500$~pc); a similar result was found by Laurent, Vidal-Madjar, \\& York (1979) using Copernicus data. The low values found for these sight lines indicate that more astration of deuterium may have occurred in these regions than in the local ISM or in Complex~C. The local star formation histories of the regions explored toward $\\delta$\\,Ori\\,A and HD\\,195965 are roughly consistent with this idea (see Hoopes et al.\\ 2003 and Jenkins et al.\\ 1999), but the number of sight lines is still too small to draw general conclusions. Eventually, with a large enough number of sight lines, it may be possible to determine whether refinements to chemical evolution models are needed. The value of D/O = $0.28\\pm0.12$ in Complex~C is very different from the D/O ratio in the local ISM or the disk of the Milky Way. H\\'ebrard \\& Moos (2003) find D/O = $0.038\\pm0.002$ for white dwarf and subdwarf sight lines within $\\sim150$ pc of the Sun. Most of this difference can be accounted for by the different metallicities of the local ISM and Complex~C. Note that even without the \\ion{H}{1} measurement, the D/O ratio strongly suggests that the metallicity in Complex~C is lower than solar. The Complex~C result is more similar to the value of D/O = $0.37\\pm0.03$ found for the $z=3.025$ absorber toward Q\\,0347-3819 (Levshakov et al. 2002) than it is to the value for the Galactic disk. However, values of D/O for two other high-redshift systems are an order of magnitude or more larger than the Complex~C value. We note that determinations of N(\\ion{O}{1}) in the high-redshift systems are challenging since the weaker \\ion{O}{1} lines occur within the \\ion{H}{1} Ly$\\alpha$ forest. As estimates of D/H in the high-redshift systems have fallen in recent years, a few local measurements with high values of D/H have received more attention. For example, the value of D/H = $(2.18\\pm^{0.36}_{0.31}) \\times10^{-5}$ for the $\\gamma^2$ Vel sight line (90\\% confidence; Sonneborn et al.\\ 2000) is indistinguishable from the values for the high-redshift ($z>2$) systems toward HS\\,0105+1619 and Q\\,1243+3407, and perhaps higher than the value for Q\\,2206--199 (see Table~10 and references therein). Preliminary FUSE results for the IX~Vel sight line, which lies $\\sim1\\degr$ from $\\gamma^2$ Vel, indicate that the value of D/H for the first $\\sim100$ pc of the $\\gamma^2$ Vel sight line is similar to that in the local ISM (Blair et al.\\ 2003). If this result is confirmed by upcoming FUSE observations, then the D/H ratio in the Vela region must be even higher than the sight line average. Note, however, that the $\\zeta$~Pup sight line, which is in the same region of the sky and extends further than the $\\gamma^2$~Vel sight line has a D/H ratio similar to that of the local ISM (see Table~17). One possible explanation for a high value in the ISM toward $\\gamma^2$ Vel is infall of ``primordial'' (metal-poor, D-rich) gas. Clouds as large as Complex~C could contribute significant amounts of deuterium to regions of the Galactic disk; a mass of M(\\ion{H}{1}) $\\gtrsim(1.2-3.0)\\times10^6$ M$_\\odot$ and D/H = $2.2\\times10^{-5}$ imply that Complex~C contains $\\gtrsim(26-66)$ M$_\\odot$ of deuterium. The effects of infalling low-metallicity gas have been considered by various authors in the context of deuterium astration and potential solutions to the well-known ``G-dwarf'' problem (Tosi 1988; Matteucci \\& Francois 1989; Tosi et al.\\ 1998; Wakker et al.\\ 1999). However, a serious problem with this interpretation for the $\\gamma^2$~Vel region is that the gas along the sight line does not show a corresponding reduction in the ratios of metals to hydrogen as would be expected if infall were the explanation for the high D/H ratio. Furthermore, the constancy of the oxygen and krypton abundances within several hundred parsecs of the Sun (Meyer, Jura, \\& Cardelli 1998; Cartledge et al.\\ 2001) supports the idea that the nearby gas is reasonably well mixed and has similar heavy element abundances. Any explanation for the high D/H ratio in the Vela region that incorporates infall must also include processing of the gas to explain the relatively ``normal'' metal abundances in the region. Other possibilities also exist for explaining the variations of D/H in the Milky Way. For example, D could be depleted onto dust grains (Jura 1982). From simple thermodynamic considerations, Draine (2003) has shown that differences in the binding energies of D and H on the peripheries of polycyclic aromatic hydrocarbons (PAHs) could produce extreme enrichment of deuterium in bound form, similar to what has been found for some simple interstellar molecules. He argues that it is plausible that roughly 20\\% of the H atoms bound to PAHs are replaced by D. If this is correct, he estimates that there are sufficient PAH binding sites in a volume of normal interstellar material to account for a decrease of free D atoms by one part in $10^5$ of H (i.e., roughly the size of the reduction observed along some lines of sight). In some places, the atomic medium can revert to the intrinsic D/H when the PAHs are destroyed by the recent passage of a shock, which could explain why D/H seems to vary. Reinforcing the concept that variable D abundances might be explained by Draine's proposal is the recent detection of the CD bond stretch features at 4.4 and 4.67 $\\mu$m toward the Orion Bar reported by Peeters (2002). She found ${\\rm D/H} = 0.17\\pm 0.07$ in bound form. This detection, although not very far above the noise, indicates that Draine's estimate of a 20\\% replacement of H by D atoms may be approximately correct. For any gas with an overall metallicity as low as that of Complex C, the concentration of PAHs would be so low that there would be no measurable reduction in the concentration of free deuterium atoms, which is consistent with our finding for the sight line toward PG\\,1259+593. Finally, Mullan \\& Linsky (1998) have proposed that deuterium production in stellar flares may cause some variations in D/H locally, but Prodanovi\\'c \\& Fields (2003) have recently reconsidered this issue and find that flare production of deuterium is not a significant source of \\ion{D}{1} on galactic scales. They concur with Mullan \\& Linsky that local variations by stellar flare production cannot be ruled out. Dust depletion and stellar flare production of deuterium are unlikely to alter the present value of D/H in Complex~C substantially. Complex~C contains no known stars, and analyses of its metal content indicate that it contains little dust. The metallicity of $Z/Z_\\odot\\approx0.17$ we derive for Complex~C through the ratio of \\ion{O}{1} and \\ion{H}{1} is very similar to the recent determination of the metallicity in Complex~C toward 3C\\,351 by Tripp et al.\\ (2003). They find [O/H] = $-0.76\\pm^{0.23}_{0.21}$. Both metallicity estimates are slightly higher than the value of $Z/Z_\\odot\\sim0.1$ inferred for the Mrk~290 line of sight through Complex~C based on measurements of \\ion{S}{2} and \\ion{H}{1} (Wakker et al.\\ 1999). The metallicity estimates may depend on environment and may change slightly as a function of position within Complex~C; Collins et al.\\ 2003 find a range from 0.1 to 0.25 solar, but some of these estimates are complicated by uncertain ionization corrections for \\ion{S}{2}. One possible explanation for the slightly varying abundances in Complex~C might be that the gas is interacting with the Galactic corona or high Galactic halo, as indicated by measurements of highly ionized gas associated with the complex (Sembach et al.\\ 2003; Fox et al.\\ 2003). Mixing of the diffuse regions of Complex~C with the more metal abundant gas of the Galactic thick disk and halo would tend to increase the observed abundances in these regions, while the abundances in denser clumps of gas, as seen toward PG\\,1259+593, should reflect more closely the original (pre-interaction) abundances in the cloud. It is interesting that our best-fit value of D/H found for Complex~C is similar to that found for high-redshift damped Ly$\\alpha$ systems (see Pettini \\& Bowen 2001 and Kirkman et al.\\ 2003). The \\ion{H}{1} column density of N(\\ion{H}{1}) $=9.0\\times10^{19}$ cm$^{-2}$ is sufficient to produce damping wings on the \\ion{H}{1} Ly$\\alpha$ and Ly$\\beta$ lines. If observed in isolation from the Milky Way by an outside observer, Complex~C would produce \\ion{H}{1} and \\ion{D}{1} absorption similar to that seen in these higher redshift systems. Complex~C also has a low nitrogen abundance inferred from the ratio of \\ion{N}{1} to \\ion{H}{1}, [N/H] $\\approx -1.9$ toward PG\\,1259+593 (Richter et al.\\ 2001b; Collins et al. 2003), about a factor of 10 lower than [O/H]. (Initial work on the amount of \\ion{N}{2} present indicates that photoionization of \\ion{N}{1} in the \\ion{H}{1} regions is not able to account for the deficit in the derived value of [N/H].) Commensurate values of [N/H] have been found for other Complex~C sight lines (Collins et al. 2003; Tripp et al. 2003). Some damped Ly$\\alpha$ systems show similar deficits of nitrogen relative to oxygen, which several authors have explained as a nucleosynthetic effect resulting from delayed release of nitrogen between massive and intermediate mass stars or a top-heavy or truncated initial mass function (Prochaska et al.\\ 2002; Molaro 2003). At the metallicity of Complex~C, the low N/O ratio observed is probably more dependent upon the secondary nature of nitrogen production than these other processes (see Figure~6 in Pettini et al. 2002). The metallicity, abundance pattern, D/H ratio, and lower distance limit all indicate that Complex~C is an external system (or the remains of an external system) falling into with the Milky Way rather than gas ejected from the Galactic disk. We report the detection of \\ion{D}{1} Lyman-series absorption in high velocity cloud Complex~C, a low-metallicity gas cloud falling onto the Milky Way. This is the first detection of atomic deuterium in the local universe in a location other than the nearby regions of the Galactic disk. The primary results of our study are as follows: \\begin{enumerate} \\item{We combine data from the {\\it Far Ultraviolet Spectroscopic Explorer}, {\\it Hubble Space Telescope}, and Westerbork Synthesis Radio Telescope to derive the column densities of \\ion{H}{1}, \\ion{D}{1}, and \\ion{O}{1} in the direction of PG\\,1259+593 ($l=120\\fdg56, b = 58\\fdg05$).} \\item{We use the FUSE and HST/STIS absorption-line data to construct a velocity model for the sight line based on the numerous \\ion{O}{1} absorption-lines detected. We identify 8 absorption-line components common to all three species (\\ion{H}{1}, \\ion{D}{1}, and \\ion{O}{1}) along the sight line. Two of these components, at $v_{\\rm LSR}\\approx -128$ and --112 \\kms, are identified with Complex~C. The remaining lower-velocity components are identified with the Intermediate Velocity Arch, the Milky Way ISM, and a positive intermediate-velocity cloud in this direction.} \\item{Our WSRT interferometer map of the \\ion{H}{1} 21\\,cm emission toward PG\\,1259+593 indicates that the sight line passes through a compact concentration of neutral gas in Complex~C. We use the WSRT data having a resolution of $\\sim1\\arcmin$ together with single-dish data from the Effelsberg 100-meter radio telescope to estimate the \\ion{H}{1} column density in Complex~C and to constrain the velocity extents of the Complex~C \\ion{H}{1} Lyman-series absorption components observed by FUSE.} \\item{We use a combination of analysis techniques to estimate the column densities of \\ion{H}{1}, \\ion{D}{1}, and \\ion{O}{1} in the absorbing regions along the sight line. We combine line-profile fitting analyses and curve-of-growth techniques with $\\chi^2$ minimization codes to estimate the most likely value and uncertainties on the column density of each species. The use of multiple analysis methods provides important consistency checks on the results of the individual methods. The good agreement found from independent analyses by members of our team confirm that the formal errors are reliable representations of the uncertainties involved in the analyses.} \\item{For each species, we describe the uncertainties involved in the estimation of the column density. Systematic errors dominate the uncertainties, especially for \\ion{D}{1}. In estimating the uncertainties on D/H, we have included the possibility that some of the absorption in the vicinity of the \\ion{D}{1} lines might be caused by a weak, high-velocity \\ion{H}{1} feature near --190 \\kms.} \\item{For the Complex~C absorption, we find N(\\ion{H}{1}) = $(9.0\\pm1.0)\\times10^{19}$ cm$^{-2}$, N(\\ion{D}{1}) = $(2.0\\pm0.6)\\times10^{15}$ cm$^{-2}$, and N(\\ion{O}{1}) = $(7.2\\pm2.1)\\times10^{15}$ cm$^{-2}$. The uncertainties quoted are $1\\sigma$ errors (68\\% confidence intervals); two sigma error estimates can be found in Table~9. The corresponding light element abundance ratios are D/H = $(2.2\\pm0.7)\\times10^{-5}$, O/H = $(8.0\\pm2.5)\\times10^{-5}$, and D/O = $0.28\\pm0.12$.} \\item{ The metallicity of the Complex~C gas in this direction is approximately 1/6 solar, as inferred from the oxygen abundance of [O/H] = $-0.79\\pm^{0.12}_{0.16}$~($1\\sigma$). This metallicity is in excellent agreement with a recent high-quality determination of [O/H] in Complex~C toward 3C\\,351, which is located $\\sim30\\degr$ from PG\\,1259+593. Two other Complex~C sight lines analyzed by Collins et al. (2003) yield similar metallicities (Mrk~279: [O/H] = $-0.71\\pm^{0.36}_{0.25}$; Mrk~817: [O/H = $-0.59\\pm^{0.25}_{0.17}$). } \\item{While we cannot rule out a value of D/H for Complex~C similar to that found for the local ISM (i.e., D/H~$\\sim1.5\\times10^{-5}$), we can reasonably exclude values as low as those determined recently for extended sight lines in the Galactic disk (D/H~$< 1\\times10^{-5}$), provided that the absorption observed is really \\ion{D}{1} and not very high velocity \\ion{H}{1}. This conclusion lends support to the hypothesis that Complex~C is a manifestation of gas originating outside the Milky Way, rather than material recirculated between the Galactic disk and halo.} \\item{A value of D/H $\\approx 2.2\\times10^{-5}$ for Complex~C is consistent with the primordial abundance of deuterium inferred from recent WMAP observations of the cosmic microwave background, combined with simple chemical evolution models that predict the amount of deuterium astration with time. Reducing the errors on the D/H estimate in Complex~C toward PG\\,1259+593 will be difficult, but may eventually be possible with a large space telescope equipped with a high-resolution ultraviolet spectrograph.} \\item{The number of sight lines for which \\ion{D}{1} can be observed in Lyman-series absorption beyond a few hundred parsecs of the Sun appears to be small because of \\ion{H}{1} velocity crowding. We considered the data for several other sight lines through Complex~C and found that they are unsuitable for a reliable \\ion{D}{1} column density determination. We continue to examine FUSE data for \\ion{D}{1} absorption in other HVCs.} \\end{enumerate} \\noindent {\\bf Acknowledgments.} We thank Van Dixon for answering numerous questions about the {\\tt CALFUSE} pipeline software, and we thank Max Pettini and Jeff Linsky for useful comments on the manuscript. This work is based on data obtained for the Guaranteed Time Team by the NASA-CNES-CSA FUSE mission operated by the Johns Hopkins University. We acknowledge the many contributions of the members of the FUSE D/H working group and the tireless efforts of the FUSE Operations Team. This work is also based in part upon observations made with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS5-26555. The Westerbork Synthesis Radio Telescope is operated by the ASTRON (Netherlands Foundation for Research in Astronomy) with support from the Netherlands Foundation for Scientific Research (NWO). The Effelsberg Telescope belongs to the Max Planck Institute for Radio Astronomy in Bonn, Germany. We are indebted to Robert Braun for assistance with the Westerbork observations. Partial financial support has been provided by NASA contract NAS5-32985 and Long Term Space Astrophysics grants NAG5-3485 (K.R.S.) and NAG5-7262 (J.M.S.). B.P.W.\\ was supported by NASA grants NAG5-9179, NAG5-9024, and NAG5-8967. T.M.T.\\ appreciates support for this work from NASA Long Term Space Astrophysics grant NAG5-11136 as well as NASA grant GO-08695.01-A from the Space Telescope Science Institute. P.R.\\ was supported by the Deutsche Forschungsgemeinschaft. French co-authors were supported by CNES. G.H.\\ and R.F.\\ acknowledge use of the {\\tt Owens.f} software developed by Martin Lemoine and the French FUSE Team. \\clearpage \\newpage" }, "0311/astro-ph0311388_arXiv.txt": { "abstract": "{We present a model of a pulsar wind nebula evolving inside its associated supernova remnant. The model uses a hydrodynamics code to simulate the evolution of this system when the pulsar has a high velocity. The simulation distinguishes four different stages of pulsar wind nebula evolution: the supersonic expansion stage, the reverse shock interaction stage, the subsonic expansion stage and ultimately the bow shock stage. The simulation bears out that, due to the high velocity of the pulsar, the position of the pulsar is off-centered with respect to its pulsar wind nebula, after the passage of the reverse shock. Subsequently the pulsar wind nebula expands subsonically untill the event of the bow shock formation, when the motion of the pulsar becomes supersonic. The bow shock formation event occurs at roughly half the crossing time, when the pulsar is positioned at 0.677 times the radius of the supernova remnant blastwave, in complete agreement with analytical predictions. The crossing time is defined by the age of the supernova remnant, when the pulsar overtakes the blastwave bounding the supernova remnant. The results of the model are applied to three supernova remnants: N157B, G327.1-1.1 and W44. We argue that the head of the pulsar wind nebula, containing the active pulsar, inside the first two systems are not bounded by a bow shock. However, in the case of W44 we argue for a scenario in which the pulsar wind nebula is bounded by a bow shock, due to the supersonic motion of the pulsar. ", "introduction": "A supernova remnant (SNR) is the relic of a supernova explosion, which injects an energy of $\\sim 10^{51}$ erg into the surrounding medium. The dynamics of a young SNR expanding into the interstellar medium (ISM) is determined by two shocks: a forward shock which propagates into the ISM and is being decelerated by sweeping up material from the ISM, and a reverse shock which results from the high pressure behind the forward shock and propagates back into the freely expanding ejecta of the SNR (McKee 1974; McKee \\& Truelove 1995). A young SNR becomes dynamically more interesting in those cases where the collapse of the progenitor star, preceding the supernova explosion, yields a pulsar: a rapidly rotating neutron star. In those cases the dynamics of the central region of the SNR is dominated by the continuous injection of energetic particles by a relativistic pulsar wind, driven by the spin-down energy of the pulsar. The pulsar wind is terminated by a strong MHD shock (Rees \\& Gunn 1974), and drives a pulsar wind nebula (PWN) in the interior of the young SNR. The dynamics of the PWN is coupled to the evolution of the SNR, because the total energy release over the pulsar's lifetime is small ($\\sim 10^{49}-10^{50}$ erg) compared with the total mechanical energy of the SNR ($\\sim 10^{51}$ erg). Several authors (Reynolds \\& Chevalier 1984; van der Swaluw et al. 2001; Blondin et al. 2001; Bucciantini et al. 2003) have considered the evolution of a centered PWN inside an evolving SNR. In these systems, the initial stage of the PWN is the supersonic expansion stage: the pulsar wind bubble is bounded by a strong PWN shock propagating through the freely expanding ejecta of the SNR. A transition to the reverse shock interaction stage takes place when the reverse shock collides with the PWN shock, and subsequently crushes the pulsar wind bubble (van der Swaluw et al. 2001, Blondin et al. 2001). This reverse shock interaction stage is characterised by an unsteady expansion of the pulsar wind bubble (van der Swaluw et al. 2001), due to the reverberations from the violent collision between the reverse shock and the PWN shock. The expansion of the PWN proceeds subsonically when these reverberations have vanished. The expansion is subsonic because the surroundings of the PWN have been reheated by the passage of the reverse shock: the PWN shock, bounding the swept-up shocked ejecta around the hot pulsar wind bubble, has disappeared. In this paper we discuss the evolution of PWNe inside SNRs, for pulsars with a constant kick velocity: initially the PWN starts its expansion at the center of the SNR, however the pulsar motion will move the pulsar wind cavity along as it moves through the SNR interior. We present results from hydrodynamical simulations for such a system, which distinguishes all three evolutionary stages mentioned above, i.e. the supersonic expansion stage, the reverse shock interaction stage and the subsonic expansion stage. However at the end of the simulation an additional stage can be distinguished, when the head of the PWN, containing the active pulsar, deforms into a bow shock, due to the supersonic motion of the pulsar. The supersonic expansion stage in the simulation shows a PWN which is off-centered with respect to the twofold shock structure of the SNR, due to the kick velocity of the pulsar. Therefore the timescale on which the reverse shock collides with the complete shock surface, bounding the PWN, can be a significant fraction of the total lifetime of the PWN when the reverse shock interaction stage starts. We use a semi-analytical approach to estimate the timescale of the collision process, which is shown to scale roughy with the pulsar velocity. The results from the simulation are in almost complete agreement with these semi-analytical calculations. Due to the high velocity of the pulsar, its position inside the PWN is strongly off-centered after the passage of the reverse shock. Ultimately, at the end of the simulation, when the pulsar is approaching the shell of its SNR, the head of the PWN, containing the active pulsar, is deformed into a bow shock at a time and position, in complete agreement with analytical predictions made by van der Swaluw et al. (1998). We determine the evolutionary stage of the PWNe inside the SNRs N157B, G327.1-1.1 and W44, using our model. For the first two systems, we argue that the position of the pulsar at the head of its PWN, is a result of the passage of the reverse shock and the high velocity of the pulsar: the head of these PWNe are not bounded by a bow shock. Therefore these PWNe are either in the {\\it reverse shock interaction stage} or the {\\it subsonic expansion stage}. The PWN inside the SNR W44 however, is shown to be a good candidate for having a bow shock nebula around its pulsar. ", "conclusions": "We have considered the case of a PWN interacting with a SNR, for which the associated pulsar is moving at a high velocity through the interior of its SNR. The model we discussed made use of a hydrodynamics code. One could distinguish four different stages in the simulation: the supersonic expansion stage, the reverse shock interaction stage, the subsonic expansion stage and the bow shock stage. Below we summarise the most important results from our model: \\begin{itemize} \\item The reverse shock interaction stage starts when the reverse shock collides with the front of the PWN shock. The timescale on which the reverse shock collides with the {\\it complete} PWN shock, bounding the pulsar wind bubble, scales with the pulsar velocity. This timescale is a significant fraction of the lifetime of the PWN, when the reverse shock interaction stage starts, for PWNe containing a high velocity pulsar. \\item The high velocity of the pulsar results in an off-centered position of the pulsar with respect to its pulsar wind bubble, after the passage of the reverse shock. \\item The morphology of a PWN, after the passage of the reverse shock, consists of a twofold structure which ultimately expands subsonically inside the relaxed Sedov-Taylor SNR: \\par 1) a roughly spherically symmetric relic PWN \\par 2) a head, containing the pulsar, directed towards the SNR shell \\item The formation of a bow shock around the head of the PWN occurs at half the crossing time, when the position of the pulsar $R_{\\rm psr}$ with respect to the SNR shock $R_{\\rm snr}$ equals $R_{\\rm psr}\\simeq 0.677 R_{\\rm snr}$. We derived a lower limit for the pulsar velocity for the bow shock formation to occur while the SNR is in its Sedov-taylor stage. \\end{itemize} From our model it follows that a SNR containing a pulsar at the head of its PWN does not imply the presence of a bow shock, bounding the head feature of the PWN. We discussed three SNRs and interpreted their morphology, using the results from our model. For the SNRs N157B and G327.1-1.1 we argued that they contain PWNe which do not have a bow shock. The PWN inside the SNR W44 seems to be a good candidate for having a bow shock around the head of its PWN." }, "0311/astro-ph0311561_arXiv.txt": { "abstract": "The outflow in SDSS J0300+0048 has the highest column density yet reported for a broad absorption line quasar. The absorption from different ions is also segregated as a function of velocity in a way that can only be explained by a disk wind outflow. Furthermore, the presence of the such large column densities of gas at the high observed outflow velocities may be incompatible with purely radiative acceleration. MHD contributions to the acceleration should be considered seriously. ", "introduction": " ", "conclusions": "" }, "0311/astro-ph0311082_arXiv.txt": { "abstract": "{ In the framework of the ESO Large Programme ``First Stars'', very high-quality spectra of some 70 very metal-poor dwarfs and giants were obtained with the ESO VLT and UVES spectrograph. These stars are likely to have descended from the first generation(s) of stars formed after the Big Bang, and their detailed composition provides constraints on issues such as the nature of the first supernovae, the efficiency of mixing processes in the early Galaxy, the formation and evolution of the halo of the Galaxy, and the possible sources of reionization of the Universe. This paper presents the abundance analysis of an homogeneous sample of 35 giants selected from the HK survey of Beers et al. (1992; 1999), emphasizing stars of extremely low metallicity: 30 of our 35 stars are in the range $-2.7 <{\\rm [Fe/H]}< -4.1$, and 22 stars have ${\\rm [Fe/H]} < -3.0$. Our new VLT/UVES spectra, at a resolving power of $R\\sim45,000$ and with signal-to-noise ratios of 100-200 per pixel over the wavelength range 330 -- 1000 nm, are greatly superior to those of the classic studies of McWilliam et al. (1995) and Ryan, Norris, \\& Beers (1996). The immediate objective of the work is to determine precise, comprehensive, and homogeneous element abundances for this large sample of the most metal-poor giants presently known. In the analysis we combine the spectral line modeling code ``Turbospectrum'' with OSMARCS model atmospheres, which treat continuum scattering correctly and thus allow proper interpretation of the blue regions of the spectra, where scattering becomes important relative to continuous absorption ($\\lambda < 400$~nm). We obtain detailed information on the trends of elemental abundance ratios and the star-to-star scatter around those trends, enabling us to separate the relative contributions of cosmic scatter and observational/analysis errors. Abundances of 17 elements from C to Zn have been measured in all stars, including K and Zn, which have not previously been detected in stars with [Fe/H] $< -$3.0. Among the key results, we discuss the oxygen abundance (from the forbidden [OI] line), the different and sometimes complex trends of the abundance ratios with metallicity, the very tight relationship between the abundances of certain elements (e.g., Fe and Cr), and the high [Zn/Fe] ratio in the most metal-poor stars. Within the error bars, the trends of the abundance ratios with metallicity are consistent with those found in earlier literature, but in many cases the scatter around the average trends is {\\it much} smaller than found in earlier studies, which were limited to lower-quality spectra. We find that the cosmic scatter in several element ratios may be as low as 0.05~dex. The evolution of the abundance trends and scatter with declining metallicity provides strong constraints on the yields of the first supernovae and their mixing into the early ISM. The abundance ratios found in our sample {\\it do not} match the predicted yields from pair-instability hypernovae, but are consistent with element production by supernovae with progenitor masses up to 100 M$_{\\sun}$. Moreover, the composition of the ejecta that have enriched the matter now contained in our very metal-poor stars appears surprisingly uniform over the range $-4.0 \\le {\\rm [Fe/H]} < -3.0$. This would indicate either that we are observing the products of very similar primordial bursts of high-mass stars, or that the mixing of matter from different bursts of early star formation was extremely rapid. In any case, it is unlikely that we observed the ejecta from individual (single) supernovae (as has often been concluded in previous work), as we do not see scatter due to different progenitor masses. The abundance ratios at the lowest metallicities ($-4.0 \\le {\\rm [Fe/H]} \\le -3.0$) are compatible with those found by McWilliam et al. (1995) and later studies. However, when elemental ratios are plotted with respect to Mg, we find no clear slopes below [Mg/H] = --3, but rather, a plateau-like behaviour defining a set of initial yields. \\keywords {Galaxy: abundances -- Galaxy: halo -- Galaxy: evolution -- Stars: abundances -- Stars: Population II -- Stars: Supernovae -- reionization} } ", "introduction": " ", "conclusions": "\\begin {figure*} \\begin {center} \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-c.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-o.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-na.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-al.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-si.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-ca.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-ti.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-sc.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-k.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-cr.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-mn.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-fe.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-co.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-ni.ps} } \\resizebox {5.8cm}{2.90cm} {\\includegraphics {abgmg-zn.ps} } \\caption {Abundance ratios [X/Mg] plotted vs. [Mg/H]. The scale is the same for all the plots: the amplitude in the coordinate [X/Mg] is 2.2 dex. For Mn the symbols are the same as in Fig. \\ref{abCr-Mn}} \\label {abMg-SiZn} \\end {center} \\end {figure*} \\subsection{Comparison with previous studies of very metal-poor stars} The trends of the relations [X/Fe] vs. [Fe/H] reported in the present study are generally in agreement with previous results in the literature, such as those of McWilliam et al. (\\cite{MPS95}), Ryan et al. (\\cite{RNB96}), and Norris, Ryan, \\& Beers (\\cite{NRB01}). However, the much smaller scatter of the ratios [X/Fe] is a notable result of the greatly improved spectra obtained for this study. Carretta et al. (\\cite{CGC02}) observed three very metal-poor MS or TO stars and two extremely metal-poor giants ([Fe/H] $< -3.0$) with Keck spectra having a figure of merit F (Norris et al. \\cite{NRB01}) larger than 600. The elements from Mg to Fe were analyzed. Within a (rather large) scatter the two analyses are compatible, except for a few elements, in particular Cr (their slope of [Cr/Fe] vs. [Fe/H] seems to be steeper than ours). \\subsection{Mg as an alternative ``reference element''}\\label{sec-refelem} Iron is a convenient ``reference element'' in high-resolution spectral analyses, because it has by far the largest number of usable lines and is represented by two ionization states. Iron may not be the best choice as a tracer of Galactic chemical evolution, however, since its nucleosynthesis channels are not very well understood, and are not necessarily even unique (e.g., explosive nucleosynthesis, Si burning in massive SNe II, or SNe Ia). From the point of view of the chemical evolution of the Galaxy oxygen would be a better choice (see Wheeler, Sneden \\& Truran \\cite{WST89}), as oxygen is the most abundant element after H and He, it comes from a single source, and its abundance should not be significantly affected in the explosive phase. However, its well-known observational difficulties would considerably degrade the accuracy of the derived trends: Oxygen could be measured in only 21 of our programme stars and the uncertainties on its abundance are large (see in particular Sect. \\ref{sec-errors} and the error bars in Fig. \\ref{ab-o}). Mg or Ca might be good alternatives. The abundances of these elements are accurately determined, and they are also formed mainly in massive SNe. We choose Mg rather than Ca here (Fig. \\ref{abMg-SiZn}), because Mg is more robust in the sense that its production is dominated by hydrostatic burning processes and it is also less affected by explosive burning and by ``fallback'' (Woosley \\& Weaver \\cite{WW95}). Note that Shigeyama \\& Tsujimoto (\\cite{ST98}) also recommended Mg rather than Fe as a useful reference element, following much the same logic. We note, however, that the ``plateau''-like behavior of [Mg/Fe] with increasing iron abundance in the range $-$4 to $-$2 implies that Mg and Fe have parallel early nucleosynthesis histories. Therefore, the trends of elemental ratios with [Fe/H] found in section \\ref{CZn-Fe} should survive if [Fe/H] is replaced by [Mg/H] as a metallicity indicator. Yet, the new diagrams (see Fig. \\ref{abMg-SiZn}) present a few notable differences. One is expected -- the scatter is never as low as in some of the earlier diagrams because [Mg/H] is determined less accurately than [Fe/H], being based on 8 lines instead of 150. The other -- new -- result is very interesting: rather than a roughly linear variation over the full range $-4.0<$ [Mg/H] $<-2.5$, there is a hint that all abundance ratios are flat in the interval $-4.0<$ [Mg/H] $<-3.0$, with something qualitatively different occurring at higher metallicity. This pattern is in fact what is expected if the first SNe are primordial and have specific yields: the plateau between $-4.0$ and $-3.0$ would then reflect a pure zero-metallicity type of SNe, whereas at higher metallicity we may observe a mix of primordial and non-primordial SNe. At still higher metallicities ($\\ge$ --2.0), the scatter is expected to decrease again because so many SN precursors are involved that any differences average out. We now consider a few elements of particular interest. $\\bullet$ Carbon\\\\ Karlsson \\& Gustafsson (\\cite{KG01}) have statistically simulated the chemical enrichment of a metal-poor system, assuming that the stellar yields are one-dimensional functions of the progenitor mass of the supernovae and the masses of the supernovae are distributed according to a Salpeter IMF. In particular, they computed the distribution of the abundance ratio [C/Mg] vs. [Mg/H] in a hypothesized sample of 500 XMP stars (their Fig. 3), adopting the yields of either Woosley \\& Weaver (\\cite{WW95}) or Nomoto et al. (\\cite{NHT97}). Our Fig. \\ref{abMg-SiZn} for C is compatible with their Fig. 3a (yields of Woosley \\& Weaver), but not with their Fig. 3b (yields of Nomoto et al.). On the other hand, both in their simulation and as observed in our present sample, the [C/Mg] ratio seems to decrease with increasing [Mg/H]. Following Karlsson and Gustafsson (\\cite{KG01}), this effect could be the result of different supernova masses operating at different metallicities. SNe producing a high [C/Mg] ratio produce only small amounts of Mg; on the contrary, SNe producing a low [C/Mg] ratio also produce substantial Mg. Karlsson \\& Gustafsson show that the patterns they predict become barely visible (or invisible) in an observational sample smaller than N $\\approx 500$ and affected by uncertainties of the order of 0.1 dex. We have examined how their [C/Mg] vs. [Mg/Fe] diagram would appear for our sample (Fig. \\ref{abmg-C-Mg}). Not only is no fine structure visible, but our diagram is considerably more extended in the vertical direction, strongly suggesting that the scatter in [C/Mg] is not explained by the theoretical yields. \\begin {figure} \\begin {center} \\resizebox {6.cm}{6.cm} {\\includegraphics {abmg-C-Mg.ps}} \\caption {[C/Mg] vs. [Mg/Fe] for all stars of our sample with temperatures higher than 4800 K (and thus presumably not mixed). Filled circles represent the most metal-poor stars ([Fe/H] $\\leq -3.0$), open circles the stars with [Fe/H]$> -3.0$. The variation of [C/Mg] is larger than expected from the computations of Karlsson \\& Gustafsson (\\cite{KG01}), suggesting that the scatter in [C/Mg] is not explained by the theoretical yields. } \\label {abmg-C-Mg} \\end {center} \\end {figure} $\\bullet$ Sodium and Aluminium\\\\ [0cm] [Na/Fe] and [Al/Fe] exhibit very different behaviours as functions of [Fe/H] (Fig. \\ref{abFe-NaAl}). However, when Mg is used as a reference element (Fig. \\ref{abMg-SiZn}), the behavior of these elements appears rather similar, and a plateau at the lowest metallicities appears for Na as well as for Al: below $\\mathrm{[Mg/H] = -3.0}$, [Na/Mg] remains constant at about [Na/Mg] = 0.9 (in fact this plateau appears also in Fig. \\ref{abFe-NaAl}: [Na/Fe] $\\approx 0.7$ for [Fe/H]$<-3.4$). However, the rise in [Na/Mg] at higher metallicity is not seen for Al. The discrepant position of CS~22952--015 in the diagrams of [Na/Mg] and [Al/Mg] vs. [Mg/H] will be discussed in Sect. \\ref{pec-stars}. $\\bullet$ Iron peak elements\\\\ There is no clear slope of [Cr/Mg] vs. [Mg/H], as was seen for [Cr/Fe] vs. [Fe/H]. This raises the suspicion that the slope in the latter diagram may be an artifact due to different NLTE corrections for the two elements as a function of metallicity (Th\\'evenin \\& Idiart \\cite{TI99}). These corrections have not been applied, as they are not known for Cr and have not been published line by line for Fe. We return to this question below. No significant slope is found for [Mn/Mg], (in agreement with absence of slope for [Mn/Fe]). Similarly, for [Mg/H]$< -3.0$ no clearly significant slope is found for the other elements (Fe, Co, Ni, Zn) . \\subsection{Cosmic vs. observational scatter in the abundance ratios} \\label{sec-scatter} A key motivation for the present programme was to explore to what extent the scatter in the observed abundance ratios is due to observational error, and to what extent it reflects physical conditions in the early Galaxy when these stars were formed. McWilliam et al. (\\cite{MPS95}) and McWilliam \\& Searle (\\cite{MS99}) already noted that the scatter in some of their diagrams of [X/Fe] vs. [Fe/H] could be entirely accounted for by observational errors. The issue was summarized by Ryan, Norris, \\& Beers (\\cite{RNB96}) as follows. ``The abundance patterns, especially those of Cr, Mn, or Co, raise the following question: why should all halo supernova ejecta around this epoch that possess a particular [Cr/Fe] ratio (or [Mn/Fe] ratio or [Co/Fe] ratio) {\\it subsequently} form into stars of the same [Fe/H]? Put differently, how do the ejecta know how much interstellar hydrogen to combine with?''. Our observations were designed to achieve twice the spectral resolution and 3--4 times the S/N ratio of the earlier data in order to test this very point. With our much lower observational errors, we can conclude that while the scatter for C, Na, Mg, Al, and Si is probably real, the very small scatter of Ca, Cr, and Ni {\\it still} do not leave room for the existence of an intrinsic scatter! Consider the case of chromium, which has the lowest observed scatter (r.m.s. 0.05 dex). The problem mentioned earlier is very acute and derives from the {\\it simultaneous} absence of scatter and presence of of a slope of [Cr/Fe] versus metallicity. Although one can argue that the amount of hydrogen swept up by the ejecta is mainly determined by the energy of the explosion of the SN (Cioffi et al. \\cite{CMB88}), thus relating the abundance ratios produced by the SN to the final [Fe/H] of the enriched gas, it is still difficult to believe that there is so little room for noise in the mixing process. However, if the slope is in fact spurious (e.g., due to neglected differential NLTE corrections between Cr and Fe), as suggested by the diagram of [Cr/Mg] vs. [Mg/H] (Fig. \\ref{abMg-SiZn}), the problem vanishes. One would simply conclude that Cr and Fe are produced together, independent of the metallicity of the SN progenitor, and the amount of mixing cannot be localized anymore along the metallicity axis. Until detailed NLTE computations for Cr and Fe become available we cannot decide if this interpretation is correct. For all the elements discussed here, our results show that the scatter of their production ratios is very small, far below the values derived earlier. This implies that we are observing either the ejecta of fairly large bursts of massive stars, so the sampling of the IMF is reasonably good, or the result of several events promptly mixed by strong turbulence. \\subsection{The nature of the first supernovae} Theoretical work (Bromm et al. \\cite{BFC01} and references therein) predict that the first stellar generation is made of very massive stars, with masses above 100 M$_{\\odot}$, because zero-metal matter lacks adequate cooling mechanisms for fragmenting down to classical supernova-progenitor masses. Results of WMAP (Kogut et al. \\cite{KSB03}) on an early reionization of the Universe have triggered further claims (Cen \\cite{Cen03}) of a very massive stellar generation.This has very important implications for early 'stellar' nucleosynthesis. According to current models, such stars either end up as pair-instability supernovae, or as collapsed black holes, in the latter case with no contribution to the metal enrichment of the ISM (Heger \\& Woosley \\cite{HW02}; Umeda \\& Nomoto \\cite{UN02}). A comparison of the yields of Heger \\& Woosley with our results show a clear disagreement, in particular the predicted strong odd-even effect, not seen in our observations, and a strong decline of Zn with metallicity also not observed. Nakamura et al. (\\cite{NUI01}) and Nomoto et al. (\\cite{NHT97}) have computed yields of SNe with a progenitor mass of 25 M$_{\\odot}$ of very high energy (also called hypernovae). Their predicted yields have some positive features at very low metallicity, such as the high [Zn/Fe] and [Co/Fe], and a low [Mn/Fe], as observed. However they also predict a lack of [O/Fe] enhancement, in clear disagreement with our observations. Our conclusion is that classical SNe are still the best candidates to explain our observational results. \\subsection{Peculiar objects} \\label{pec-stars} \\subsubsection{CS~22949--037} The highly peculiar abundances of CS~22949--037 were studied in detail by Depagne et al. (\\cite{DHS02}). They may be explained by a single progenitor or by an enrichment event dominated by massive SNe II with substantial fallback; this applies also to the similar star CS~29498--043 analysed by Aoki et al. (\\cite{ANR02a}, \\cite{ANR02b}). Tsujimoto \\& Shigeyama (\\cite{TS03}) propose another interpretation of CS~22949--037 and CS~29498--043. The high [Mg/Fe] ratio could be due to a low-energy explosion, strong enough to eject the layers containing the light elements, but ejecting little iron and other iron-peak elements. In this interpretation, the lack of Fe relative to the lighter elements O and Mg is also associated with a large fallback on the remnant, but due to a low explosion energy rather than a large mass of the collapsed core as in the model adopted by Depagne et al. (\\cite{DHS02}). However, the normal Cr/Mn/Fe/C/Ni ratios observed in this star would rather suggest a normal explosion energy, the larger fallback being due to a larger mass of the single or multiple progenitors. Further exploration of the precise abundance patterns of these two putative low-energy supernova descendents should be quite interesting. This interpretation would link the exceptional cases of the most extreme metal-poor stars with rather low-mass progenitors. It would, however, conflict with the usual interpretation which associate the more massive progenitors with both the earliest explosions and the largest volume of hydrogen swept up, resulting in low metallicity in the ISM. The problem clearly requires further investigation. \\subsubsection{CS~22952--015} CS~22952--015 is known to be Mg deficient (McWilliam et al., \\cite{MPS95}). In the abundance ratio plots vs. [Mg/H] (Fig. \\ref{abMg-SiZn}), it does indeed appear Fe-rich, but its most notable characteristic is the large values of [Na/Mg] and [Al/Mg]. An interesting point is that this effect is not seen as clearly in the diagrams of [Na/Fe] and [Al/Fe] vs. [Fe/H] (Fig. \\ref{abFe-NaAl}), because all three elements Na, Al, and Fe are enhanced relative to Mg in CS~22952--015. \\subsubsection{CS~22169--035} CS~22169--035 appears to be particularly deficient in Ti (Fig. \\ref{ab-mgti}), but in fact all the ratios [Mg/Fe], [Si/Fe], [Ca/Fe], [Co/Fe], [Ni/Fe], and [Zn/Fe] are low as well. When Mg is used as the reference element, this star has a normal position in the diagrams, and the abundance anomalies are most simply characterised as a deficiency of Fe. \\subsection{Yields of the first supernovae} With Mg chosen as the reference element (Fig. \\ref{abMg-SiZn}), the most metal-poor stars in our sample ([X/H]$<-2.9$) define a plateau at abundance ratios [X/Mg] corresponding to the yields of the first supernovae, thus providing constraints on these yields. The mean value of [X/Mg] of each plateau is given in Table \\ref{yieldsuper} and represent our best estimate of the yields from the first metal producers in the Galaxy. \\begin {table}[t] \\caption {Mean [X/Mg] ratios for stars with [Mg/H]$<-2.9$, corresponding to the yields of the first supernovae. The r.m.s. is the scatter around the mean.} \\label{yieldsuper} \\begin {center} \\begin {tabular}{lrrrr} \\hline & &$r.m.s.$& n \\\\ $[$Na/Mg] & $-$0.84 & 0.22 & 11 \\\\ $[$Al/Mg] & $-$0.33 & 0.15 & 13 \\\\ $[$Si/Mg] & $+$0.21 & 0.14 & 14 \\\\ $[$K/Mg] & $-$0.14 & 0.14 & 13 \\\\ $[$Ca/Mg] & $+$0.06 & 0.09 & 14 \\\\ $[$Sc/Mg] & $-$0.17 & 0.14 & 14 \\\\ $[$Ti/Mg] & $-$0.01 & 0.09 & 14 \\\\ $[$Cr/Mg] & $-$0.63 & 0.09 & 14 \\\\ $[$Mn/Mg] & $-$0.65 & 0.20 & 14 \\\\ $[$Fe/Mg] & $-$0.21 & 0.10 & 14 \\\\ $[$Co/Mg] & $+$0.13 & 0.17 & 14 \\\\ $[$Ni/Mg] & $-$0.23 & 0.13 & 14 \\\\ $[$Zn/Mg] & $+$0.15 & 0.19 & 14 \\\\ \\hline \\end {tabular} \\end {center} \\end {table} In Fig. \\ref{compayields}a, these mean values are compared to the values predicted by Woosley \\& Weaver (\\cite{WW95}) for their zero-metal supernova models 15A, 25B and 35C (progenitor masses 15, 25 and 35 $M_{\\odot}$). The Woosley \\& Weaver values of [X/Mg] for the elements from C to Ca have been enhanced in order to bring their predicted value of [Fe/Mg] into agreement with our observations. In Fig. \\ref{compayields}b we compare with the zero-metal models of Chieffi \\& Limongi (\\cite{CL03}) for 15, 20, 35, 50 $M_{\\odot}$ stars. The predicted [X/Mg] values were already adjusted by Chieffi \\& Limongi to obtain [Mg/Fe]$\\approx +0.45$). These comparisons suggest that some adjustment of the models is indeed required. The connection between the abundances observed in these very old stars and those observed in intergalactic clouds -- in particular the damped Ly$_\\alpha$ systems (DLA) -- will be further discussed in subsequent papers. \\begin {figure} \\begin {center} \\resizebox {7.cm}{3.5cm} {\\includegraphics {woosley-nous.ps} } \\resizebox {7.cm}{3.5cm} {\\includegraphics {limongi-nous.ps} } \\caption {Mean [X/Mg] values for [Mg/H]$<-3$ (Fig. \\ref{ab-mgti} and Table \\ref{yieldsuper}) compared to the model yields by: {\\bf a)} Woosley and Weaver for 15 (dotted line), 25 (dashed), and 35 $M_{\\odot}$ SNe (full), and {\\bf b)} Limongi and Chieffi for 15 (dotted line), 20 (dotted-dashed), 35 (dashed), and 50 $M_{\\odot}$ (full). } \\label{compayields} \\end {center} \\end {figure}" }, "0311/nucl-th0311055_arXiv.txt": { "abstract": "An effective field theory technique that combines the standard nuclear physics approach and chiral perturbation theory is applied to the $hen$ process, ${}^{3}\\mbox{He}+n\\rightarrow {}^{4}\\mbox{He}+ \\gamma$. For the initial and final nuclear states, high-precision wave functions are generated via the variational Monte Carlo method using the Argonne $v_{14}$ potential and Urbana VIII trinucleon interactions, while the relevant transition operators are calculated up to ${\\cal O}(Q^4)$ in HB$\\chi$PT. The imposition of the renormalization condition that the magnetic moments of ${}^{3}\\mbox{He}$ and ${}^{3}\\mbox{H}$ be reproduced allows us to carry out a parameter-free calculation of the $hen$ cross section. The result, $\\sigma=(60\\pm 3\\pm 1)~\\mu b$, is in reasonable agreement with the experimental values, $(54\\pm 6)~\\mu b$ and $(55\\pm 3)~\\mu b$. This agreement demonstrates the validity of the calculational method previously used for estimating the reaction rate of the solar $hep$ process. ", "introduction": " ", "conclusions": "" }, "0311/astro-ph0311269_arXiv.txt": { "abstract": "We examine the ability of the Sunyaev-Zel'dovich effect (SZE) statistics proposed by Lee (2002) using numerical simulations. The statistics describe the distribution of the peak heights in the noise-contaminated SZE sky maps, and provide an analysis technique to sort out the noise-contamination effect and estimate the number density of real clusters. The method is devised to be suitable for the interferometric SZE observations in drift-scanning mode like AMiBA experiment. We apply the proposed method to a set of realistic SZE sky maps constructed from large-scale cosmological simulations, and show that the method indeed allows us to estimate the number density of clusters efficiently. The efficiency of the method is demonstrated in two aspects: (i) it can reconstruct the number density of clusters even though the clusters contribute only a small fraction ($\\sim 10 \\%$) of the total number of peaks in each SZE map; (ii) it can count even those clusters whose amplitudes are low enough to be compatible with the noise level while conventional method of cluster identification using the threshold cutoff cannot count them properly within the same observation time. Thus, the proposed method may be useful for the study of cluster abundance exploiting future interferometric SZE surveys such as AMiBA. ", "introduction": "The thermal Sunyaev-Zel'dovich effect (SZE) by galaxy clusters represents the spectral distortion of the cosmic microwave background radiation (CMBR) through the inverse Compton scattering by the hot intra-cluster electrons \\citep{sun-zel72}. Since the SZE is independent of the redshift, it provides a unique tool to detect high-redshift clusters. Thus, the blank-field SZE survey can measure the formation epoch and the abundance evolution of galaxy clusters, which can be used as a probe to put constraints on the cosmological parameters \\citep{bar-etal96,hen97,bah-fan98,via-lid99,bor-etal99,fan-chi01, gre-etal01,mol-etal02}. The SZE flux diminution is directly proportional to the cluster Comptonization parameter $y$, defined as \\beq y(\\bth) = \\frac{\\sigma_{T}k_{B}}{m_{e}c^{2}} \\int n_{e}(l\\hbth)T_{e}(l\\hbth)dl, \\label{eqn:compton} \\eeq where $\\bth$ represents the location vector along the line-of-sight ($\\hbth \\equiv \\bth/\\vert \\bth\\vert$) and the integral is over the CMBR path in the line-of-sight direction. Hence the SZE survey at a fixed frequency basically provides a map of the $y$-parameter. The number density of galaxy clusters, however, cannot be directly measured from the observed SZE maps because of the presence of noise. Conventionally, signals from the noise-contaminated maps are identified as the local peaks whose flux amplitudes (for the SZE observation, proportional to the peak height of the Comptonization parameter, $y$) significantly exceed the noise level. Therefore, to detect lower-amplitude signals using the conventional approach, one has to decrease the noise level by increasing the observational time. Unfortunately, the necessary observational time increases inversely proportional to the squares of the imposed noise level. It is clearly inefficient to use the conventional signal-selection method for the estimate of the cluster number counts, because of the high costs of the interferometric observations. Recently \\citet{lee02} proposed a statistical strategy to estimate efficiently the number density of galaxy clusters from the noise-contaminated SZE maps constructed from the interferometric observation in drift-scanning mode, with the AMiBA experiment as a target model. By {\\it the efficient estimate}, we mean that the method is capable of estimating the number density of clusters which contribute only a small fraction of the total number density of peaks in the SZE maps, counting properly even those clusters whose peak heights are low enough to be compatible with the noise level without increasing the observational time. The key feature of the method is to exploit the property that the noise-contamination is described by a Gaussian statistic. Then the cluster number density can be estimated by simply deconvolving a one-dimensional Gaussian function and the zero-th order approximation to the cluster number density. Although the method has been shown to render reasonable results for idealized Monte-Carlo simulations \\citep{lee02}, it is necessary to test its ability by applying it to more realistic simulations before using it in real SZE surveys. In this {\\it Letter}, we test the ability of the SZE statistical strategy using the outputs of a large cosmological simulation. In $\\S 2$, we provide a concise overview of the model experiment AMiBA and the SZE statistics. In $\\S 3$, we describe the numerical simulation, explain how to apply the statistical strategy to the simulated SZE maps, and compare the reconstructed cluster number densities with the original ones. In $\\S 4$, we summarize the results, and give concluding remarks. ", "conclusions": "We have tested the validity of the SZE statistics proposed by \\citet{lee02} using numerical simulations. The statistical method is devised for the interferometric SZE observations in drift scanning mode, targeting at AMiBA as a model experiment, and applied to estimate the number density of cluster signals that occupy only a small fraction ($\\sim 10 \\%$) of the noise-contaminated SZE maps more efficiently than the conventional method based on the threshold cut-off. The method is developed to model the dominant noise contamination effect on the SZE peak distribution by a Gaussian scatter and to deconvolve the noise scatter and the SZE peak distribution. The stability of the deconvolution process has been achieved by convolution with a Wiener optimal filter estimated from the measurable power spectrum of the SZE peaks. We have applied the method to a large number of realistic SZE sky maps, and found that it indeed works for all realizations, giving a fairly good estimate of the cluster number density, including even those clusters whose peak heights are compatible with noise-levels. In the conventional approach, those low-amplitude peaks are usually all disregarded as noise, so that a large fraction of low-amplitude cluster signals are missing. Our method allows us to count those missing low-amplitude clusters as quickly and accurately as possible. Hence, we conclude that it will provide an efficient method to determine the number density of galaxy clusters in the future interferometric SZE surveys." }, "0311/astro-ph0311096_arXiv.txt": { "abstract": "The COBE-DMR 4-year maps displayed a strong non-Gaussian signal in the ``inter-scale'' components of the bispectrum: their observed values did not display the scatter expected from Gaussian maps. We re-examine this and other suggested non-Gaussian features in the light of WMAP. We find that they all disappear. Given that it was proved that COBE-DMR high noise levels and documented systematics could at most {\\it dilute} the observed non-Gaussian features, we conclude that this dataset must have contained non-negligible undocumented systematic errors. It turns out that the culprit is a combination of QuadCube pixelization and data collected during the ``eclipse season''. ", "introduction": "The possibility of non-Gaussianity in the COBE-DMR 4 year maps led to a rather protracted story. Using ``single-$\\ell$'' bispectrum analysis, \\citet{fmg} found strong evidence for non-Gaussianity in the anisotropy of the Cosmic Microwave Background (CMB) temperature. This detection was followed by similar claims by \\citet{nfs98} and \\citet{pvl98}, and caused considerable consternation among theorists (see \\cite{nvs} for a discussion). Further work, however, showed that claims other than those based on the bispectrum could not be reproduced \\citep{muk}. The bispectrum claims were confirmed by~\\cite{bt99}. Later~\\cite{banday} cast serious doubts upon the cosmological origin of the observed signal. A systematic was identified which removed the observed ``single-$\\ell$'' bispectrum signal -- the so-called eclipse season data, which should never have been used. However, when an extension to ``inter-$\\ell$'' bispectrum components was sought, a new non-Gaussian signal was found for inter-$\\ell$ separations of $\\Delta l=1$~\\citep{joao}. Specifically, their observed values were found to concentrate uncannily close to zero instead of displaying the scatter expected from Gaussian maps. This signal could not be blamed on any systematic effects studied by~\\cite{banday} or otherwise \\citep{joao}. It was also proved that the instrument high noise levels could at most dilute the observed signal, in spite of the noise correlations and anisotropy \\citep{joao}. It was also found that the observed signal did not extend to higher inter-$\\ell$ separations $\\Delta l>1$ \\citep{haav}. What shall we make of these claims in the light of the recent observations \\citep{wmap} by the Wilkinson Microwave Anisotropy Probe (WMAP)? In this paper we show that they don't survive the new data, which displays consistency with Gaussianity on the angular scales probed by COBE. This might seem at first suprising, given that the WMAP data is considerably less noisy than the COBE-DMR dataset, and that noise can at most hide a non-Gaussian signal. However one should never forget the issue of systematics. Even though documented COBE-DMR systematics were shown not to correlate with the observed non-Gaussian signal, the new data allows us to cast a new look at the problem. We find that a highly non-subtle combination of QuadCube pixelization systematics and the ``eclipse data'' are to be blamed for the observed effect. This, we believe, closes the story. The moral is clear: care must be exercised regarding similar claims currently being made with WMAP. ", "conclusions": "The conclusion is that a combination of pixelization effects and the ``eclipse'' systematic is behind the inter-$\\ell$ non-Gaussian effect previously reported for COBE maps. Curiously if we take a difference map (COBE minus WMAP) the anomalous alignment of the $J^{3}_{l}$ is not present (see top panel of Fig.~\\ref{fig3}). On the scales we are considering such maps picture COBE noise, plus COBE systematics, minus WMAP systematics (assumed to be small). Thus one would expect the COBE $J^{3}_{l}$ signal to be enhanced in the difference map. The fact that it is not results from the non-linearity of the statistic being used, plus the subtle interplay of signal and systematic via the QuadCube pixelization scheme. In this paper we have concentrated on $J^3_\\ell$, because the single scale bispectrum anomalies have been explained long ago. However we have checked that no new anomalies on large angular scales emerge in the WMAP data. We reserve to a future publication a complete study of the bispectrum of WMAP of smaller angular scales. To our mind the work of \\cite{ngwmap} is just the beginning." }, "0311/astro-ph0311575_arXiv.txt": { "abstract": "Simulations of large-scale structure in the universe have played a vital role in observational cosmology since 1980's in particular. Their important role will definitely continue to be true in the 21st century. Rather the requirements for simulations in the precision cosmology era will become more progressively demanding; they are supposed to fill the missing link in an accurate and reliable manner between the ``initial'' condition at z=1000 revealed by WMAP and the galaxy/quasar distribution at z=0 - 6 surveyed by 2dF and SDSS. In this review, I will summarize what we have learned so far from the previous cosmological simulations, and discuss several remaining problems for the new millennium. ", "introduction": "Cosmological $N$-body simulations started in late 1970's, and since then have played an important part in describing and understanding the nonlinear gravitational clustering in the universe. As far as I know, the cosmological N-body simulation in a comoving periodic cube, which is quite conventional now, was performed for the first time by Miyoshi \\& Kihara (1975) using $N=400$ particles. Figure 1 plots the evolution of the number of particles employed in cosmological N-body simulations. Here I consider only the ``high-resolution'' simulations including Particle-Particle, Particle-Particle--Particle-Mesh, and tree algorithms which are published in refereed journals (excluding, e.g., conference proceedings). I found that the evolution is well fitted by \\begin{eqnarray} \\label{eq:Nyear} N= 400 \\times 10^{0.215({\\rm Year}-1975)} , \\end{eqnarray} where the amplitude is normalized to the work of Miyoshi \\& Kihara (1975). Just for comparison, the total number of CDM particles of mass $m_{\\rm CDM}$ in a box of the universe of one side $L$ is \\begin{eqnarray} N= \\frac{\\Omega_\\CDM \\rho_{\\rm cr} L^3}{m_\\CDM} \\approx 10^{83} \\left(\\frac{\\Omega_\\CDM}{0.23} \\right) \\left(\\frac{L}{1 h^{-1}{\\rm Gpc}}\\right)^3 \\left(\\frac{1 {\\rm keV}}{m_\\CDM}\\right) \\left(\\frac{0.71}{h}\\right) . \\end{eqnarray} If I simply extrapolate equation (1) and adopt the WMAP parameters (Spergel et al. 2003), then the number of particles that one can simulate in a $(1 h^{-1}{\\rm Gpc})^3$ box will reach the real number of CDM particles in December 2348 and February 2386 for $m_\\CDM= 1{\\rm keV}$ and $10^{-5}{\\rm eV}$, respectively. I have not yet checked the above arithmetic, but the exact number should not change the basic conclusion; simulations in the new millennium will be {\\it unbelievably} realistic. \\begin{figure}[tbh] \\begin{center} \\leavevmode\\epsfxsize=7.50cm \\epsfbox{Nyear.eps} \\caption{Evolution of the number of particles in ``high-resolution'' cosmological N-body simulations. \\label{fig:Nyear}} \\end{center} \\end{figure} ", "conclusions": "It turned out that I was able to review the cosmological simulations only for a time-scale of decades. Still the material may be heavily biased, which I have to apologize for the organizers and possible readers of these proceedings. A millennium is definitely too long for any scientist to make any reliable prediction for its eventual outcome. Thus the number of simulation particles that I predicted in Introduction might sound ridiculous, but in reality the progress in the new millennium may be even more drastic that whatever one can imagine. For instance, it is unlikely that one still continues to use currently popular particle- or mesh-based simulation techniques over next hundreds of years. In that case, the number of particles may turn out to be a totally {\\it useless} measure of the progress or reliability of simulations. Nevertheless I believe that a historical lesson that I learned in preparing this talk will be still true even at the end of this millennium; good science favors the prepared mind, not the largest simulation at the time. \\bigskip" }, "0311/astro-ph0311605_arXiv.txt": { "abstract": "We explore possible correlations between light profile shapes, as parameterized by the S\\'ersic index $n$ or the concentration index $C_{re}(1/3)$, and relevant stellar population parameters in early-type galaxies. Mean luminosity weighted ages, metallicities and abundance ratios were obtained from spectra of very high signal--to--noise and stellar population models that synthesize galaxy spectra at the resolution given by their velocity dispersions $\\sigma$, in combination with an age indicator (H$\\gamma_{\\sigma}$) that is virtually free of the effects of metallicity. We do not find any significant correlation between $n$ (or $C_{re}(1/3)$) and mean age or metallicity, but a strong positive correlation of the shape parameters with Mg/Fe abundance ratio. This dependence is as strong as the Mg/Fe -- $\\sigma$ and $C_{re}(1/3)$ -- $\\sigma$ relations. We speculate that early-type galaxies settle up their structure on time-scales in agreement with those imposed by their Mg/Fe ratios. This suggest that the global structure of larger galaxies, with larger Mg/Fe ratios and shorter time-scales, was already at place at high $z$, without experiencing a significant time evolution. ", "introduction": "\\label{Sec:intro} A successful galaxy formation model should be able to explain the main scaling relations followed by early-type (E/SO) galaxies. These galaxies show up a tight relation between the size (by means of the effective radius $r_e$), surface brightness, and central velocity dispersion, $\\sigma$, which is known as the Fundamental Plane (Djorgovski \\& Davis 1987; Dressler et al. 1987). Early-type galaxies also follow the Color-Magnitude Relation, CMR, i.e. brighter galaxies look redder (Visvanathan \\& Sandage 1977; Bower et al. 1992). The CMR and the Mg$_2$--$\\sigma$ relation (e.g. Bender, Burstein \\& Faber 1992; Colless et al. 1999) links the mass of a galaxy, through its luminosity, to its constituent stellar populations. However, little is known about the relation between the morphological shape and the stellar populations of early-type galaxies (Conselice 2003), most likely because these two fields rely on methods that have been traditionally disconnected. From a stellar population point of view, measured colors and absorption line-strengths are translated to mean ages and metallicities via stellar population synthesis models. Using these models the CMR is often interpreted as a metallicity sequence (e.g., Arimoto \\& Yoshii 1987), i.e. elliptical galaxies are old and passively evolving systems (e.g., Ellis et al. 1997; Stanford, Eisenhardt \\& Dickinson 1998). However, this view is questioned by some studies, which show evidence for a significant intermediate age population in some ellipticals (e.g., Gonz\\'alez 1993; J\\o rgensen 1999). Element ratios provide further constraints on the star formation histories (e.g. Worthey 1998). These studies have shown compelling evidence for a systematic departure from solar element ratios of some elements as a function of $\\sigma$ (Trager et al. 2000). Recent modeling efforts have predicted spectra of stellar populations at moderately high resolution (Vazdekis 1999; Schiavon et al. 2002; Bruzual \\& Charlot 2003), leading to more accurate stellar population parameter estimates (Vazdekis \\& Arimoto 1999; Vazdekis et al. 2001a; Kuntschner et al. 2002). From the point of view of galaxy structure, the shape of the overall surface brightness as parametrized by S\\'ersic (1968) index, $n$, correlates with the global properties of early-type galaxies: total luminosity, central surface brightness and effective radius, $r_e$ (Graham, Trujillo \\& Caon 2001 and references therein). Nowadays it is clear that the above luminosity--dependent departures from de Vaucouleurs profile are real, as clearly proven by the existence of strong correlations between $n$ (or equivalently galaxy central concentration of light $C_{re}(1/3)$ (Trujillo, Graham \\& Caon 2001)) and the photometric--independent properties, $\\sigma$ (Graham, Trujillo \\& Caon 2001) and galaxy's supermassive black hole mass (Graham et al. 2001b). Aiming at exploring possible correlations between galaxy light profiles and their stellar populations we combine in this letter the results obtained from applying these modeling developments to high quality spectroscopic data. We find a tight correlation between [Mg/Fe] ratio and the shape parameters. \\begin{deluxetable*}{rrrrrrrrrr} \\tablecaption{Galaxy measurements \\label{tbl-1}} \\tablewidth{0pt} \\tablehead{\\colhead{NGC}&\\colhead{n}&\\colhead{Ref}&\\colhead{$C_{re}(1/3)$}&\\colhead{$\\sigma$(${\\rm km~s^{-1}}$)}& \\colhead{Age (Gyr)}& \\colhead{[M/H]}& \\colhead{[Fe/H]}& \\colhead{[Mg/H]}& \\colhead{[Mg/Fe]} } \\startdata \\multicolumn{10}{c}{Virgo galaxies}\\\\[4pt] 4239 & 0.67 & 2 & 0.17 & 82 & 5.5$^{+ 1.9}_{-1.7}$&-0.21$^{+0.11}_{-0.10}$& -0.21$^{+0.09}_{-0.09}$& -0.16$^{+0.15}_{-0.14}$& 0.05$^{+0.10}_{-0.09}$ \\\\[2pt] 4339 & 3.50 & 1 & 0.44 & 142 & 12.6$^{+ 5.3}_{-2.8}$&-0.06$^{+0.10}_{-0.06}$& -0.17$^{+0.07}_{-0.07}$& +0.13$^{+0.15}_{-0.15}$& 0.30$^{+0.11}_{-0.11}$ \\\\[2pt] 4365 & 6.08 & 1 & 0.55 & 245 & 20.4$^{+ 3.1}_{-6.4}$&-0.02$^{+0.10}_{-0.08}$& -0.28$^{+0.16}_{-0.08}$& +0.21$^{+0.21}_{-0.16}$& 0.49$^{+0.08}_{-0.11}$ \\\\[2pt] 4387 & 1.92 & 1 & 0.34 & 105 & 15.1$^{+ 5.4}_{-4.8}$&-0.20$^{+0.11}_{-0.11}$& -0.22$^{+0.10}_{-0.08}$& -0.09$^{+0.19}_{-0.10}$& 0.13$^{+0.12}_{-0.05}$ \\\\[2pt] 4458 & 2.55 & 1 & 0.40 & 105 & 16.6$^{+ 3.9}_{-5.6}$&-0.31$^{+0.13}_{-0.05}$& -0.40$^{+0.05}_{-0.10}$& -0.05$^{+0.16}_{-0.07}$& 0.35$^{+0.14}_{-0.03}$ \\\\[2pt] 4464 & 2.61 & 1 & 0.38 & 135 & 21.0$^{+ 2.3}_{-7.1}$&-0.32$^{+0.12}_{-0.06}$& -0.40$^{+0.09}_{-0.07}$& -0.03$^{+0.12}_{-0.13}$& 0.37$^{+0.06}_{-0.09}$ \\\\[2pt] 4467 & 1.54 & 2 & 0.28 & 75 & 16.0$^{+ 5.2}_{-5.3}$&-0.23$^{+0.13}_{-0.12}$& -0.35$^{+0.13}_{-0.08}$& -0.08$^{+0.18}_{-0.13}$& 0.27$^{+0.08}_{-0.08}$ \\\\[2pt] 4472 & 6.27 & 1 & 0.55 & 306 & 10.4$^{+ 9.5}_{-4.1}$& 0.07$^{+0.17}_{-0.20}$& -0.12$^{+0.15}_{-0.23}$& +0.46$^{+0.16}_{-0.37}$& 0.58$^{+0.04}_{-0.17}$ \\\\[2pt] 4473 & 5.29 & 1 & 0.52 & 180 & 10.3$^{+ 3.1}_{-2.1}$& 0.15$^{+0.07}_{-0.11}$& -0.06$^{+0.06}_{-0.08}$& +0.48$^{+0.08}_{-0.16}$& 0.54$^{+0.04}_{-0.10}$ \\\\[2pt] 4478 & 1.98 & 1 & 0.33 & 135 & 10.0$^{+ 2.7}_{-2.8}$& 0.04$^{+0.15}_{-0.12}$& -0.08$^{+0.12}_{-0.06}$& +0.23$^{+0.14}_{-0.17}$& 0.31$^{+0.05}_{-0.14}$ \\\\[2pt] 4489 & 2.50 & 4 & 0.37 & 73 & 3.1$^{+ 0.6}_{-0.8}$& 0.20$^{+0.18}_{-0.16}$& +0.20$^{+0.15}_{-0.15}$& +0.23$^{+0.17}_{-0.21}$& 0.03$^{+0.06}_{-0.10}$ \\\\[2pt] 4551 & 1.89 & 1 & 0.32 & 105 & 7.2$^{+ 1.5}_{-0.9}$& 0.19$^{+0.07}_{-0.09}$& +0.07$^{+0.05}_{-0.10}$& +0.41$^{+0.07}_{-0.12}$& 0.34$^{+0.05}_{-0.05}$ \\\\[2pt] 4621 & 6.14 & 1 & 0.55 & 230 & 10.3$^{+ 2.4}_{-1.8}$& 0.26$^{+0.08}_{-0.08}$& -0.04$^{+0.04}_{-0.04}$& +0.64$^{+0.05}_{-0.14}$& 0.60$^{+0.02}_{-0.11}$ \\\\[2pt] 4697 & 3.70 & 3 & 0.45 & 168 & 11.3$^{+ 4.5}_{-2.7}$& 0.06$^{+0.10}_{-0.11}$& -0.10$^{+0.06}_{-0.10}$& +0.30$^{+0.08}_{-0.17}$& 0.40$^{+0.04}_{-0.09}$ \\\\[2pt] \\multicolumn{10}{c}{Field galaxies}\\\\[4pt] 584 & 4.70 & 2 & 0.50 & 217 & 8.6$^{+ 1.9}_{-2.3}$& - & -0.03$^{+0.04}_{-0.08}$& +0.31$^{+0.06}_{-0.10}$&+0.34$^{+0.07}_{-0.05}$ \\\\[2pt] 720 & 4.96 & 3 & 0.51 & 247 & 10.9$^{+ 7.0}_{-2.7}$& - & -0.16$^{+0.08}_{-0.08}$& +0.58$^{+0.12}_{-0.19}$&+0.74$^{+0.07}_{-0.14}$ \\\\[2pt] 821 & 4.00 & 2 & 0.47 & 199 & 12.3$^{+12.5}_{-5.6}$& - & -0.15$^{+0.17}_{-0.17}$& +0.26$^{+0.18}_{-0.23}$&+0.41$^{+0.04}_{-0.08}$ \\\\[2pt] 1700 & 5.99 & 2 & 0.54 & 233 & 6.4$^{+ 3.1}_{-1.7}$& - & +0.04$^{+0.08}_{-0.09}$& +0.37$^{+0.09}_{-0.15}$&+0.34$^{+0.09}_{-0.09}$ \\\\[2pt] 3379 & 4.64 & 3 & 0.49 & 201 & 17.8$^{+ 7.4}_{-6.9}$& - & -0.29$^{+0.08}_{-0.10}$& +0.19$^{+0.19}_{-0.10}$&+0.48$^{+0.12}_{-0.03}$ \\\\[2pt] 5831 & 4.72 & 3 & 0.50 & 166 & 5.9$^{+ 3.5}_{-1.9}$& 0.32$^{+0.15}_{-0.13}$& +0.14$^{+0.15}_{-0.16}$& +0.56$^{+0.06}_{-0.18}$&+0.42$^{+0.04}_{-0.10}$ \\\\[2pt] 7454 & 2.55 & 4 & 0.38 & 112 & 6.5$^{+ 3.3}_{-1.8}$& - & -0.20$^{+0.09}_{-0.10}$& -0.07$^{+0.13}_{-0.14}$&+0.13$^{+0.06}_{-0.06}$ \\enddata \\tablecomments{The numbers in column 3 indicate the $n$ parameter sources: 1 Caon, Capaccioli \\& D'Onofrio (1993). 2 HST F555W archive images. 3 Fits to the profiles of Peletier et al. (1990). 4 V--band images from J. Blakeslee (private communication).} \\end{deluxetable*} ", "conclusions": "\\label{Sec:discussion} To test how strong is the new correlation between the shape parameters and [Mg/Fe], we show in Fig.~{\\ref{Fig:c1s}} the fits obtained for the already known $C_{re}(1/3)$ -- $\\sigma$ and [Mg/Fe] -- $\\sigma$ relations. We use the same galaxies plotted in Fig.~{\\ref{Fig:c1a}} and Fig.~{\\ref{Fig:c1m}}. The [Mg/Fe] -- $n$ relation shows a correlation value as strong as the ones plotted in Fig.~{\\ref{Fig:c1s}}. \\begin{figure}[t] \\centerline{\\psfig{file=f3.eps,width=3.6in}} \\caption{Galaxy velocity dispersion versus $n$ and $C_{re}(1/3)$ and [Mg/Fe]. Symbols have same meaning as in Fig.~{\\ref{Fig:c1a}} \\label{Fig:c1s}} \\end{figure} It is well known that early-type galaxies show non-solar abundance ratios (e.g., Trager et al. 2000). The [Mg/Fe] ratio, which spans a super-solar range (see Fig.~{\\ref{Fig:c1m}}), provides us with important clues on the star-formation time-scales. The $\\alpha$ elements, such as Mg, are ejected to the ISM by SN~II, which explode on short time-scales ($\\le10^6-10^7$ yr). On the other hand, SN~Ia are the main producers of Fe peak elements, which require time-scales around $\\sim$1~Gyr. It has been proposed that if the duration of the star formation is shorter than SN~Ia time-scales the [Mg/Fe] should be larger than 0 (e.g., Worthey et al. 1992). Other possibilities have also been proposed, such as a top-heavy initial mass function (IMF), which could also be combined with a star formation that stopped before the products of SN~Ia could be incorporated into the stars (e.g., Worthey et al. 1992; Vazdekis et al. 1996). The fact that the central ($r_{e}/10$) $\\sigma$ and [Mg/Fe] measurements are strongly correlated (see lower panel of Fig.~{\\ref{Fig:c1s}}) suggests that the innermost regions of the more massive galaxies, with larger potential wells, stopped their star formation on shorter time-scales. Furthermore since the light concentration index strongly correlates with [Mg/Fe] suggests that early-type galaxies settled up their global (and not only the innermost regions) structure on time-scales according to their [Mg/Fe] ratios. This scenario would be reinforced if no [Mg/Fe] radial gradients were present. Almost nothing is known about [Mg/Fe] gradients outward galaxy effective radii, there are indications that this ratio is roughly constant at least out to 1~$r_e$ (Halliday 1999; Mehlert et al. 2003). Mehlert et al. suggest that the constraint of short formation time-scales is a global feature and not only applies to galaxy centers. We do not see a clear correlation between $n$ and galaxy mean luminosity weighted ages but [Mg/Fe] suggesting that the formation time-scale, rather than burst peppering as a function of time, is the main process linked to galaxy global structure. Since our galaxies are old and the time scales of the larger galaxies very short, the new correlation suggests that the global structure of these galaxies was already at place at high $z$, without experiencing a significant time evolution. This result is supported by Chiosi \\& Carraro (2002) models, which suggest that the gravitational potential well (and therefore mass and shape of the galaxy) dictates the efficiency and extention of the star formation. The favoured picture is that early-type galaxies are assembled at high $z$ (e.g., Fassano et al. 1998; van Dokkum \\& Ellis 2003), defining a tight sequence in the CM diagram at least out to $z \\sim$1.2 (Blakeslee et al. 2003), and their stellar populations evolve passively (e.g., Im et al. 2002; Schade et al. 1999). Furthermore, Mobasher et al. (2003) find that early-type galaxies show higher rest-frame B-band concentration indices than late-type spirals out to $z \\sim 1$, which suggest that by that redshift value early-type systems have already developed large central light concentrations. Such high light concentrations are predicted by both the monolithic scenario by means of a rapid gas collapse which turns into stars, and within hierarchical galaxy formation framework via mergers of stellar and gaseous systems that take place at high z. Conselice (2003) shows that the light concentration is a key parameter to trace the past evolutionary history of galaxy formation. This research will benefit from studies reaching deeper gradients of line-strengths and abundance ratios, and from theoretical efforts linking them to the light profiles (Angeletti \\& Giannone 2003). Including other element ratios in this analysis would provide further clues, since each element is expelled to the ISM at different epochs, and because line-strengths do not only depend on abundance but on other stellar population parameters. S\\'anchez-Bl\\'azquez et al. (2003) found that the CN is stronger in Virgo than in Coma galaxies. The near-IR Ca{\\sc II} triplet, which is sensitive to both the Ca abundance and the IMF (Vazdekis et al. 2003), decreases as a function of $\\sigma$ (Saglia et al. 2002; Cenarro et al. 2003)." }, "0311/astro-ph0311119_arXiv.txt": { "abstract": "We report on the detection of Main Sequence stars belonging to the recently identified Canis Major galaxy in a field located at $\\simeq 4.2\\degr$ from the center of the stellar system. With Main Sequence fitting we obtain a distance modulus $(m-M)_0=14.5\\pm 0.3$ to the dwarf, corresponding to a distance of $D_{\\sun}\\simeq 8.0 \\pm 1.2$ kpc, in full agreement with previous estimates based on the photometric parallax of M-giants. From the comparison with theoretical isochrones we constrain the age of the main population of the Canis Major system in the range $\\sim 4-10$ Gyr. A blue plume of likely younger stars (age $< 1-2$ Gyr) is also identified. The available Colour Magnitude Diagrams of open clusters that may be projected onto the main body of Canis Major are also briefly analyzed. The position, distance and stellar population of the old open clusters AM-2 and Tombaugh~2 strongly suggest that they are physically associated with the Canis Major galaxy. Using our own photometry and data from 2MASS and the GSC2.2 surveys we demonstrate that the claim by Momany et al. that the CMa overdensity is entirely due to the Galactic warp is not supported by the existing observations, once all the available pieces of information are taken into account. It is shown that the CMa overdensity clearly emerges at a heliocentric distance of $\\sim 8$ kpc above any overdensity possibly produced by the Galactic warp. ", "introduction": "The process of hierarchical merging \\citep{wr78,wf91} is generally accepted as the driving mechanism of the formation of giant galaxies. The study of the local (e.g. Galactic or in the Local Group) relics of such processes provides an unprecedentedly detailed insight of the current cosmological model as well as a formidable testbed for theories of galaxy formation. The case of the Sagittarius dwarf galaxy \\cite[Sgr dSph]{s1,s2} is emblematic, since we are presently observing it during its disruption into a giant stream contributing to the continued assembly of the Galactic halo with stars \\cite[see][and references therein]{ibata01,ibata02,ivez,newberg,majewski} and globular clusters \\citep{bell03a,bell03b,bell03c}. A similar case has been reported in M31 \\citep{m31s1,m31s2} and systematic searches in galaxies outside the Local Group are beginning \\citep{poh}. An even more interesting case is provided by the recently discovered ring structure \\cite[the Monoceros Ring, hereafter the Ring, for brevity][]{newberg,yanny,ibaring,majewski,rochap} near the Galactic plane. The more recent observational results \\citep{crane,martin} suggests that the Ring is the stream remnant of an in-plane accretion that may be contributing to the build-up of the thick disk \\cite[see also][]{helmi,abadi}. In particular \\citet[][hereafter Pap-I]{martin}, while tracing the whole extent of the Ring using M-giants from the 2MASS database\\footnote{See Cutri et al. (2003), Explanatory Supplement to the 2MASS All Sky Data Release, http://www.ipac.caltech.edu/2mass/releases/allsky/doc/explsup.html} discovered a large overdensity of \"Ring-like\" M-giants in the Canis Major constellation. The elliptical shape, the overall structure of this overdensity and its coincidence with a noticeable grouping of globular clusters \\citep{bell03c} led \\citet{martin} to the conclusion that it is the relic of the dwarf galaxy whose disruption generated the Ring. The discovered relic (hereafter Canis Major galaxy, CMa) is approximately centered at galactic coordinates (l,b)$\\sim(240^o,-7^o)$, it has a FWHM extension in the sky of $\\sim 12 \\degr$ in the latitude direction and it is located at $\\sim 7$ kpc from the Sun (i.e., $(m-M)_0\\sim 14.25$), as obtained from the photometric parallax to M-giants introduced by \\citet{majewski}. The mass, luminosity and characteristic dimensions of Canis Major appear quite similar to those of the Sgr dSph. The similarity with Sgr extends to other relevant characteristics: the two objects seem to host a similar grouping of globular clusters, and the upper RGB of both galaxies is dominated by stars of similar colour (M-giants), which are present in similar numbers in both galaxies. The analogy between the two systems may provide a useful guideline in the study of the newly discovered galaxy. \\citet{martin} obtained a near infrared (NIR) Colour Magnitude Diagram (CMD) of the discovered structure (e.g., their Fig.~8) showing a Red Giant Branch (RGB) extending to $(J-K_S)_0\\simeq 1.3$ and a pronounced Red Clump (RC) at $K_{S0}\\ge 12.5$ and $(J-K_S)_0\\simeq 0.6$. A Blue Plume (BP) of stars (of more uncertain nature) is also detected around $(J-K_S)_0\\simeq 0.3$. The Main Sequence (MS) of the CMa population is beyond the reach of 2MASS photometry. The degeneracy among distance, age, metallicity and reddening (which is particularly high and highly variable at such low latitudes) prevents any firm conclusion about the nature of the stellar population in the CMa structure. Furthermore the distance scale based on photometric parallax of M-giants is intrinsically uncertain \\citep{bell03b,majewski} and it is likely affected by a systematic effect leading to the underestimate of true distances \\citep{martin,newberg2}. Hence, the detection of fainter stars belonging to the newly discovered galaxy, the access to alternative distance indicators than M-giants and the study of its stellar population in other wavelength ranges (than the NIR) are highly desirable to confirm the discovery presented in Pap-I and to obtain a clearer characterization of the system. Here we present the first detection of the Main Sequence (MS) of the CMa system in a wide field located $\\sim 4.2\\degr$ apart from its center, from deep B and V photometry of very recent ESO-WFI images taken from the ESO archive. We also take advantage of the fact that the low latitude region of sky covered by the CMa galaxy hosts a number of galactic open clusters (OC), most of them in the foreground. Thus we searched for existing wide field CCD photometry of these open clusters to look for evidence of a background population that may be related to the CMa galaxy. Unfortunately, modern wide field photometries of open clusters are quite rare since, typically, a field of few arcmin is sufficient to characterize the population of these clusters and the background population is an undesired contaminant for these kind of studies. Nevertheless, some interesting clues of the presence of CMa stars in these fields have been obtained, providing useful indications on the spatial extent of the system and guidance for future observations. The list of the fields considered in the current study is reported in Table~1, where we report the position of the fields in Galactic coordinates, the field of view covered by the available observations, the angular distance from the center of CMa, an estimate of the reddening and the source of the data. The position of the various fields with respect to the CMa system is sketched in Fig.~1. Most of the photometry (catalogues in electronic form) used in this paper have been retrieved from the WEBDA\\footnote{\\tt http://obswww.unige.ch/webda/webda.html} database \\citep{mermi}. The adopted reddening values are taken from the maps by \\citet[][SFD98]{schlegel}, if not otherwise stated. All over the paper we adopt the reddening laws by \\citet{dean} for the (Cousins') I band and by \\citet{rl85} for all the other passbands. We anticipate that the data presented in this paper are not sufficient to definitely break the above quoted {\\em distance/age/metallicity/reddening} degeneracy and that reliable spectroscopic abundances and dedicated optical photometry are still needed to obtain a fully consistent picture of the main characteristics of the galaxy. On the other hand, the results presented here should be regarded as a significant progress in the understanding of the CMa system as well as a confirmation of its existence. Criticizing the results presented in \\citet{martin} and in a previous preliminary version of the present paper circulated as a preprint, \\citet[][hereafter M04]{moma4} claimed that the CMa overdensity is completely due to the North-South asymmetry produced by the Galactic warp \\cite[see][and references therein]{yusi,kuij}. We discuss this point in Sect.~4, demonstrating that the M04 hypothesis is clearly confuted by several independent observational evidences. \\begin{table*} \\centering \\begin{minipage}{140mm} \\caption{Explored Fields.} \\begin{tabular}{@{}lrrcccc@{}} \\hline Field Name & l$\\degr$ & b$\\degr$ & f.o.v. & dist\\footnote{Angular distance from the centroid of CMa, in degrees.}&E(B-V)\\footnote{From the reddening maps by \\citet{schlegel}} & Ref. \\\\ \\hline F-XMM & 244.2 & -8.2 & $34\\arcmin\\times 33\\arcmin$ &4.2$\\degr$ &0.14 & This work\\\\ NGC~2477 (B,V) & 253.6 & -5.8 & $34\\arcmin\\times 33\\arcmin$ & 13.6$\\degr$&0.77 & \\citet{moma} \\footnote{The (V,I) photometry by \\citet{kas} has been also considered (f.o.v=$14.7\\arcmin\\times 14.7\\arcmin$).}\\\\ Tombaugh~1& 232.3 & -7.3 & $7.0\\arcmin\\times 7.0\\arcmin$ & 7.7$\\degr$&0.43& \\citet{tom1}\\\\ Berkeley~33& 225.4 & -4.6 & $10.5\\arcmin\\times 10.5\\arcmin$ & 14.8$\\degr$&0.75& \\citet{be33}\\\\ NGC~2204 & 226.0 & -16.1& $14.7\\arcmin\\times 14.7\\arcmin$& 15.9$\\degr$&0.10&\\citet{kas} \\\\ Melotte 66 & 259.6 & -14.3& $14.7\\arcmin\\times 14.7\\arcmin$& 20.2$\\degr$&0.22&\\citet{kas} \\\\ NGC~2243 & 239.5 & -18.0& $15.0\\arcmin\\times 15.0\\arcmin$& 10.5$\\degr$&0.07 & \\citet{2243kal}\\\\ &{\\bf Open Clusters}\\footnote{Possibly associated with CMa, see Sect.~5.}& & & & & \\\\ Arp-Madore 2& 248.1&-5.9 & $5.5\\arcmin\\times 5.5\\arcmin$& 8.3$\\degr$ &0.73&\\citet{am2} \\\\ Tombaugh~2 & 232.9& -6.9& $10.4\\arcmin\\times 10.4\\arcmin$& 6.9$\\degr$&0.39 & \\citet{tom2}\\\\ \\hline \\end{tabular} \\end{minipage} \\end{table*} \\begin{figure} \\includegraphics[width=84mm]{fig1.ps} \\caption{Relative positions on the sky (Galactic coordinates) of the CMa galaxy and of the fields considered in the present analysis. The center of CMa is represented as a cross, the two concentric circles have radius r$=6^o$ deg and r$=12^o$, corresponding to the HWHM and FWHM of the distribution of CMa M-giants as derived in Pap-I, respectively. The clusters that are possible members of the CMa system are represented as open stars. } \\end{figure} \\begin{figure*} \\includegraphics[width=168mm]{fig2.ps} \\caption{CMD of the $\\simeq 0.5\\degr \\times 0.5\\degr$ XMM field, located at $\\simeq 4.2\\degr$ from the center of the CMa system. The error bars at $B-V\\simeq 2.2$ are the average photometric errors as a function of magnitude. The dotted, dashed and continuous lines indicate the locus of the CMD where the joint Completeness Factor (CF) is 80\\%, 50\\% and 5\\%, respectively. The joint CF as a function of V magnitude (computed at $B-V=1.0$) is shown in the inset. } \\end{figure*} ", "conclusions": "We have detected, for the first time, the Main Sequence of the main stellar population of the newly discovered galaxy in Canis Major (Pap-I), from photometry of a wide field located at $\\simeq 4.2\\degr$ from its center (the XMM field). From the analysis of the CMD of the XMM field we conclude that: \\begin{itemize} \\item The stellar population {\\em observed} in the XMM field has no counterpart in the R03 Galactic model. \\item The distance modulus of CMa is $(m-M)_0\\simeq 14.5\\pm 0.3$, where the reported uncertainty mainly reflects the uncertainty in the actual average metallicity of the system, that we assumed to lie in the range $-0.3\\le [M/H]\\le -0.7$, according to \\citet{martin,crane}. The corresponding distance is $D_{\\sun}\\simeq 8.0 \\pm 1.2$ kpc. \\item The age of the main population of CMa lies in the range $4\\sol$ age$\\sol 10$ Gyr, slightly depending on the assumed metallicity. \\item There is a sparse population of younger stars (age$\\sol 1-2$ Gyr) that is also associated with the CMa system (Blue Plume). \\end{itemize} Tentative detections of stars related to the CMa/Ring system are presented for the fields surrounding the open clusters NGC~2477, Tombaugh~1 and Berkeley~33. The cases of two open clusters (AM~2 and Tombaugh~2) that may be physically associated with the CMa system have also been discussed. The results presented in this study provide independent confirmation, full support and complement the results of \\citet{martin}, that described for the first time the CMa stellar system. The claim by \\citet{moma4} that the CMa overdensity could be entirely due to the South/North asymmetry produced by the Galactic warp has been demonstrated to lack the essential observational basis. Using the F-XMM data we have shown that the warp model encapsulated in the R03 model (used by M04 to substantiate their claim) fails to reproduce both the star counts {\\em and} CMD morphology observed in this field, located at $\\simeq 4.2\\degr$ from the center of the CMa system. Using Red Clump and M giant stars from 2MASS we have shown that (a) the CMa overdensity is much larger than that produced by the warp, and (b) it is found only at a well defined distance not compatible with the warp, fully confirming the results of Pap-I and demonstrating that if all the available information is taken into account the CMa overdensity clearly emerges {\\em above} the ``noise'' provided by the Galactic warp. Using GSC2.2 data we demonstrated that the Blue Plume population is a distinctive characteristic of the CMa system that has no comparable counterpart neither in Northern Control Fields or in fields sampling the real warp. A better characterization of the stellar content and star formation history of the CMa system must wait for dedicated photometry and spectroscopy fully suited for the study of this difficult target, immersed in very contaminated and extincted regions of the sky. Here we tried to obtain the best possible characterization of the system from existing and publicly available data, as an immediate follow-up of the discovery. Future searches for CMa stars in the background of other targets may prove extremely useful in characterizing the spatial distribution of the galaxy and, in particular, of the associated Ring. The relevant change of view is that what may be perceived as an undesired contaminant of a given observation may hide more astrophysical information than the target itself. This is true either if the background object is an unknown stellar system (as CMa or Sgr) either if it is a poorly characterized Galactic component, as the stellar warp of the Milky Way \\citep{binney,kuij}. The lessons provided by the Sgr dSph (and Stream) and by the CMa galaxy (and Ring) suggests to all of us a renewed attitude in the analysis of astronomical data, to avoid missing relevant pieces of information hidden in the background of our main targets." }, "0311/astro-ph0311355_arXiv.txt": { "abstract": "The mechanisms which could lead to chemo-thermal instabilities and fragmentation during the formation of primordial protostars are investigated analytically. We introduce new analytic approximations for \\HH cooling rates bridging the optically thin and thick regimes. These allow us to discuss chemo--thermal instabilities up to densities when protostars become optically thick to continuum radiation ($n\\equiv\\rho/m_H\\lsim10^{16-17}\\;{\\rm cm^{-3}}$). During the proto-stellar collapse instabilities are active in two different density regimes. In the well known ``low density'' regime ($n\\sim10^8-10^{10}\\;{\\rm cm^{-3}}$), instability is due to 3-body reactions quickly converting atomic hydrogen into H$_2$. In the ``high density'' regime ($n\\gsim 10^{14}\\;{\\rm cm^{-3}}$), another instability is triggered by the strong increase in the cooling rate due to \\HH Collisional Induced Emission (CIE). In agreement with the three dimensional simulations, we find that the ``low density'' instabilities cannot lead to fragmentation, both because fluctuations are too small to survive turbulent mixing, and because their growth times are too slow. The situation for the newly found ``high density'' instability is analytically similar. This continuum cooling instability is as weak as ``low density'' instability, with similar ratios of growth and dynamical time scales, as well as allowing for the necessary fragmentation condition $t_{cool}\\lsim t_{dyn}$. Because the instability growth timescale is always longer than the free fall timescale, it seems unlikely that fragmentation could occur in this high density regime. Consequently, one expects the first stars to be very massive, not to form binaries nor harbour planets. Nevertheless, full three dimensional simulations are required to be certain. Such 3D calculations could become possible using simplified approaches to approximate the effects of radiative transfer, which we show to work very well in 1D calculations, giving virtually indistinguishable results from calculations employing full line transfer. This indicates that the effects of radiative transfer during the initial stages of formation of primordial proto--stars are local corrections to cooling rather than influencing the energetics of distant regions of the flow. ", "introduction": "In current scenarios for the formation of the first stars, molecular hydrogen plays a prominent role, providing the only cooling mechanism for metal-free gas at $T\\lsim 10^4\\;{\\rm K}$. \\HH cooling is responsible for both the collapse of the gas inside the small ($M_{\\rm halo} \\sim 10^6\\;\\msun$) cosmological halos where primordial star formation is believed to occur (\\eg Couchman \\& Rees 1986; Tegmark \\etal 1997, Abel \\etal 1998), and for the fragmentation of the gas itself into ``clumps'' with masses of $\\sim 100-1000\\;\\msun$, as is seen in the simulations of Abel, Bryan \\& Norman (1998, 2000) (see also Bromm, Coppi \\& Larson 2002). The fate of these objects is unclear. Several previous studies, based on analytical arguments (\\ie stability analysis) or single-zone models, led to different conclusions about the fragmentation properties of primordial clouds. One important instability-triggering mechanism was first suggested by Palla, Salpeter \\& Stahler (1983; hereafter PSS83): the onset of 3-body \\HH formation at number densities $n\\gsim 10^8\\;{\\rm cm^{-3}}$ (where $n\\equiv\\rho/m_{\\rm H}$) causes a fast increase in the cooling rate, possibly leading to fragmentation on a mass scale of $\\sim 0.1\\msun$. A similar result was obtained by Silk (1983; hereafter S83) by applying a criterion for ``chemo-thermal instability'' (first discussed by Sabano \\& Yoshii, 1977) to an object where 3-body \\HH formation is important. Numerical simulations have recently shed new light upon the issue, and in particular Abel, Bryan \\& Norman (2002; hereafter ABN02) were able to reach the 3-body reactions density regime finding that these reactions actually lead to the formation of a single fully molecular core (with a mass of $\\sim 1\\;\\msun$) at the centre of each clump. They point out that in their simulations they see several chemo-thermally unstable regions, but that the instabilities do not lead to fragmentation because turbulence efficiently mixes the gas, erasing each fluctuation before it can grow significantly. Recently, Omukai \\& Yoshii (2003; OY03) found a somewhat different explanation by improving the original S83 investigation and pointing out that instabilities growth was too slow. The further evolution of such an object is still uncertain: due to the high molecular fraction, many of the numerous \\HH roto-vibrational lines (accounting for most of the cooling) become optically thick just after the formation of the molecular core. For this reason, it becomes impossible to estimate the \\HH cooling rate without a complete treatment of radiative transfer, which currently can not be included into full 3-D hydrodynamical simulations. Simpler 1-D studies, such as those in Omukai \\& Nishi (1998; hereafter ON98) and Ripamonti \\etal (2002; hereafter R02), can follow the further evolution, giving useful predictions about the physical processes driving the collapse, or about the properties of the small hydrostatic protostellar core which finally forms in the centre; on the other hand, their intrinsic spherical symmetry prevents them from giving direct predictions about fragmentation. An interesting result of ON98 and R02 is that during the collapse, the protostellar object experiences a phase when cooling is dominated by the effects of \\HH Collision-Induced Emission (CIE), the opposite process of the more commonly known Collision-Induced Absorption, or CIA. Both of them find that, at the beginning of the CIE-cooling phase, the molecular core is optically thin to CIE-produced continuum photons, and there is a very fast increase in the cooling rate until the continuum optical depth exceeds unity. However, neither of these previous studies pointed out that during this phase the core conditions resemble the ones that, at the onset of 3-body \\HH formation, led to the chemo-thermal instability: the core is optically thin (which is believed to be a necessary condition for fragmentation; see Rees 1976) and the cooling rate is undergoing a dramatic increase. In the next sections, we will investigate more thoroughly the issue of chemo-thermal instability, with special attention to the effects of CIE cooling. In section 2 we will describe the main cooling processes, giving a brief account of CIE emission, and some useful approximations for \\HH line cooling. In section 3 we examine the conditions for the insurgence of chemo-thermal instabilities, and argue whether they can cause fragmentation. Finally, we discuss the results in section 4. ", "conclusions": "We have investigated the chemical and thermal instabilities of primordial metal-free gas during its collapse towards the formation of a protostar. We have focused on the Collision Induced Emission (CIE) and on its effects on fragmentation. We point out that even when \\HH line cooling becomes optically thick at densities $n\\gsim10^{10}\\;{\\rm cm^{-3}}$ this does not necessarily set a minimum mass scale for fragmentation due to an opacity limit. At higher densities ($n\\gsim10^{14}\\;{\\rm cm^{-3}}$) the optically thin continuum (CIE) cooling dominates, and the collapsing protostellar cloud can once again fulfill the necessary (but not sufficient) fragmentation criterion $t_{\\rm cool}